QC Chapter2
QC Chapter2
31
32 QUANTUM COMPUTING
inner product): ⟨ψ|ψ⟩ = 1. Suppose two states |ψ1 ⟩ and |ψ2 ⟩ are physi-
cal states of the system. Then their linear superposition c1 |ψ1 ⟩ + c2 |ψ2 ⟩
(ck ∈ C) is also a possible state of the same system. This is called the
superposition principle.
A2 For any physical quantity (i.e., observable) a, such as energy, spin,
position and momentum, there exists a corresponding Hermitian oper-
ator A acting on the Hilbert space H. When we measure a in the state
|ψ⟩, we obtain one of the eigenvalues λj of the operator A and the state
undergoes an abrupt change to the corresponding eigenvector |λj ⟩. This
phenomenon is called the collapse of wave function. If we prepare
an ensemble of many identical state |ψ⟩ and measure a, the probability
of obtaining the outcome |λj ⟩ is pj = |⟨λj |ψ⟩|2 .
expand |ψ⟩ in terms of a complete orthonormal basis {|λj ⟩} as
Let us P
|ψ⟩ = k ck |λk ⟩, where ck = ⟨λk |ψ⟩ ∈ C is called the probability
amplitude, and consider ⟨ψ|A|ψ⟩. We find
X X X
⟨ψ|A|ψ⟩ = c∗j ck ⟨λj |A|λk ⟩ = c∗j ck λk δjk = λk |ck |2 .
j,k j,k k
where Pj = |λj ⟩⟨λj | is the projection operator, and the state immedi-
ately after the measurement is |λj ⟩ or equivalently
P |ψ⟩
p j , (2.4)
⟨ψ|Pj |ψ⟩
where the overall phase has been ignored.
The discussion is similar but more involved if H has an infinite dimen-
sion.
The Schrödinger equation (2.2) in Axiom A3 is formally solved to yield
i t
Z
|ψ(t)⟩ = T exp − H(t)dt |ψ(0)⟩ (2.6)
ℏ 0
if H depends on t, where T is the time-ordering operator defined by
A(t1 )B(t2 ), t1 > t2
T [A(t1 )B(t2 )] = ,
B(t2 )A(t1 ), t2 ≥ t1
34 QUANTUM COMPUTING
for a product of two operators. Note that T is required since H(t1 ) and
H(t2 ) do not commute with each other in general if t1 ̸= t2 . General-
ization to products of more than two operators should be obvious. We
write Eqs. (2.5) and (2.6) as
i t
Z
|ψ(t)⟩ = U (t)|ψ(0)⟩, where U (t) = T exp − H(t)dt .
ℏ 0
0 0 0 1
1
|ψ⟩ = √ (| ↑⟩ ⊗ | ↑⟩ + | ↓⟩ ⊗ | ↓⟩) (2.8)
2
1
c1 d2 = c2 d1 = 0, c1 d1 = c2 d2 = √
2
We can use the SVD to express |ψ⟩ = r,s crs |e1,r ⟩ ⊗ |e2,s ⟩ ∈ Cm ⊗ Cn in
P
a simpler form and detect whether |ψ⟩ is entangled or not as follows. For the
SVD of
Xk
(crs ) = U ΣV † = sj |uj ⟩⟨vj |,
j=1
∗
we can set |f1,j ⟩ = |uj ⟩ and |f2,j ⟩ = |vj ⟩ , the conjugate of |vj ⟩. Then
X k
X
|ψ⟩ = crs |e1,r ⟩ ⊗ |e2,s ⟩ = sj |f1,j ⟩ ⊗ |f2,j ⟩.
r,s j=1
Clearly, |ψ⟩ is a product state if and only if (crs ) has rank one, i.e., k = 1 and
s1 = 1.
EXAMPLE 2.2.1. Let {|e1,1 ⟩, |e1,2 ⟩, |e1,3 ⟩} and {|e2,1 ⟩, |e2,2 ⟩⟩} be the stan-
dard bases of C3 and C2 , and consider the bipartite state
1
|ψ⟩ = (|e1,1 ⟩|e2,1 ⟩ + |e1,1 ⟩|e2,2 ⟩ + i|e1,3 ⟩|e2,1 ⟩ + i|e1,3 ⟩|e2,2 ⟩) ∈ C2 ⊗ C3
2
whose coefficients form a matrix
11
1
C = 0 0.
2
i i
It is immediate that the matrix is rank one, and i,j |cij |2 = 1. So, |ψ⟩ is a
P
product state. We can find a simple representation of |ψ⟩ using the SVD of
C obtained in Example 1.6.2. We have
where
3
X 1
|f1,1 ⟩ = Ui1 |e1,i ⟩ = √ (|e1,1 ⟩ + i|e1,3 ⟩)
i=1
2
and
2
X
∗ 1
|f2,1 ⟩ = Vj1 |e2,j ⟩ = √ (|e2,1 ⟩ + |e2,2 ⟩).
j=1
2
where si > 0 are called the Schmidt coefficients and {|fa,i ⟩} is an orthonor-
mal set of Ha . The number r ∈ N is called the Schmidt number of |ψ⟩.
Proof. This is a direct consequence of SVD introduced in §1.6. Let |ψ⟩ be ex-
panded as in Eq. (2.7). Note that the coefficients cij form a dimH1 × dimH2
matrix C. We apply the SVD to obtain C = U ΣV † , where U and V are
unitary matrices and Σ is a matrix whose (j, j) elements are nonnegative real
numbers while all the other elements vanish. Now |ψ⟩ of Eq. (2.7) is put in
the form X
|ψ⟩ = Uik Σkl Vjl∗ |e1,i ⟩ ⊗ |e2,j ⟩.
i,j,k,ℓ
P P ∗
Now define |f1,k ⟩ = i Uik |e1,i ⟩ and |f2,k ⟩ = j Vjk |e2,j ⟩. Unitarity of U and
V guarantees that they are orthonormal vectors of H1 and H2 , respectively.
By noting that the (k, l) entry of Σ is sk δkl , we obtain
r
X
|ψ⟩ = sj |f1,j ⟩ ⊗ |f2,j ⟩,
j=1
It follows from the above proposition that a bipartite state |ψ⟩ is separable
if and only if its Schmidt number r is 1.
Generalization to a system with more components, i.e., a multipartite
system, should be obvious. A system composed of N components has a
Hilbert space
H = H1 ⊗ H 2 ⊗ . . . ⊗ H N , (2.10)
where Ha is the Hilbert space to which the ath component belongs. Classifi-
cation of entanglement in a multipartite system is far from obvious, and an
analogue of the Schmidt decomposition is not known to date for N ≥ 3.∗
The no-cloning theorem may be proved by using the special theory of rela-
tivity, which assumes no information can propagate faster than the speed of
light. [27]
Suppose Alice and Bob share a Bell state
1 1
|Ψ− ⟩ = √ (|0⟩|1⟩ − |1⟩|0⟩) = √ (|−⟩|+⟩ − |+⟩|−⟩). (2.12)
2 2
where |±⟩ = √12 (|0⟩ ± |1⟩). Readers are encouraged to verify the second
equality. Alice keeps the first qubit while Bob keeps the second. If Alice
wants to send Bob a bit “0”, she measures her qubit in {|0⟩, |1⟩} basis while
if she wants to send “1”, she employs {|+⟩, |−⟩} basis for her measurement.
Bob always measures his qubit in {|0⟩, |1⟩} basis.
After Alice’s measurment and before Bob’s measurment, Bob’s qubit is |0⟩
or |1⟩ if Alice sent “0” while it is |+⟩ or |−⟩ if Alice sent “1”.
Suppose Bob is able to clone his qubit. He makes many copies of his qubit
and measures them in {|0⟩, |1⟩} basis. If Alice sent “0”, Bob will obtain
0, 0, 0, . . . or 1, 1, 1, . . . while if she sent “1”, Bob will obtain approximately
50% of 0’s and 50% of 1’s. Suppose Bob received |±⟩ and made N clones, then
the probability of obtaining the same outcome is 1/2N −1 , which is negligible
if N is sufficiently large. Note that Bob obtains the bit Alice wanted to send
immediately after Alice’s measurement assuming it does not take long to clone
his qubit. This could happen even if Alice and Bob are separated many light
years apart, thus in contradiction with the special theory of relativity.
Quantum mechanics: Hilbert space formalism 39
2.4 Qubits
A (Boolean) bit assumes two distinct values, 0 and 1. Bits constitute the
building blocks of the classical information theory founded by C. Shannon.
Quantum information theory, on the other hand, is based on qubits.
2
while |1⟩ represents a state
1
| ⟩ = √ (| ↔⟩ − | ↕⟩).
↔
2
Note that if | ↔⟩ (| ↕⟩) corresponds to an eigenstate of σz with the eigen-
value +1 (−1), respectively, then | ⟩ (| ⟩) corresponds to an eigenstate
↔
In any case, we have to fix a set of basis vectors when we carry out quantum
information processing. All the physics should be described with respect to
this basis. In the following, the basis is written in an abstract form as {|0⟩, |1⟩},
unless otherwise stated.
Note that the third example of a qubit above suggests that a quantum
system with more than two states may be employed for information storage
and information processing. If a quantum system admits three different states,
it is called a qutrit, while if it takes d different states, it is called a qudit.
A spin S particle, for example, takes d = 2S + 1 spin states and works as a
qudit. The significance of qutrits and qudits in information processing is still
to be explored.
Qubits can be used as a Hardware random number generator or true
random number generator to generate cryptographic keys. (See [28] for
background and references for random number generation.) Based on the
quantum postulates, one can generate a zero-one sequence by measuring a
sequence of quibts in a state √12 (|0⟩ + |1⟩) one by one.
It is convenient to assume the vector |0⟩ corresponds to the classical value
0, while |1⟩ to 1 in quantum computation. Moreover it is possible for a qubit
to be in a superposition state:
|ψ⟩ = a|0⟩ + b|1⟩ with a, b ∈ C, |a|2 + |b|2 = 1. (2.14)
The fundamental requirement of quantum mechanics is that if we make mea-
surement on |ψ⟩ to see whether it is in |0⟩ or |1⟩, the outcome will be 0 (1) with
the probability |a|2 (|b|2 ), and the state immediately after the measurement
is |0⟩ (|1⟩).
Although a qubit may take infinitely many different states, it should be kept
in mind that we can extract from it as the same amount of information as that
of a classical bit. Information can be extracted only through measurements.
When we make measurement on a qubit, the state vector “collapses” to the
eigenvector that corresponds to the eigenvalue observed. Suppose that a spin
is in a state a|0⟩ + b|1⟩. If we observe that the z-component of the spin
is +1/2, the system immediately after the measurement is definitely in the
state |0⟩. This happens with probability ⟨ψ|0⟩⟨0|ψ⟩ = |a|2 . The outcome
of a measurement on a qubit is always one of the eigenvalues, which we call
abstractly 0 and 1, just like for a classical bit. As a result, if we can make
measurements of a large number of copies of unknown state, we may estimate
|a| and |b|, but we cannot estimate the coefficients a and b.
It is useful, for many purposes, to express a state of a single qubit graph-
ically. Note that every qubit |ψ⟩ = a|0⟩ + b|1⟩ corresponds to the rank one
orthogonal projection
1 1 + nz nx − iny 1
ρ = |ψ⟩⟨ψ| = = (I2 + nx σx + ny σy + nz σz ),
2 nx + iny 1 − nz 2
where σx , σy , σz are the Pauli matrices, (nx − iny )/2 = ab̄, nz = 2|a|2 − 1.
As we will see in the next chapter, the matrix ρ is a density matrix corre-
Quantum mechanics: Hilbert space formalism 41
sponding to a pure state. We use the notation in Proposition 1.5.8, and let
n̂ = (nx , ny , nz ) ∈ R3 be a unit vector, σ = (σx , σy , σz ), and
nz nx − iny
A = n̂ · σ = .
nx + iny −nz
Then
A = P1 − P2 with P1 = (I + A)/2 and P2 = (I − A)/2,
where |ψ⟩⟨ψ| = P1 is precisely the eigenprojection of the eigenvalue 1 of |ψ⟩⟨ψ|.
In particular, we may parametrize a one-qubit state |ψ⟩ with θ and ϕ as
θ θ
|ψ⟩ = |ψ(θ, ϕ)⟩ = cos |0⟩ + eiϕ sin |1⟩, θ ∈ [0, π], ϕ ∈ [0, 2π), (2.15)
2 2
where we take advantage of the freedom of the phase to make the coefficient
of |0⟩ real. Then
1
|ψ⟩⟨ψ| = (I + n̂ · σ)
2
with n̂ = (sin θ cos ϕ, sin θ sin ϕ, cos θ) corre-
sponding to the unique point n̂ on the unit
sphere in R3 using the polar co-ordinates. So,
there is a natural correspondence between a
unit vector n̂(θ, ϕ) and a state vector |ψ(θ, ϕ)⟩.
Namely, a state |ψ(θ, ϕ)⟩ is expressed as a unit
vector n̂(θ, ϕ) on the surface of the unit sphere,
The Bloch sphere
called the Bloch sphere.
Pr
A more general density matrix ρ can be constructed as ρ = j=1 pr |ψj ⟩⟨ψj |,
where |ψ1 ⟩⟨ψ1 |, . . . , |ψr ⟩⟨ψr | are rank one orthogonal projections and
p1 , . . . , pr are positive numbers summing up to one. Then ρ has the form
1 X
ρ= I+ uj σi , (2.16)
2 j=x,y,z
The set
1 1
{|Φ+ ⟩ = √ (|00⟩ + |11⟩), |Φ− ⟩ = √ (|00⟩ − |11⟩),
2 2
(2.18)
1 1
|Ψ+ ⟩ = √ (|01⟩ + |10⟩), |Ψ− ⟩ = √ (|01⟩ − |10⟩)}
2 2
is an orthonormal basis of a two-qubit system and is called the Bell basis.
Each vector is called the Bell state or the Bell vector. Note that all the
Bell states are entangled.
Among three-qubit entangled states, the following two states are important
for various reasons and hence deserve special names. The state
1
|GHZ⟩ = √ (|000⟩ + |111⟩) (2.19)
2
is called the Greenberger-Horne-Zeilinger state and is often abbreviated
as the GHZ state [21]. Another important three-qubit state is the W state
[22],
1
|W⟩ = √ (|100⟩ + |010⟩ + |001⟩). (2.20)
3
The GHZ state has a different entanglement pattern as the W state. If any
qubit in the GHZ state is measured, the resulting state is a separable state
while if any qubit in the W state is measured, there is a probability of 2/3 that
the resulting state is entangled. Accordingly. it is impossible to transform
the GHZ state to the W state by local operations and vice versa.
2.5 Measurement
Classical information theory is formulated independently of measurements of
the system under consideration. This is because the readout of the result
is always the same for anyone and at any time, provided that the system
processes the same information without error. This is completely different
in quantum information processing. Measurement is an essential part of the
theory as we see below.
By making a measurement on a system, we project the state vector to one of
the basis vectors that the measurement equipment defines.† Suppose we have
a state vector |ψ⟩ = a|0⟩ + b|1⟩ and measure it to see if it is in the state |0⟩
or |1⟩. Depending on the system, this means if a spin points up or down or a
photon is polarized horizontally or vertically, for example. The result is either
0 or 1. In the first case, the state “collapses” to |0⟩ while in the second case, to
|1⟩. We find, after many measurements, the probability of obtaining outcome
0 (1) is |a|2 (|b|2 ). Here, we may assume that the observable is associate with
the Hermitian operator M = λ0 |0⟩⟨0| + λ1 |1⟩⟨1|. Note that the eigenvalues λ0
and λ1 represent the readings produced by the apparatus if the states |0⟩ and
|1⟩ are measured, respectively. They may be determined or limited by the set
up of the measuring device.
To be more formal, P we construct a Hermitian matrix corresponding to the
observable M = m λm Pm such that the eigenprojection Pm will serve as
the measurement operator Pm such that the probability of obtaining the
outcome λm in the state |ψ⟩ is
Pm |ψ⟩
|m⟩ = p . (2.22)
p(m)
In the above example, the measurement operators are nothing but projection
operators; P0 = |0⟩⟨0| and P1 = |1⟩⟨1|. In fact, we have
and
P |ψ⟩ a
p0 = |0⟩ ≃ |0⟩,
p(0) |a|
and similarly for the other case P1 . It should be noted that a quantum state
is defined up to a phase and hence a/|a| does not play any role. ‡
P many copies of a particular state |ψ⟩. If we measure
Suppose we are given
an observable M = m λm Pm in each of the copies, the expectation value of
M is given, in terms of the projection operators, by
X X
E(M ) = λm p(m) = λm ⟨ψ|Pm |ψ⟩
m m
X
= ⟨ψ| λm Pm |ψ⟩ = ⟨ψ|M |ψ⟩, (2.23)
m
P
where use has been made of the spectral decomposition M = m λm Pm . The
standard deviation of the measurement outcomes of M is given by
p p
∆(M ) = ⟨(M − ⟨M ⟩)2 ⟩ = ⟨M 2 ⟩ − ⟨M ⟩2 . (2.24)
‡ We will consider some more general measurement operators. One may see [18] if desired.
Quantum mechanics: Hilbert space formalism 45
into one of the subspaces Hmi randomly with the probability |cmi |2 .
Measurement gives an alternative viewpoint to entangled states. A state
is not entangled if a measurement of a qubit does not affect the state of the
other qubits. Suppose the first qubit of the state
1
√ (|00⟩ + |11⟩)
2
was measured to be 0 (1). Then the outcome of the measurement of the
second qubit is definitely 0 (1). Therefore the measurement of the first qubit
affects the outcome of the measurement on the second qubit, which shows
that the initial state is an entangled state. In other words, there exists a
strong correlation between the two qubits. This correlation may be used for
information processing as will be shown later. In contrast with this, the state
1
√ (|00⟩ + |01⟩) is not entangled since it can be written as
2
1 1
√ (|00⟩ + |01⟩) = |0⟩ ⊗ √ (|0⟩ + |1⟩).
2 2
Regardless of the measurement of the second qubit, the measurement of the
first qubit definitely yields 0. Moreover, the second qubit is measured to be
0 (1) with the probability 1/2, independently of whether the first qubit is
measured or not.
FIGURE 2.1
EPR pair produced by a source in the middle. One qubit is sent to Alice and
the other to Bob.
|01⟩ (|10⟩), and Bob will definitely observe |1⟩ (|0⟩) in his measurement. The
change of the state
1
√ (|01⟩ − |10⟩) → |01⟩ or |10⟩ (2.26)
2
takes place instantaneously even when they are separated by a large distance.
It seems that Alice’s measurement propageted to Bob’s qubit instantaneously,
and it violates the special theory of relativity. This is the very point EPR pro-
posed to defeat quantum mechanics, which claims that nothing can propagate
faster than the speed of light.
Note, however, that nothing has propagated from Alice to Bob and vice
versa, upon Alice’s measurement. Clearly no energy has propagated. What
about information? It is impossible for Alice to control her and hence Bob’s
readings. Therefore it is impossible to use EPR pairs to send a sensible
message from Alice to Bob. If they could, the message would be sent in-
stantaneously, which certainly violates the special theory of relativity. If a
large number of EPR pairs are sent to Alice and Bob and they independently
measure their qubits, they will observe random sequences of 0 and 1. They
notice that their readings are strongly correlated only after they exchange
their sequences by means of classical communication, which can be done at
most with the speed of light.
apply one of his two measurement schemes, say, S and T , each will produce
a measured value in {1, −1}.
Let us consider
QS + RS + RT − QT = (Q + R)S + (R − Q)T.
1 1
E(QS) = ⟨Ψ− |Q ⊗ S|Ψ− ⟩ = √ , E(RS) = ⟨Ψ− |R ⊗ S|Ψ− ⟩ = √ ,
2 2
1 −1
E(RT ) = ⟨Ψ− |R ⊗ T |Ψ− ⟩ = √ , E(QT ) = ⟨Ψ− |Q ⊗ T |Ψ− ⟩ = √ ,
2 2
and hence √ √
E(QS + RS + RT − QT ) = 4/ 2 = 2 2. (2.28)
This equality clearly violates the Bell inequality.
To determine whether (2.27) or (2.28) is valid, Alice and Bob can estimate
E(QS) by performing measurements on many copies of |Ψ− ⟩, and record their
results. After the experiments, they can multiply their measurements when
they used the measurement schemes Q and S, respectively. Similarly, they
can estimate E(RS), E(RT ), E(QT ), so as to obtain an estimate of E(QS +
RS + RT − QT ).
Quantum mechanics: Hilbert space formalism 49
ℏ
∆(Q)∆(P ) ≥ .
2
Note that AB − BA = γI with γ ̸= 0 can only happen for infinite dimensional
operators as Tr (AB − BA) = 0 for finite dimensional operators. The Uncer-
tainty principle in terms of standard deviation has been formulated first in [7]
and [8].
d|ψ(t)⟩
iℏ = H|ψ(t)⟩
dt
where |ψ(t)⟩ = U (t)|ψ(0)⟩. The expectation value of an observable a at t is
given by ⟨ψ(t)|A|ψ(t)⟩. This description with time-dependent state |ψ(t)⟩ and
time-independent operator A is called the Schrödinger picture.
There is another equivalent description in which the state is time-
independent while the operator depends on time, called the Heisenberg
picture. Let
A(t) = U (t)† AU (t) (2.30)
and consider ⟨ψ(0)|A(t)|ψ(0)⟩. We find
which shows the two pictures give the same expectation value at any t. It is
also shown that any matrix element satisfies
⟨ϕ(t)|A|ψ(t)⟩ = ⟨ϕ(0)|A(t)|ψ(0)⟩
FIGURE 2.2
Probability P↑ (t) with which a spin is observed in the ↑-state and P↓ (t) ob-
served in the ↓-state.
The wave function |ψ(t)⟩ (t > 0) is then found from Eq. (2.5) to be
ω
|ψ(t)⟩ = exp i σx t |ψ(0)⟩. (2.36)
2
The matrix exponential function in this equation is evaluated with the help of
Eq. (1.23) and we find
! !
cos ωt/2 i sin ωt/2 1 cos ωt/2
|ψ(t)⟩ = = . (2.37)
i sin ωt/2 cos ωt/2 0 i sin ωt/2
ω ω
|ψ(t)⟩ = cos t|σz = +1⟩ + i sin t|σz = −1⟩.
2 2
Therefore we find the spin is in the spin-up state with the probability P↑ (t) =
cos2 (ωt/2) and in the spin-down state with the probability P↓ (t) = sin2 (ωt/2)
as depicted in Fig. 2.2. Of course, the total probability is independent of time
since cos2 (ωt/2) + sin2 (ωt/2) = 1. This result is consistent with classical
spin dynamics. The Hamiltonian (2.35) depicts a spin under a magnetic field
along the x-axis. Our initial condition signifies that the spin points the z-
direction at t = 0. Then the spin starts precession around the x-axis, and the
z-component of the spin oscillates sinusoidally as is shown above.
Next let us take the initial state
1 1
|ψ(0)⟩ = √ ,
2 1
Quantum mechanics: Hilbert space formalism 53
d
i |ϕ(t)⟩ = H̃|ϕ(t)⟩, (2.46)
dt
where
−iωσz t/2 iωσz t/2 −iωσz t/2 d iωσz t/2 1 −ω0 + ω ω1
H̃ = e He − ie e =
dt 2 ω1 ω0 − ω
δ ω1
= − σz + σx (2.47)
2 2
is, in fact, time-independent. Here δ = ω0 − ω stands for the “detuning”
between ω and ω0 . Note that the Hamiltonian H̃ can be put into the form
(2.39) as
∆ ω1 δ
q
H̃ = σx − σz , ∆ ≡ δ 2 + ω12 . (2.48)
2 ∆ ∆
Now it is easy to solve Eq. (2.46). The time evolution operator is obtained
using Eq. (2.40) as
∆t ω1 δ ∆t
Ũ (t) = cos I −i σx − σz sin
2 ∆ ∆ 2
∆t δ ∆t ω1 ∆t
cos + i sin −i sin
2 ∆ 2 ∆ 2
= . (2.49)
ω1 ∆t ∆t δ ∆t
−i sin cos − i sin
∆ 2 2 ∆ 2
The wave function |ϕ(t)⟩ with the initial condition |ϕ(0)⟩ = (1, 0)t is
∆t δ ∆t
cos + i sin
2 ∆ 2
|ϕ(t)⟩ = Ũ (t)|ϕ(0)⟩ = . (2.50)
ω1 ∆t
−i sin
∆ 2
We find |ψ(t)⟩ from Eq. (2.45) as
iωt/2 ∆t δ ∆t
e cos + i sin
2 ∆ 2
|ψ(t)⟩ = eiωσz t/2 |ϕ(t)⟩ =
. (2.51)
ω1 ∆t
−ie−iωt/2 sin
∆ 2
Quantum mechanics: Hilbert space formalism 55
Suppose the applied field is in resonance with the energy difference of two
levels, namely ω = ω0 . We obtain δ = 0 and ∆ = ω1 in this case. The wave
function |ψ(t)⟩ at later time t > 0 is
ω1 t
eiω0 t/2 cos
2
|ψ(t)⟩ = eiωσz t/2 |ϕ(t)⟩ =
. (2.52)
−iω0 t/2 ω1 t
−ie sin
2
The probability with which the system is found in the ground (excited) state
of H0 is given by
⟨ψ(θ, ϕ)|σ|ψ(θ, ϕ)⟩ = (sin θ cos ϕ, sin θ sin ϕ, cos θ), (2.56)
(⟨ψ(θ, ϕ)|σx |ψ(θ, ϕ)⟩, ⟨ψ(θ, ϕ)|σy |ψ(θ, ϕ)⟩, ⟨ψ(θ, ϕ)|σz |ψ(θ, ϕ)⟩).
Quantum mechanics: Hilbert space formalism 57
EXERCISE 2.5. Find the density matrix of a pure state (2.15) and write
it in the form of Eq. (2.16).
EXERCISE 2.7. The Bell basis is obtained from the binary basis {|00⟩, |01⟩,
|10⟩, |11⟩} by a unitary transformation. Write down the unitary transforma-
tion explicitly.
where |x⟩ stands for the tensor product state |bn−1 bn−2 . . . b0 ⟩ with x =
bn−1 2n−1 + bn−2 2n−2 + . . . + b0 and N is the normalization constant. The
first register is for the input x, while the second one is for the corresponding
output f (x). Note that Uf acts on all possible states simultaneously.
Let f (x) = ax mod N , where a and N are coprime, and consider the state
" 511
# 511
1 X 1 X
Uf √ |x⟩|0⟩ = √ |x⟩|ax mod N ⟩
512 x=0 512 x=0
with a = 6 and N = 91. Suppose the measurement of the first register results
in (1) x = 11, (2) x = 23 and (3) x = 35. What is the state immediately after
each measurement?
References
[1] P. A. M. Dirac, Principles of Quantum Mechanics (4th ed.), Clarendon
Press (1981).
[2] L. I. Shiff, Quantum Mechanics (3rd ed.), McGraw-Hill (1968).
[3] A. Messiah, Quantum Mechanics, Dover (2000).
[4] J. J. Sakurai and J. Napolitano, Modern Quantum Mechanics (3rd Edi-
tion), Cambridge University Press, Cambridge (2020).
58 QUANTUM COMPUTING