0% found this document useful (0 votes)
9 views21 pages

PH4401 03 Entanglement

Chapter 3 of PH4401 discusses quantum entanglement and the quantum states of multi-particle systems, emphasizing how the postulates of quantum mechanics lead to phenomena such as entangled states that cannot be expressed as definite states of individual particles. It also covers the implications of partial measurements on multi-particle systems and introduces the EPR paradox, which illustrates the counterintuitive nature of quantum mechanics and the instantaneous effect of measurements on entangled particles, regardless of distance. The chapter highlights the complexities of quantum states and the significant implications for quantum computing and the understanding of reality.

Uploaded by

Abraham Cano
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
9 views21 pages

PH4401 03 Entanglement

Chapter 3 of PH4401 discusses quantum entanglement and the quantum states of multi-particle systems, emphasizing how the postulates of quantum mechanics lead to phenomena such as entangled states that cannot be expressed as definite states of individual particles. It also covers the implications of partial measurements on multi-particle systems and introduces the EPR paradox, which illustrates the counterintuitive nature of quantum mechanics and the instantaneous effect of measurements on entangled particles, regardless of distance. The chapter highlights the complexities of quantum states and the significant implications for quantum computing and the understanding of reality.

Uploaded by

Abraham Cano
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 21

Y. D.

Chong PH4401: Quantum Mechanics III

Chapter 3: Quantum Entanglement

They don’t think it be like it is,


but it do.

Oscar Gamble

I. QUANTUM STATES OF MULTI-PARTICLE SYSTEMS

So far, we have studied quantum mechanical systems consisting of single particles. The
next important step is to look at systems of more than one particle. We shall see that
the postulates of quantum mechanics, when applied to multi-particle systems, give rise to
interesting and counterintuitive phenomena such as quantum entanglement.
Suppose we have two particles labeled A and B. If each individual particle is treated as
a quantum system, the postulates of quantum mechanics require that its state be described
by a vector in a Hilbert space. Let HA and HB denote the respective single-particle Hilbert
spaces. Then the Hilbert space for the combined system of two particles is
H = HA ⊗ HB . (3.1)
The symbol ⊗ refers to a tensor product, a mathematical operation that combines two
Hilbert spaces to form another Hilbert space. It is most easily understood in terms of explicit
basis vectors: let HA be spanned by a basis {|µ1 i, |µ2 i, |µ3 i, . . . }, and HB be spanned by
{|ν1 i, |ν2 i, |ν3 i, . . . }. Then HA ⊗ HB is a space spanned by basis vectors consisting of
pairwise combinations of basis vectors drawn from the HA and HB bases:
n o
|µi i ⊗ |νj i for all |µi i, |νj i . (3.2)

Thus, if HA has dimension dA and HB has dimension dB , then HA ⊗ HB has dimension


dA dB . Any two-particle state can be written as a superposition of these basis vectors:
X
|ψi = cij |µi i ⊗ |νj i. (3.3)
ij

The inner product between the tensor product basis states is defined as follows:
    
|µi i ⊗ |νj i , |µp i ⊗ |νq i ≡ hµi | ⊗ hνj | |µp i ⊗ |νq i ≡ hµi |µp i hνj |νq i = δip δjq . (3.4)

In other words, the inner product is performed “slot-by-slot”. We calculate the inner product
for A, calculate the inner product for B, and then multiply the two resulting numbers. You
can check that this satisfies all the formal requirements for an inner product in linear algebra
(see Exercise 1).
For example, suppose HA and HB are both 2D Hilbert spaces describing spin-1/2 degrees
of freedom. Each space can be spanned by an orthonormal basis { |+zi, |−zi }, representing
“spin-up” and “spin-down”. Then the tensor product space H is a 4D space spanned by
n o
|+zi ⊗ |+zi , |+zi ⊗ |−zi , |− zi ⊗ |+zi , |−zi ⊗ |−zi . (3.5)

37
Y. D. Chong PH4401: Quantum Mechanics III

We now make an important observation. If A is in state |µi and B is in state |νi, then
the state of the combined system is fully specified: |µi ⊗ |νi ∈ HA ⊗ HB . But the reverse
is not generally true! There exist states of the combined system that cannot be expressed
in terms of definite states of the individual particles. For example, consider the following
quantum state of two spin-1/2 particles:
1  
|ψi = √ |+zi ⊗ |−zi − |−zi ⊗ |+zi . (3.6)
2
This state is constructed
√ from two of the four basis states in (3.5), and you can check that
the factor of 1/ 2 ensures the normalization hψ|ψi = 1 with the inner product rule (3.4).
It is evident from looking at Eq. (3.6) that neither A nor B possesses a definite |+zi or |−zi
state. Moreover, we shall show (in Section VII) that there’s no choice of basis that allows
this state to be expressed in terms of definite individual-particle states; i.e.,

6 |ψA i ⊗ |ψB i for any |ψA i ∈ HA , |ψB i ∈ HB .


|ψi = (3.7)

In such a situation, the two particles are said to be entangled.


It is cumbersome to keep writing ⊗ symbols, so we will henceforth omit the ⊗ in cases
where the tensor product is obvious. For instance,
1   1  
√ |+zi ⊗ |−zi − |−zi ⊗ |+zi ≡ √ |+zi|−zi − |−zi|+zi . (3.8)
2 2

For systems of more than two particles, quantum states can be defined using multiple
tensor products. Suppose a quantum system contains N particles described by the individual
Hilbert spaces {H1 , H2 , . . . , HN } having dimensionality {d1 , . . . , dN }. Then the overall
system is described by the Hilbert space

H = H1 ⊗ H2 ⊗ · · · ⊗ HN , (3.9)

which has dimensionality d = d1 d2 · · · dN . The dimensionality scales exponentially with the


number of particles! For instance, if each particle has a 2D Hilbert space, a 20-particle system
has a Hilbert space with 220 = 1 048 576 dimensions. Thus, even in quantum systems with
a modest number of particles, the quantum state can carry huge amounts of information.
This is one of the motivations behind the active research field of quantum computing.
Finally, a proviso: although we refer to subsystems like A and B as “particles” for nar-
rative convenience, they need not be actual particles. All this formalism applies to general
subsystems—i.e., subsets of a large quantum system’s degrees of freedom. For instance, if
a quantum system has a position eigenbasis for 3D space, the x, y, and z coordinates are
distinct degrees of freedom, so each position eigenstate is really a tensor product:

| r = (x, y, z) i ≡ |xi |yi |zi.

Also, if the subsystems really are particles, we are going to assume for now that the particles
are distinguishable. There are other complications that arise if the particles are “identical”,
which will be the subject of the next chapter (if you’re unsure what this means, just read
on).

38
Y. D. Chong PH4401: Quantum Mechanics III

II. PARTIAL MEASUREMENTS

Let us recall how measurements work in single-particle quantum theory. Each observable
Q is described by some Hermitian operator Q̂, which has an eigenbasis {|qi i} such that

Q̂|qi i = qi |qi i. (3.10)


For simplicity, let the eigenvalues {qi } be non-degenerate. Suppose a particle initially has
quantum state |ψi. This can always be expanded in terms of the eigenbasis of Q̂:
X
|ψi = ψi |qi i, where and ψi = hqi |ψi. (3.11)
i

The measurement postulate of quantum mechanics states that if we measure Q, then


(i) the probability of obtaining the measurement outcome qi is Pi = |ψi |2 , the absolute
square of the coefficient of |qi i in the basis expansion; and (ii) upon obtaining this outcome,
the system instantly “collapses” into state |qi i.
Mathematically, these two rules can be summarized using the projection operator
Π̂(qi ) = |qi ihqi |. (3.12)
Applying this operator to |ψi gives the non-normalized state vector
|ψ 0 i = |qi ihqi |ψi. (3.13)
From this, we glean two pieces of information:

• The probability of obtaining this outcome is hψ 0 |ψ 0 i = |hqi |ψi|2 .


• The post-collapse state is obtained by the re-normalization |ψ 0 i → |qi i.

For multi-particle systems, there is a new complication: what if a measurement is per-


formed on just one particle?
Consider a system of two particles A and B, with two-particle Hilbert space HA ⊗HB . We
perform a measurement on particle A, corresponding to a Hermitian operator Q̂A that acts
upon HA and has eigenvectors {|µi | µ = 1, 2, . . . } (i.e., the eigenvectors are enumerated by
some index µ). We can write any state |ψi using the eigenbasis of Q̂A for the HA part, and
an arbitrary basis {|νi} for the HB part:
X
|ψi = ψµν |µi|νi
µν
X X (3.14)
= |µi|ϕµ i, where |ϕµ i ≡ ψµν |νi ∈ HB .
µ ν

Unlike the single-particle case, the “coefficient” of |µi i in this basis expansion is not a
complex number, but a vector in HB .
Proceeding by analogy, the probability of obtaining the outcome labelled by µ should be
the “absolute square” of this “coefficient”, hϕµ |ϕµ i. Let us define the partial projector
ˆ
Π̂(µ) = |µihµ| ⊗ I. (3.15)

39
Y. D. Chong PH4401: Quantum Mechanics III

The A slot of this operator contains a projector, |µihµ|, while the B slot leaves the HB part
of the two-particle space unchanged. Applying the partial projector to the state given in
Eq. (3.14) gives
|ψ 0 i = Π̂(µ) |ψi = |µi|ϕµ i. (3.16)
Now we follow the same measurement rules as before. The outcome probability is
X
Pµ = hψ 0 |ψ 0 i = hµ|µi hϕµ |ϕµ i = |ψµν |2 . (3.17)
ν

The post-measurement collapsed state is obtained by the re-normalization


1 X
|ψ 0 i → pP ψµν |µi|νi. (3.18)
2
ν 0 |ψµν 0 | ν

Example—A system of two spin-1/2 particles is in the “singlet state”


1  
|ψi = √ |+zi|−zi − |−zi|+zi . (3.19)
2

For each particle, |+z i and |−z i denote eigenstates of the operator Ŝz , with eigenvalues
+~/2 and −~/2 respectively. Suppose we measure Sz on particle A. What are the
probabilities of the possible outcomes, and the associated post-collapse states?

• First outcome: +~/2.


ˆ
– The partial projector is |+zih+z| ⊗ I.

– Applying the projection to |ψi yields |ψ 0 i = (1/ 2) |+zi|−zi.
1
– The outcome probability is P+ = hψ 0 |ψ 0 i = .
2
1
– The post-collapse state is √ |ψ 0 i = |+zi|−zi
P+
• Second outcome: −~/2.
ˆ
– The partial projector is |−zih−z| ⊗ I.

– Applying the projection to |ψi yields |ψ 0 i = (1/ 2) |−zi|+zi.
1
– The outcome probability is P− = hψ 0 |ψ 0 i = .
2
1
– The post-collapse state is = √ |ψ 0 i = |−zi|+zi.
P−

The two possible outcomes, +~/2 and −~/2, occur with equal probability. In either
case, the two-particle state collapses so that A is in the observed spin eigenstate, and B
has the opposite spin. After the collapse, the two-particle state is no longer entangled.

40
Y. D. Chong PH4401: Quantum Mechanics III

III. THE EINSTEIN-PODOLSKY-ROSEN “PARADOX”

In 1935, Einstein, Podolsky, and Rosen (EPR) formulated a thought experiment, now
known as the EPR paradox, that highlights the counter-intuitive features of quantum
entanglement [3]. They tried to use this thought experiment to argue that quantum theory
cannot serve as a fundamental description of reality. Subsequently, however, it was shown
that the EPR paradox is not an actual paradox; physical systems really do have the strange
behavior that the thought experiment highlighted.
Consider an entangled state, like the following “singlet state” of two spin-1/2 particles:
1  
|ψi = √ |+zi|−zi − |−zi|+zi . (3.20)
2
As before, let the two particles be labeled A and B. Measuring Sz on A collapses the system
into a two-particle state that is unentangled, where each particle has a definite spin. If the
measurement outcome is +~/2, the new state is |+zi|−zi, whereas if the outcome is −~/2,
the new state is |−zi|+zi.
The postulates of quantum theory seem to indicate that the state collapse happens in-
stantaneously, regardless of the distance separating the particles. Imagine that we prepare
the two-particle state in a laboratory on Earth. Particle A is then transported to the lab-
oratory of Alice, in the Alpha Centauri star system, and particle B is transported to the
laboratory of Bob, in the Betelgeuse system, separated by ∼ 640 light years. In principle,
this can be done carefully enough to avoid disturbing the two-particle quantum state.

Once ready, Alice measures Ŝz on particle A, which induces an instantaneous collapse
of the two-particle state. Immediately afterwards, Bob measures Ŝz on particle B, and
obtains—with 100% certainty—the opposite spin. During the time interval between these
two measurements, no classical signal could have traveled between the two star systems,
not even at the speed of light. Yet the state collapse induced by Alice’s measurement has a
definite effect on the result of Bob’s measurement.
There are three noteworthy aspects of this phenomenon:
First, it dispels some commonsensical but mistaken “explanations” for quantum state
collapse in terms of perturbative effects. For instance, it is sometimes explained that if we
want to measure a particle’s position, we need to shine a light beam on it, or disturb it
in some way, and this disturbance generates an uncertainty in the particle’s momentum.
The EPR paradox shows that such stories don’t capture the full weirdness of quantum state
collapse, for we can collapse the state of a particle by doing a measurement on another
particle far away!

41
Y. D. Chong PH4401: Quantum Mechanics III

Second, our experimentalists have a certain amount of control over the state collapse, due
to the choice of what measurement to perform. So far, we have considered Sz measurements
performed by Alice on particle A. But Alice can choose to measure the spin of A along
another axis, say Sx . In the basis of spin-up and spin-down states, the operator Ŝx has
matrix representation  
~ 0 1
Ŝx = . (3.21)
2 1 0
The eigenvalues and eigenvectors are
~ 1  
sx = , |+xi = √ |+zi + |−zi
2 2
(3.22)
~ 1  
sx = − , |−xi = √ |+zi − |−zi .
2 2
Conversely, we can write the Ŝz eigenstates in the {|+xi, |−xi} basis:
1  
|+zi = √ |+xi + |−xi
2
(3.23)
1  
|−zi = √ |+xi − |−xi .
2
This allows us to write the two-particle entangled state in the Ŝx basis:
1  
|ψi = √ |−xi|+xi − |+xi|−xi . (3.24)
2
Alice’s measurement still collapses the particles into definite spin states with opposite spins—
but now spin states of Sx rather than Sz .
Third, this ability to choose the measurement axis does not allow for superluminal com-
munication. Alice can choose whether to (i) measure Sz or (ii) measure Sx , and this choice
instantaneously affects the quantum state of particle B. If Bob can find a way to distinguish
between the cases (i) and (ii), even statistically, this would serve as a method for instanta-
neous communication, violating the theory of relativity! Yet this turns out to be impossible.
The key problem is that quantum states themselves cannot be measured; only observables
can be measured. Suppose Alice’s measurement is Ŝz , which collapses B to either |+zi or
|−zi, each with probability 1/2. Bob must now choose which measurement to perform. If
he measures Sz , the outcome is +~/2 or −~/2 with equal probabilities. If he measures Sx ,
the probabilities are:
1 2 1 2 1
P (Sx = +~/2) = h+x|+zi + h+x|−zi =
2 2 2 (3.25)
1 2 1 2 1
P (Sx = −~/2) = h−x|+zi + h−x|−zi = .
2 2 2
The probabilities are still equal! Repeating this analysis for any other choice of spin axis, we
find that the two possible outcomes always have equal probability. Thus, Bob’s measurement
does not yield any information about Alice’s choice of measurement axis.
Since quantum state collapse does not allow for superluminal communication, it is con-
sistent in practice with the theory of relativity. However, state collapse is still nonlocal,

42
Y. D. Chong PH4401: Quantum Mechanics III

in the sense that unobservable ingredients of the theory (quantum states) can change faster
than light can travel between two points. For this reason, EPR argued that quantum theory
is philosophically inconsistent with relativity.
EPR suggested an alternative: maybe quantum mechanics is an approximation of some
deeper theory, whose details are currently unknown, but which is deterministic and local.
Such a “hidden variable theory” may give the appearance of quantum state collapse in
the following way. Suppose each particle has a definite but “hidden” value of Sz , either
Sz = +~/2 or Sz = −~/2; let us denote these as [+] or [−]. We can hypothesize that the
two-particle quantum state |ψi is not an actual description of reality; rather, it corresponds
to a statistical distribution of “hidden variable” states, denoted by [+; −] (i.e., Sz = +~/2
for particle A and Sz = −~/2 for particle B), and [−; +] (the other way around).

When Alice measures Sz , the value of the hidden variable is revealed. A result of +z
implies [+; −], whereas −z implies [−; +]. When bob subsequently measures Sz , the result
obtained is the opposite of Alice’s result. But those were simply the values all along—there
is no instantaneous physical influence traveling between their two laboratories.
Clearly, there are many missing details in this hypothetical description. Any actual hidden
variable theory would also need to replicate the huge list of successful predictions made by
quantum theory. Trying to come up with a suitable theory of this sort seems difficult, but
with enough hard work, one might imagine that it is doable.

IV. BELL’S THEOREM

In 1964, John S. Bell published a bombshell paper showing that the predictions of quan-
tum theory are inherently inconsistent with hidden variable theories [4]. The amazing thing
about this result, known as Bell’s theorem, is that it requires no knowledge about the
details of the hidden variable theory, just that it is deterministic and local. Here, we present
a simplified version of Bell’s theorem due to Mermin [5].
We again consider spin-1/2 particle pairs, with particle A sent to Alice at Alpha Centauri,
and particle B to Bob at Betelgeuse. Each experimentalist can measure the particle’s spin
along three distinct choices of spin axis. These spin observables are denoted by S1 , S2 , and
S3 . We will not specify the actual directions of these spin axes until later in the proof. For
now, just note that the axes need not correspond to orthogonal spatial directions.
We repeatedly prepare the particle pairs in the singlet state
1  
|ψi = √ |+zi|−zi − |−zi|+zi , (3.26)
2

43
Y. D. Chong PH4401: Quantum Mechanics III

and send the respective particles to Alice and Bob. During each round of the experiment,
each experimentalist randomly chooses one of the three spin axes S1 , S2 , or S3 , and performs
that spin measurement. It doesn’t matter which experimentalist performs the measurement
first; the experimentalists can’t influence each other, as there is not enough time for a
light-speed signal to travel between the two locations. Many rounds of the experiment are
conducted; for each round, both experimentalists’ choices of spin axis are recorded, along
with their measurement results.

At the end, the experimental records are brought together and examined. We assume
that the results are consistent with the predictions of quantum theory. Among other things,
this means that whenever the experimentalists happen to choose the same measurement
axis, they always find opposite spins. (For example, this is the case during “Experiment 4”
in the above figure, where both experimentalists happened to measure S3 .)
Can a hidden variable theory reproduce the results predicted by quantum theory? In a
hidden variable theory, each particle must have a definite value for each spin observable. For
example, particle A might have S1 = +~/2, S2 = +~/2, S3 = −~/2. Let us denote this by
[+ + −]. To be consistent with the predictions of quantum theory, the hidden spin variables
for the two particles must have opposite values along each direction. This means that there
are 8 distinct possibilities, which we can denote as
[+ + +; − − −], [+ + −; − − +], [+ − +; − + −], [+ − −; − + +],
[− + +; + − −], [− + −; + − +], [− − +; + + −], [− − −; + + +].
For instance, [+ + −; − − +] indicates that for particle A, S1 = S2 = +~/2 and S3 = −~/2,
while particle B has the opposite spin values, S1 = S2 = −~/2 and S3 = +~/2. So far,
however, we don’t know anything about the relative probabilities of these 8 cases.
Let’s now focus on the subset of experiments in which the two experimentalists happened
to choose different spin axes (e.g., Alice chose S1 and Bob chose S2 ). Within this subset,
what is the probability for the two measurement results to have opposite signs (i.e., one +
and one −)? To answer this question, we first look at the following 6 cases:
[+ + −; − − +], [+ − +; − + −], [+ − −; − + +],
[− + +; + − −], [− + −; + − +], [− − +; + + −].
These are the cases which do not have all + or all − for each particle. Consider one of these,
say [+ + −; − − +]. The two experimentalists picked their measurement axes at random

44
Y. D. Chong PH4401: Quantum Mechanics III

each time, and amongst the experiments where they picked different axes, there are two
ways for the measurement results to have opposite signs: (S1 , S2 ) or (S2 , S1 ). There are four
ways to get the same sign: (S1 , S3 ), (S2 , S3 ), (S3 , S1 ) and (S3 , S2 ). Thus, for this particular
set of hidden variables, the probability for measurement results with opposite signs is 1/3.
If we go through all 6 of the cases listed above, we find that in call cases, the probability for
opposite signs is 1/3.
Now look at the remaining 2 cases:

[+ + +; − − −], [− − −; + + +].

For these, Alice and Bob always obtain results with opposite signs. Combining this with
the findings from the previous paragraph, we obtain the following statement:
Given that the two experimentalists choose different spin axes, the probability that their
results have opposite signs is P ≥ 1/3.
This is called Bell’s inequality. If we can arrange a situation where quantum theory
predicts a probability P < 1/3 (i.e., a violation of Bell’s inequality), that would mean that
quantum theory is inherently inconsistent with local deterministic hidden variables. This
conclusion would hold regardless of the “inner workings” of the hidden variable theory. In
particular, note that the above derivation made no assumptions about the relative proba-
bilities of the hidden variable states.
To complete the proof, we must find a set {S1 , S2 , S3 } such that the predictions of quantum
mechanics violate Bell’s inequality. One simple choice is to align S1 with the z axis, and
align S2 and S3 along the x-z plane at 120◦ (2π/3 radians) from S1 , as shown below:

The corresponding spin operators can be written in the eigenbasis of Ŝz :


~
Ŝ1 = σ3
2
~
Ŝ2 = [cos(2π/3)σ3 + sin(2π/3)σ1 ] (3.27)
2
~
Ŝ3 = [cos(2π/3)σ3 − sin(2π/3)σ1 ] .
2

Suppose Alice chooses S1 , and obtains +~/2. Particle A collapses to state | + zi, and
particle B collapses to state |−zi. Bob is assumed to choose a different spin axis. If the
choice is S2 , the expectation value is
~h i
h −z | S2 | −z i = cos(2π/3)h −z |σ3 | −z i + sin(2π/3)h −z |σ1 | −z i
2 (3.28)
~ 1
= ·
2 2

45
Y. D. Chong PH4401: Quantum Mechanics III

If P+ and P− respectively denote the probability of measuring +~/2 and −~/2 in this
measurement, the above equation implies that P+ − P− = +1/2. Moreover, P+ + P− = 1 by
probability conservation. Hence, the probability of obtaining a negative value (the opposite
sign from Alice’s measurement) is P− = 1/4. All the other possible scenarios are worked out
similarly. The result is that the overall probability of the two experimentalists obtaining
opposite results (in the cases where they choose different measurement axis) is 1/4. Bell’s
inequality is violated!
Last of all, we must consult Nature itself. Is it possible to observe, in an actual ex-
periment, probabilities that violate Bell’s inequality? In the decades following Bell’s 1964
paper, many experiments were performed to answer this question. These experiments are
all substantially more complicated than the simple two-particle spin-1/2 model that we’ve
studied, and they are subject to various uncertainties and “loopholes” that are beyond the
scope of our discussion. But in the end, the experimental consensus appears to be a clear
yes: Nature really does behave according to quantum mechanics, and in a manner that can-
not be replicated by deterministic local hidden variables! A summary of the experimental
evidence is given in a review paper by Aspect [6].

V. QUANTUM CRYPTOGRAPHY

One of the most remarkable consequences of Bell’s thought experiment is that it provides
a way to perform cryptography that is more secure, in certain respects, than conventional
cryptography. This possibility was first raised by Ekert [7], and it has led to a huge amount
of research into quantum cryptography, which is poised to be one of the most important
technological applications of quantum mechanics.
Ekert’s quantum cryptography scheme allows two participants, Alice and Bob, to share
with each other a string of random binary digits (0 or 1), called a “key”, in such a manner
that no one else can learn the key by eavesdropping on their communications. Once Alice
and Bob have established a secret shared key, it can be used to encrypt subsequent messages
between them, which nobody else can decipher (e.g., by using one-time pads).
The scheme follows almost immediately from the Bell thought experiment of Section IV.
In each round, a pair of spin-1/2 particles is prepared in the singlet state, with particle A
sent to Alice and B sent to Bob. Alice and Bob each randomly choose a measurement axis
(S1 , S2 , or S3 ), and measure the spin of their particle along that axis.
After an appropriate number of rounds, Alice and Bob publicly announce their choices
of measurement axes. These announcements are assumed to take place over a classical
communication channel that cannot be jammed or manipulated by any hostile party (though
it can be eavesdropped upon). From the announcements, Alice and Bob determine the
rounds in which they happened to pick the same axes. Their measurement results during
these rounds are guaranteed to be the opposites of each other. Hence, they have established
a random binary string known to each other but to no one else.
How might an eavesdropper, Eve, attempt to foil this scheme? Suppose Eve can intercept
some or all of the particles B destined for Bob. She might try to substitute her own
measurements, in a manner that could let her work out the secret key. However, Eve is
hampered by the fact that she is unable to predict or influence Bob’s choices of measurement

46
Y. D. Chong PH4401: Quantum Mechanics III

axes (i.e., Bob’s choices are truly random), nor is she able to impersonate Bob during the
announcements of the axis choices (i.e., the classical communication channel is unjammable).
Under these assumptions, it can be shown that any attempt by Eve to substitute her own
measurements can be detected by Alice and Bob, by performing a statistical analysis of
their measurement results in the rounds with different different axis choices. The detection
of the eavesdropper turns out to be essentially the same as checking for Bell’s inequality.
For details, refer to Ref. [7].
Alternatively, Eve might try to “clone” the quantum state of particle B before passing it
along to Bob. If this can be done, Eve can retain the cloned quantum state, wait for Bob to
announce his choice of measurement axis for that round, and then perform the corresponding
measurement to reproduce Bob’s result. Though plausible at first glance, this turns out to
be fundamentally unworkable, as it is incompatible with the laws of quantum mechanics.
The so-called no-cloning theorem can be proven as follows. Eve desires to clone an
arbitrary state of a spin-half particle B onto another spin-half particle C. The two-particle
Hilbert space is H ⊗ H . With particle C initially prepared in some state |0i, Eve must
devise a unitary operation Û , representing the cloning process, such that

Û |ψi|0i = eiφ |ψi|ψi (3.29)

for all |ψi ∈ H , and for some phase factor φ that could depend on |ψi. Note that the value
of φ does not affect the outcomes of measurements.
Now replace |ψi in the above equation with two arbitrary states denoted by |ψ1 i and
|ψ2 i, and take their inner product. According to Eq. (3.29),
     
hψ1 |h0|Û † Û |ψ2 i|0i = hψ1 |hψ1 |e−iφ1 eiφ2 |ψ2 i|ψ2 i (3.30)
 2
= e−i(φ1 −φ2 ) hψ1 |ψ2 i . (3.31)

Here, φ1 and φ2 are the phase factors from Eq. (3.29) for the two chosen states. On the
other hand, since Û is unitary,
  

hψ1 |h0|Û Û |ψ2 i|0i = hψ1 |h0| |ψ2 i|0i (3.32)
= hψ1 |ψ2 i. (3.33)

Here we have used the fact that h0|0i = 1. Comparing the magnitudes of (3.31) and (3.33),
2
hψ1 |ψ2 i = hψ1 |ψ2 i ⇒ hψ1 |ψ2 i = 0 or 1. (3.34)

But aside from the trivial case of a one-dimensional Hilbert space, this cannot be true for
arbitrary |ψ1 i and |ψ2 i. For instance, for a two-dimensional space spanned by an orthonormal
basis {|0i, |1i}, we can pick
1  1
|ψ1 i = |0i, |ψ2 i = √ |0i + |1i ⇒ hψ1 |ψ2 i = √ . (3.35)
2 2

47
Y. D. Chong PH4401: Quantum Mechanics III

VI. DENSITY OPERATORS

We now introduce the density operator, which helps to streamline many calculations
in multi-particle quantum mechanics.
Consider a quantum system with a d-dimensional Hilbert space H . Given an arbitrary
state |ψi ∈ H , define
ρ̂ = |ψi hψ|. (3.36)
This is just the projection operator for |ψi, but in this context we call it a “density operator”.
Some other authors call it a density matrix, based on the fact that linear operators can
be represented as matrices. It has the following noteworthy features:

1. It is Hermitian.
2. Suppose Q̂ is an observable with eigenvalues {qµ } and eigenstates {|µi} (where µ is
some label that enumerates the eigenstates. If we do a Q̂ measurement on |ψi, the
probability of obtaining qµ is
2
Pµ = hµ|ψi = hµ| ρ̂ |µi. (3.37)

3. Moreover, the expectation value of the observable is


X X  
hQi = q µ Pµ = qµ hµ|ρ̂|µi = Tr Q̂ ρ̂ . (3.38)
µ µ

In the last equality, Tr[· · · ] denotes the trace, which is the sum of the diagonal ele-
ments of the matrix representation of the operator. The value of the trace is basis-
independent.

Now consider, once again, a composite system consisting of two subsystems A and B, with
Hilbert spaces HA and HB . Let’s say we are interested in the physical behavior of A, that
is to say the outcome probabilities and expectation values of any measurements performed
on A. These can be calculated from |ψi, the state of the combined system; however, |ψi
also carries information about B, which is not relevant to us as we only care about A.
There is a more economical way to encode just the information about A. We can define
the density operator for subsystem A (sometimes called the reduced density operator):
 
ρ̂A = TrB ρ̂ . (3.39)

Here, TrB [· · · ] refers to a partial trace. This means tracing over the HB part of the Hilbert
space H = HA ⊗ HB , which yields an operator acting on HA .
To better understand Eq. (3.39), let us go to an explicit basis. Let Q̂A be an observable
for HA with eigenbasis {|µi}, and let Q̂B be an observable for HB with eigenbasis {|νi}. If
the density operator of the combined system is ρ̂ = |ψihψ|, then
X   
ρ̂A = ˆ ˆ
I ⊗ hν| |ψihψ| I ⊗ |νi . (3.40)
ν

48
Y. D. Chong PH4401: Quantum Mechanics III

This is a Hermitian operator acting on the HA space. In the {|µi} basis, its diagonal matrix
elements are
X   
hµ|ρ̂A |µi = hµ|hν| |ψihψ| |µi|νi
ν
" !#
(3.41)
X
= hψ| |µihµ| ⊗ |νihν| |ψi
 ν
= hψ| |µihµ| ⊗ IˆB |ψi.

According to the rules of partial measurements discussed in Section II, this is precisely the
probability of obtaining qµ when measuring Q̂A on subsystem A:

Pµ = hµ|ρ̂A |µi. (3.42)

It follows that the expectation value for observable M̂ is


X h i
hQA i = qµ hµ|ρ̂A |µi = Tr Q̂A ρ̂A . (3.43)
µ

These results hold for any choice of basis. Hence, knowing the density operator for A, we
can determine the outcome probabilities of any partial measurement performed on A.
To better understand the properties of ρ̂A , let us write |ψi explicitly as
X
|ψi = ψµν |µi|νi, (3.44)
µν

|ψµν |2 = 1. Then
P
where µν
X
ρ̂ = ψµν ψµ∗ 0 ν 0 |µi|νi hµ0 |hν 0 |
µµ0 νν 0
X
ρ̂A = ψµν ψµ∗ 0 ν |µihµ0 |
µµ0 ν

X X
!
X
! (3.45)
= ψµν |µi ψµ∗ 0 ν hµ0 |
ν µ µ0
X X
= |ϕν ihϕν |, where |ϕν i = ψµν |µi.
ν µ

|ψµν |2 ≤ 1. Let us define


P
But |ϕν i is not necessarily normalized to unity: hϕν |ϕν i = µ

1 X
|ϕ̃ν i = √ |ϕν i, where Pν = |ψµν |2 . (3.46)
Pν µ

Note that each Pν is a non-negative real number in the range [0, 1]. Then
(
X each Pν is a real number in [0, 1], and
ρ̂A = Pν |ϕ̃ν ihϕ̃ν |, where (3.47)
ν
each |ϕ̃ν i ∈ HA , with hϕ̃ν |ϕ̃ν i = 1.

49
Y. D. Chong PH4401: Quantum Mechanics III

In general, we can define a density operator as any operator that has the form of Eq. (3.47),
regardless of whether or not it was formally derived via a partial trace. We can interpret
it as describing a ensemble of quantum states weighted by a set of classical probabilities.
Each term in the sum consists of (i) a weighting coefficient Pν which can be regarded as a
probability (the coefficients are all real numbers in the range [0, 1], and sum to 1), and (ii) a
projection operator associated with some normalized state vector |ϕ̃ν i. Note that the states
in the ensemble do not have to be orthogonal to each other.
From this point of view, a density operator of the form |ψihψ| corresponds to the special
case of an ensemble containing only one quantum state |ψi. Such an ensemble is called a
pure state, and describes a quantum system that is not entangled with any other system.
If an ensemble is not a pure state, we call it a mixed state; it describes a system that is
entangled with some other system.
We can show that any linear operator ρ̂A obeying Eq. (3.47) has the following properties:

1. ρ̂A is Hermitian.
2. hϕ|ρ̂A |ϕi ≥ 0 for any |ϕi ∈ HA (i.e., the operator is positive semidefinite).
3. For any observable Q̂A acting on HA ,
X
hQA i ≡ Pν hϕ̃ν |Q̂A |ϕ̃ν i
ν
X 
= Pν hϕ̃ν |µi hµ|Q̂|ϕ̃ν i using some basis {|µi}
µν
! (3.48)
X X
= hµ|Q̂ |ϕ̃ν ihϕ̃ν | |µi
µ ν
h i
= Tr Q̂ ρ̂A .
This property can be used to deduce the probability of obtaining any measurement
outcome: if |µi is the eigenstate associated with the outcome, the outcome probability
is hµ|ρ̂A |µi, consistent with Eq. (3.42). To see this, take Q̂ = |µihµ| in Eq. (3.48).
4. The eigenvalues of ρ̂A , denoted by {p1 , p2 , . . . , pdA }, satisfy
dA
X
pj ∈ R and 0 ≤ pj ≤ 1 for j = 1, . . . , dA , with pj = 1. (3.49)
j=1

In other words, the eigenvalues can be interpreted as probabilities. This also implies
that Tr[ρ̂A ] = 1.
This property follows from Property 3 by taking Q̂ = |ϕihϕ|, where |ϕi is any eigen-
vector of ρ̂A , and then taking Q̂ = IˆA .

VII. ENTANGLEMENT ENTROPY

Previously, we said that a multi-particle system is entangled if the individual particles


lack definite quantum states. It would be nice to make this statement more precise, and in

50
Y. D. Chong PH4401: Quantum Mechanics III

fact physicists have come up with several different quantitive measures of entanglement. In
this section, we will describe the most common measure, entanglement entropy, which
is closely related to the entropy concept from thermodynamics, statistical mechanics, and
information theory.
We have seen from the previous section that if a subsystem A is (possibly) entangled
with some other subsystem B, the information required to calculate all partial measurement
outcomes on A is stored within a reduced density operator ρ̂A . We can use this to define a
quantity called the entanglement entropy of A:
n  o
SA = −kb TrA ρ̂A ln ρ̂A . (3.50)

In this formula, ln[· · · ] denotes the logarithm of an operator, which is the inverse of the
exponential: ln(P̂ ) = Q̂ ⇒ exp(Q̂) = P̂ . The prefactor kb is Boltzmann’s constant, and
ensures that SA has the same units as thermodynamic entropy.
The definition of the entanglement entropy is based on an analogy with the entropy con-
cept from classical thermodynamics, statistical mechanics and information theory. In those
classical contexts, entropy is a quantitative measure of uncertainty (i.e, lack of information)
about a system’s underlying microscopic state, or “microstate”. Suppose a system P has W
possible microstates that occur with probabilities {p1 , p2 , . . . , pW }, satisfying i pi = 1.
Then we define the classical entropy
W
X
Scl. = −kb pi ln(pi ). (3.51)
i=1

In a situation of complete certainty where the system is known to be in a specific microstate


k (pi = δik ), the formula gives Scl. = 0. (Note that x ln(x) → 0 as x → 0). In a situation
of complete uncertainty where all microstates are equally probable (pi = 1/W ), we get
Scl. = kb ln W , the entropy of a microcanonical ensemble in statistical mechanics. For any
other distribution of probabilities, it can be shown that the entropy lies between these two
extremes: 0 ≤ Scl. ≤ kb ln W . For a review of the properties of entropy, see Appendix C.
The concept of entanglement entropy aims to quantify the uncertainty arising from a
quantum (sub)system’s lack of a definite quantum state, due to it being possibly entangled
with another (sub)system. When formulating it, the key issue we need to be careful about
is how to extend classical notions of probability to quantum systems. We have seen that
when performing a measurement on A whose possible outcomes are {qµ }, the probability
of getting qµ is Pµ = hµ|ρ̂A |µi. However, it is problematic to directly substitute these
probabilities {Pµ } into the classical entropy formula, since they are basis-dependent (i.e.,
the set of probabilities is dependent on the choice of measurement). Eq. (3.50) bypasses this
problem by using the trace, which is basis-independent.
In the special case where {|µi} is the eigenbasis for ρ̂A , the connection is easier to see.
From (3.49), the eigenvalues {pµ } are all real numbers between 0 and 1, and summing to
unity, so they can be regarded as probabilities. Then the entanglement entropy is
X
SA = −kb hµ|ρ̂A ln(ρ̂A )|µi
µ
X (3.52)
= −kb pµ ln(pµ ).
µ

51
Y. D. Chong PH4401: Quantum Mechanics III

Therefore, in this particular basis the expression for the entanglement entropy is consistent
with the classical definition of entropy, with the eigenvalues of ρ̂A serving as the relevant
probabilities.
By analogy with the classical entropy formula (see Appendix C), the entanglement entropy
has the following bounds:
0 ≤ SA ≤ kb ln(dA ), (3.53)
where dA is the dimension of HA .
The lower bound SA = 0 holds if and only if system A is in a pure state (i.e., it is not
entangled with any other system). This is because the bound corresponds to a situation
where ρ̂A has one eigenvalue that is 1, and all the other eigenvalues are 0 (see Appendix C).
If we denote the eigenvector associated with the non-vanishing eigenvalue by |ψi, then the
density matrix can be written as ρ̂A = |ϕihϕ|, which has the form of a pure state.
As a corollary, if we find that SA 6= 0, then ρ̂A cannot be written as a pure state |ψihψ|
for any |ψi, and hence it must describe a mixed state.
A system is said to be maximally entangled if it saturates the upper bound of (3.53),
SA = kb ln(dA ). This occurs if and only if the eigenvalues of the density operator are all
equal: i.e., pj = 1/dA for all j = 1, . . . , dA .

Example—Consider the following state of two spin-1/2 particles:


1  
|ψi = √ |+zi|−zi − |−zi|+zi . (3.54)
2
The density operator for the two-particle system is
1  
ρ̂(ψ) = |+zi|−zi − |−zi|+zi h+z|h−z| − h−z|h+z| . (3.55)
2
Tracing over system B (the second slot) yields the reduced density operator
1 
ρ̂A (ψ) = |+zih+z| + |−zih−z| . (3.56)
2
This can be expressed as a matrix in the {|+ zi, |− zi} basis:
1 
0
ρ̂A (ψ) = 2 1 . (3.57)
0 2

Now we can use ρ̂A to compute the entanglement entropy


1
ln 12
 
0 
SA = −kb Tr {ρ̂A ln(ρA )} = −kb Tr 2
1 = kb ln(2). (3.58)
0 2
ln 12

Hence, the particles are maximally entangled.

52
Y. D. Chong PH4401: Quantum Mechanics III

VIII. THE MANY WORLDS INTERPRETATION

We conclude this chapter by discussing a set of compelling but controversial ideas arising
from the phenomenon of quantum entanglement: the Many Worlds Interpretation as
formulated by Hugh Everett [8].
So far, when describing the phenomenon of state collapse, we have relied on the mea-
surement postulate (see Section II), which is part of the Copenhagen Interpretation of
quantum mechanics. This is how quantum mechanics is typically taught, and how physicists
think about the theory when doing practical, everyday calculations.
However, the measurement postulate has two bad features:

1. It stands apart from the other postulates of quantum mechanics, for it is the only
place where randomness (or “indeterminism”) creeps into quantum theory. The other
postulates do not refer to probabilities. In particular, the Schrödinger equation

i~ |ψ(t)i = Ĥ(t)|ψ(t)i (3.59)
∂t
is completely deterministic. If you know Ĥ(t) and are given the state |ψ(t0 )i at some
time t0 , you can in principle determine |ψ(t)i for all t. This time-evolution consists
of a smooth, non-random rotation of the state vector within its Hilbert space. A
measurement process, however, has a completely different character: it causes the
state vector to jump discontinuously to a randomly-selected value. It is strange that
quantum theory contains two completely different ways for a state to change.
2. The measurement postulate is silent on what constitutes a measurement. Does mea-
surement require a conscious observer? Surely not: as Einstein once exasperatedly
asked, are we really expected to believe that the Moon exists only when we look at
it? But if a given device interacts with a particle, what determines whether it acts via
the Schrödinger equation, or performs a measurement?

The Many Worlds Interpretation seeks to resolve these problems by positing that the
measurement postulate is not a fundamental postulate of quantum mechanics. Rather, what
we call “measurement”, including state collapse and the apparent randomness of measure-
ment results, is an emergent phenomenon that can be derived from the behavior of complex
many-particle quantum systems obeying the Schrödinger equation. The key idea is that a
measurement process can be described by applying the Schrödinger equation to a quantum
system containing both the thing being measured and the measurement apparatus itself.
We can study this using a toy model formulated by Albrecht [9]. Consider a spin-1/2
particle, and an apparatus designed to measure Sz . Let HS be the spin-1/2 Hilbert space
(which is 2D), and HA be the Hilbert space of the apparatus (which has dimension d). We
will assume that d is very large, as actual experimental apparatuses are macroscopic objects
containing 1023 or more atoms! The Hilbert space of the combined system is
H = HS ⊗ HA , (3.60)
and is 2d-dimensional. Let us suppose the system is prepared in an initial state
 
|ψ(0)i = a+ |+ zi + a− |− zi ⊗ |Ψi, (3.61)

53
Y. D. Chong PH4401: Quantum Mechanics III

where a± ∈ C are the quantum amplitudes for the particle to be initially spin-up or spin-
down, and |Ψi ∈ HA is the initial state of the apparatus.
The combined system now evolves via the Schrödinger equation. We aim to show that if
the Hamiltonian has the form
Ĥ = Ŝz ⊗ V̂ , (3.62)
where Ŝz is the operator corresponding to the observable Sz , then time evolution has an
effect equivalent to the measurement of Sz .
It turns out that we can show this without making any special choices for V̂ or |Ψi. We
only need d  2, and for both V̂ and |Ψi to be “sufficiently complicated”. We choose |Ψi to
be a random state vector, and choose random matrix components for the operator V̂ . The
precise generation procedures will be elaborated on later. Once we decide on |Ψi and V̂ , we
can evolve the system by solving the Schödinger equation
 
i
|ψ(t)i = U (t)|ψ(0)i, where Û (t) = exp − Ĥt . (3.63)
~
Because the part of Ĥ acting on the HS subspace is Ŝz , the result necessarily has the
following form:
 
|ψ(t)i = Û (t) a+ |+zi + a− |−zi ⊗ |Ψi
(3.64)
= a+ |+zi ⊗ |Ψ+ (t)i + a− |−zi ⊗ |Ψ− (t)i.
Here, |Ψ+ (t)i and |Ψ− (t)i are apparatus states that are “paired up” with the |+zi and |−zi
states of the spin-1/2 subsystem. At t = 0, both |Ψ+ (t)i and |Ψ− (t)i are equal to |Ψi; for
t > 0, they rotate into different parts of the state space HA . If the dimensionality of HA
is sufficiently large, and both V̂ and |Ψi are sufficiently complicated, we can guess (and we
will verify numerically) that the two state vectors rotate into completely different parts of
the state space, so that
hΨ+ (t)|Ψ− (t)i ≈ 0 for sufficiently large t. (3.65)
Once this is the case, the two terms in the above expression for |ψ(t)i can be interpreted
as two decoupled “worlds”. In one world, the spin has a definite value +~/2, and the
apparatus is in a state |Ψ+ i (which might describe, for instance, a macroscopically-sized
physical pointer that is pointing to a “Sz = +~/2” reading). In the other world, the spin
has a definite value −~/2, and the apparatus has a different state |Ψ− i (which might describe
a physical pointer pointing to a “Sz = −~/2” reading). Importantly, the |Ψ+ i and |Ψ− i
states are orthogonal, so they can be rigorously distinguished from each other. The two
worlds are “weighted” by |a+ |2 and |a− |2 , which correspond to the probabilities of the two
possible measurement results.
The above description can be tested numerically. Let us use an arbitrary basis for the
apparatus space HA ; in that basis, let the d components of the initial apparatus state vector
|Ψi be random complex numbers:
Ψ0
 
1  Ψ1 
|Ψi = √  .  , where Re(Ψj ), Im(Ψj ) ∼ N (0, 1). (3.66)
N  .. 
Ψd−1

54
Y. D. Chong PH4401: Quantum Mechanics III

In other words, the real and imaginary parts of each complex number Ψj are independently
drawn from the the standard normal (Gaussian) distribution, denoted by N (0, 1). The
normalization constant N is defined so that hΨ|Ψi = 1.
Likewise, we generate the matrix elements of V̂ according to the following random scheme:

Aij ∼ uij + ivij , where uij , vij ∼ N (0, 1)


1   (3.67)
V̂ = √  + † .
2 d
This scheme produces a d × d matrix with random components, √ subject to the requirement
that the overall matrix be Hermitian. The factor of 1/2 d is relatively unimportant; it
ensures that the eigenvalues of V̂ lie in the range [−2, 2], instead of scaling with d [10].
The Schrödinger equation can now be solved numerically. The results are shown below:


In the inital state, we let a+ = 0.7, so a− = 1 − 0.72 = 0.71414 . . . The upper panel
plots the overlap between the two apparatus states, |hΨ+ |Ψ− i|2 , versus t. In accordance
with the preceding discussion, the overlap is unity at t = 0, but subsequently decreases to
nearly zero. For comparison, the lower panel plots the entanglement entropy between the
two subsystems, SA = −kb TrA {ρ̂A ln ρ̂A }, where ρ̂A is the reduced density matrix obtained
by tracing over the spin subspace. We find that SA = 0 at t = 0, due to the fact that the
spin and apparatus subsystems start out with definite quantum states in |ψ(0)i. As the
system evolves, the subsystems become increasingly entangled, and SA increases up to
 
SAmax /kb = − |a+ |2 ln |a+ |2 + |a− |2 ln |a− |2 ≈ 0.693 (3.68)

This value is indicated in the figure by a horizontal dashed line, and corresponds to the
result of the classical entropy formula for probabilities {|a+ |2 , |a− |2 }. Moreover, we see
that the entropy reaches SAmax at around the same time that |hΨ+ |Ψ− i|2 reaches zero. This
demonstrates the close relationship between “measurement” and “entanglement”.
For details about the numerical linear algebra methods used to perform the above calcu-
lation, refer to Appendix D.

55
Y. D. Chong PH4401: Quantum Mechanics III

The “many worlds” concept can be generalized from the above toy model to the universe
as a whole. In the viewpoint of the Many Worlds Interpretation of quantum mechanics, the
entire universe can be described by a mind-bogglingly complicated quantum state, evolving
deterministically according to the Schrödinger equation. This evolution involves repeated
“branchings” of the universal quantum state, which continuously produces more and more
worlds. The classical world that we appear to inhabit is just one of a vast multitude. It is
up to you to decide whether this conception of reality seems reasonable. It is essentially a
matter of preference, because the Copenhangen Interpretation and the Many Worlds Inter-
pretation have identical physical consequences, which is why they are referred to as different
“interpretations” of quantum mechanics, rather than different theories.

Exercises

1. Let HA and HB denote single-particle Hilbert spaces with well-defined inner products.
That is to say, for all vectors |µi, |µ0 i, |µ00 i ∈ HA , that Hilbert space’s inner product
satisfies the inner product axioms
(a) hµ|µ0 i = hµ0 |µi∗
(b) hµ|µi ∈ R+ 0 , and hµ|µi = 0 if and only if |µi = 0.
(c) hµ| |µ i + |µ00 i = hµ|µ0 i + hµ|µ00 i
0


(d) hµ| c|µ0 i = chµ|µ0 i for all c ∈ C,




and likewise for vectors from HB with that Hilbert space’s inner product.
In Section I, we defined a tensor product space HA ⊗ HB as the space spanned by the
basis vectors {|µi ⊗ |νi}, where the |µi’s are basis vectors for HA and the |νi’s are
basis vectors for HB . Prove that we can define an inner product using
  
hµ| ⊗ hν| |µ i ⊗ |ν i ≡ hµ|µ0 i hν|ν 0 i = δµµ0 δνν 0
0 0
(3.69)
which satisfies the inner product axioms.
2. Consider the density operator
1 1
ρ̂ = |+zih+z| + |+xih+x| (3.70)
2 2
where |+xi = √12 (|+zi + |−zi). This can be viewed as an equal-probability sum of
two different pure states. However, the density matrix can also be written as
ρ̂ = p1 |ψ1 ihψ1 | + p2 |ψ2 ihψ2 | (3.71)
where |ψ1 i and |ψ2 i are the eigenvectors of ρ̂. Show that p1 and p2 are not 1/2.
3. Consider two distinguishable particles, A and B. The 2D Hilbert space of A is spanned
by {|mi, |ni}, and the 3D Hilbert space of B is spanned by {|pi, |qi, |ri}. The two-
particle state is

1 1 1 2 1 1
|ψi = |mi|pi + √ |mi|qi + √ |mi|ri + |ni|pi + √ |ni|qi + |ni|ri. (3.72)
3 6 18 3 3 3
Find the entanglement entropy.

56
Y. D. Chong PH4401: Quantum Mechanics III

Further Reading

[1] Bransden & Joachain, §14.1—14.4, §17.1–17.5


[2] Sakurai, §3.9

[3] A. Einstein, B. Podolsky, and N. Rosen, Can Quantum-Mechanical Description of Phys-


ical Reality Be Considered Complete?, Physical Review 47, 777 (1935). [link]

[4] J. S. Bell, On the Einstein-Podolsky-Rosen paradox, Physics 1, 195 (1964). [link]


[5] N. D. Mermin, Bringing home the atomic world: Quantum mysteries for anybody, Amer-
ican Journal of Physics 49, 940 (1981). [link]
[6] A. Aspect, Bell’s inequality test: more ideal than ever, Nature (News and Views) 398,
189 (1999). [link]
[7] A. K. Ekert, Quantum Cryptography Based on Bell’s Theorem, Physical Review Letters
67, 661 (1991). [link]
[8] H. Everett, III, The Theory of the Universal Wave Function (PhD thesis), Princeton
University (1956) [link]
[9] A. Albrecht, Following a “collapsing” wave function, Physical Review D 48, 3768 (1993).
[link]
[10] A. Edelman and N. R. Rao, Random matrix theory, Acta Numerica 14, 233 (2005). [link]

57

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy