QO Note 1 Review QM
QO Note 1 Review QM
QO Note 1 Review QM
Jhih-Sheng Wu
2024
Contents
1 Wavefunction 3
2 Dirac Notations 4
5 Harmonic Oscillators 14
1
7 Density Operator Formulation 19
2
Quantum mechanics is a modern mathematical theory used to describe the quantum phenom-
ena. Although many scientists think it is not the ultimate theory, it is the most accurate theory
today that describes the experiments. Quantum mechanics is formulated under the postulates,
which are made by the scientists to explain the experiments. In classical mechanics, a physical
system consists of physical quantities which have definite values. For examples, the position x
and the momentum p of a particle at any given time t are assumed to be some numbers. On
the contrary, a physical system in quantum mechanics is described by a state |ψi. The notation
|ψi is called a ket. In a closed system, the state |ψi contains all the information of the systems.
The exotic part of quantum mechanics is that even |ψi is complete, the outcome of observed
quantities are still probabilistic.
1 Wavefunction
Let’s use the wavefunction to elaborate the nature of probability. The wavefunction of a particle
is obtained by writing |ψi in the x basis |xi,
ψ(x) ≡ hx|ψi. (1)
For a given wavefunction ψ(x), the probability to find the particle to be at x is |ψ(x)|2 dx. Since
the total probability is one, the normalization of a state requires that
Z
|ψ(x)|2 dx = 1. (2)
Eq. (4), the Schrödinger’s equation, is only one example of the Hamiltonian. In quantum me-
chanics (or classical mechanics), a Hamiltonian is roughly speaking the total energy of a system.
Thus, the Schrödinger’s equation only describes the kinetic energy and the potential energy of
an electron. If we want to describe the other energies, we need to use the other Hamiltonians.
This will be elaborated in Postulate 2.
3
Example 1: Plane Wave
Let |pi be the eigenstate of the momentum operator p̂ such that
p̂|pi = p|pi.
The state |pi has a well-defined momentum p. A plane wave is indeed the projection of
|pi onto the x basis.
px
hx|pi = ei ~ .
If instead, we project the state |pi onto the p basis,
That is, a plane wave in the p space is a delta function. This relation is analogous to the
Fourier transform of a single frequency signal eiωt is a delta function.
2 Dirac Notations
In quantum mechanics, the Bra-Ket notations are convenient tools. Any states are written as
kets |ψ1 i, |ψ2 i, |ψ3 i,... You can think a ket as a column vector. However, the representation of
a column vector depends on the bases. For example, in the position basis, a ket can be defined
as:
ψ(x1 )
ψ(x2 )
|ψi = .. . (7)
.
ψ(xN )
The role of a bra is similar to row vectors in linear algebra. A bra is defined as the hermitian
conjugate of a ket. For examples, in the position basis, a bra can be defined as:
hψ| = ψ ∗ (x1 ) ψ ∗ (x2 ) . . . ψ ∗ (xN ) . (9)
4
The inner product of two states |ψi and |φi is
hψ|φi, (11)
which is a complex number. The inner product hψi |ψi i is the probability to find the particle in
the ith state. The outer product of two states |ψi and |φi is
|φihψ| (12)
which is a matrix.
3. Calculate |aihb| and |biha|. Are they complex conjugate of each other?
If |ψi is to describe a single particle, the normalization of a state requires the inner product
hψ|ψi = 1 (15)
or in a specific basis
X
|ψi |2 = 1, (16)
i
5
and for a continuous variable like x,
Z
dx|ψ(x)|2 = 1. (17)
As it should be, the operator x̂ is a diagonal matrix in the position basis. In the Dirac’s notation,
the expectation value of x is
The state vector |ψi contains all the information. The state vector can be written as a sum of
other (basis) vectors.
X
|ψi = αi |ψi i (22)
i
The probability to find the system in the ith state is |αi |2 . The simplest example is the qubit,
6
Figure 1: Bloch sphere.
Without losing the generality, the qubit can be written as (|α|2 + |β|2 = 1)
θ θ iφr
iφg
|ψi = e cos |0i + sin e |1i , (24)
2 2
where φg is the global phase, and φr is the relative phase between the |0i and |1i states. Without
comparing with another qubit, the phase ψg does not have much meaning. The degrees of
freedoms of a qubit are given by θ and φr , which correspond to a surface of a sphere. The
space of a qubit is called the Bloch sphere.
7
H 0
U (t, t 0 ) = e−i ~ (t −t) , (27)
eX eY = eZ
1 1 1
Z = X + Y + [X, Y ] + [X, [X, Y ]] − [Y , [X, Y ]] + · · · ,
2 12 12
• H = H†
8
In the case of an electron, the angular momentum operator is σ = (σx , σy , σz ), where σi are the
Pauli matrices (2 by 2 matrices). Hence, the dimension of the Hamiltonian is two.
Let’s consider a system with N levels of the energies E1 , E2 , ..., EN . The energy eigenstates,
|Ei i, satisfy
H|Ei i = Ei |Ei i. (32)
The Hamiltonian in the energy bases |Ei i is diagonal
E1 0 0 0
0 E2 0 0
(33)
H = .. .. . . .. .
.
. . .
0 0 · · · EN
Postulate 3: Measurement
Quantum measurement (collapse). A measurement makes a system |ψi collapse randomly
into some state |ψi i. The possible outcome states |ψi i depend on the measurements.
For example, if we measure the position of a particle, the outcome states are |xi with
−∞ < x < ∞. A measurement is described by a set of operators {Mm }, where m denotes
all the possible outcome states. After a measurement, the state becomes
Mm |ψi
q (35)
†
hψ|Mm Mm |ψi
9
Exercise 3: Qubit Measurement
√
3
The initial qubit state is 12 |0i + 2 |1i.
Postulate 4: Hermiticity
Any physical observables are Hermitian operators. For example, in the position basis, the
position and momentum operators are
x̂ = x (40)
∂
p̂ = −i~ (41)
∂x
Let  be the physical observable operator. The expectation value of  of a state |ψi is
Note that the eigenvectors of a Hermitian operator form a complete set of bases of the
space.
The eigenvectors |Ai i of A forms a complete set of bases of the state space. The eigenstates are
orthogonal and normal,
10
The completeness implies that the identity 1 is,
X
1= |Ai ihAi | (47)
i
[A, B] = c (49)
The state |ψi of a physical system contains all the informations, but is not the direct observable.
We do not directly see or measure the state |ψi, but rather we measure the physical quantities
11
such position, momentum, and so on. These physical quantities are the expectations of the
observable operators (such as x̂ and p̂). For example, when we measure the position of a state,
the averaged outcome is
x(t) = hψ(t)|x̂|ψ(t)i.
This outcome is called the observable.
More generally, let’s consider a physical observable operator Â, and its expectation is
The number A(t) is what we measured in the experiment. To obtain the evolution of A(t)
belongs to the subject of quantum dynamics. In quantum dynamics, there are three main
pictures to solve the problems.
Consider that the observable operator A is static and the states |ψ(t)i is evolving.
iHt
|ψ(t)i ≡ |ψ(t)iS = e− ~ |ψ(0)i (55)
Consider that the observable operator A(t) is dynamic and the states |ψ(t)i is static.
iHt iHt
Ah ≡ A(t) = e ~ Ae− ~ , (57)
∂Ah
i~ = [Ah , H]. (59)
∂t
12
Exercise 5: Proof of the Heisenberg’s equation
iHt
Let U (t) = e− ~ so that Ah = U † AU . Differentiating Ah with respect to t gives
∂Ah ∂U † ∂U
= AU + U † A (60)
∂t ∂t ∂t
First, show that the derivative of U (t) is
∂
i~ U (t) = HU (t). (61)
∂t
Use the two above equations to prove the Heisenberg’s equation.
At t = 0, x̂h (0) = x̂(0). The operator x̂h (t) at a later time, can be obtained by
∂x̂h
i~ = [x̂h , H].
∂t
You can show that
∂x̂h
= −mω2 x̂h .
∂t
When the Hamiltonian includes two terms: one is the original Hamiltonian H0 and the inter-
action with the external system V (t), it is convenient to use the interaction picture, where both
the states and the operator are evolving. The total Hamiltonian is H = H0 + V (t). The state
|ψiI is
H0 t
|ψiI = ei ~ |ψ(t)iS , (62)
13
and the operator AI is
H0 t H0 t
AI = ei ~ Ae−i ~ , (63)
∂
i~ |ψi = VI (t)|ψiI , (64)
∂t I
H0 t H0 t
VI (t) ≡ ei ~ V (t)e−i ~ . (65)
VI
Note that the solution to Eq. (64) is not |ψ(t)iI = e−i ~ t |ψ(0)iI because the VI (t) is time-
dependent. The solution to to Eq. (64) is
∂AI
i~ = [AI , H0 ]. (68)
∂t
5 Harmonic Oscillators
p2 mω2 x2
H= + , (69)
2m 2
√
where ω = k/m and k is the spring constant. We define the creation operator a† and the
annihilation operator a,
r
mω ip
a= x+ , (70)
2~ mω
r
mω ip
†
a = x− . (71)
2~ mω
[a, a† ] = 1. (72)
14
Use the relation [x, p] = i~.
[N , a] = −a, (77)
[N , a† ] = a† , (78)
As a result, we have
N a† |ni = a† N + a† |ni = (n + 1)a† |ni, (79)
N a|ni = (aN − a) |ni = (n − 1)a|ni, (80)
√
a|ni = n|n − 1i, (85)
√
a† |ni = n + 1|n + 1i. (86)
15
Note 2: Representation in the number basis
The number n is the number of the energy quanta. The smallest number of n is n = 0.
The physical meaning of |ni is a state containing n energy quanta. Thus, |ni is called the
number state. The energy of a harmonic oscillator is
1
En = n + ~ω (87)
2
The 12 ~ω is interpreted as the vacuum energy since it exists even when n = 0. Applying
√
a creation operator on the |ni, the state |ni becomes n + 1|n + 1i, that is, the a† will
create one single quantum to the original state. Similarly, the a will annihilate one energy
quantum from the system. We can also prove that
(a† )n
|ni = √ |0i. (88)
n!
The position operator x and momentum operator p can be expressed as
r
~
x= (a + a† ) (89)
2mω
r
mω~
p=i (−a + a† ) (90)
2
As the familiar wave function ψ(x), we can express the |ni in the x bases. The wavefunctions
are ψn (x) ≡ hx|ni. Let’s solve the ground states first ψ0 (x). We start with
a|0i = 0 (91)
⇒hx|a|0i = 0 (92)
r
mω ip
⇒ x x+ 0 =0 (93)
2~ mω
!
~ ∂
⇒ x+ ψ (x) = 0 (94)
mω ∂x 0
2
1 − 12 xx
⇒ψ0 (x) = 1/4 √ e 0 , (95)
π x0
q
where x0 = ~
mω
16
Figure 2: Wavefunction ψn (x).
~
σ (x)σ (p) = . (96)
2
Using Eqs. (88) and (95), we obtain the expression for φn (x),
!n 2
1 2 ∂ − 21 xx
(97)
ψn (x) = √ x − x0
e 0 .
π1/4 2n n!x0n+1/2 ∂x
17
In terms of x and p, Eqs. (101) and (102) read
!
ip(t) ip(0) −iωt
a(t) = x(t) + = x(0) + e , (103)
mω mω
!
ip(t) ip(0) iωt
†
a (t) = x(t) − = x(0) − e . (104)
mω mω
Coherence refers to many meanings in different circumstances. We consider its usages in the
context of physics. Roughly speaking, coherence means that two (or more than two) states
(waves, particles) have a well-defined correlation as time t or positions x change.
For instance, there are two waves with the complex amplitudes ψ1 (x, t) and ψ2 (x, t), respec-
tively. In terms of the absolute amplitudes A and phases φ, they can be written as ψ1 (x, t) =
A1 (x, t) exp(iφ1 (x, t)) and ψ2 (x, t) = A2 (x, t) exp(iφ2 (x, t)). Coherence means that the differ-
ence between their phases φ1 − φ2 is constant as t or x changes. In the two-slit experiments,
coherent light source is required to produce interference patterns. In this definition, two plane
waves exp(ik1 x − ω1 t + φ1 ) and exp(ik2 x − ω2 t + φ2 ) are coherent if k1 = k2 and ω1 = ω2 . The
18
reason for the same frequencies or the same wavenumber is as following. When we measure the
interference, frequently we collect the data for a long time over many periods . The interference
signal is the time-average of the product ψ1∗ (t)ψ2 (t). |ψ1 (t)|2 and |ψ2 (t)|2 are the background
intensity. The time-averaged interference is
RT h i
∗
2Re ψ1 (t)ψ2 (t)
Iinterference = lim 0 . (108)
T →∞ T
If the two waves have different frequencies, the time-average vanishes.
Another question is that are any two waves fully coherent if they have the same frequencies.
The answer is not necessary. Why? It is because the phases φ1 and φ2 can fluctuate. The
coherence implies that δ = φ1 − φ2 is a constant as time t and position x changes. In practical
situations, as the waves propagate, the environment provides noises to the phases. As a result,
the time-average becomes smaller. This process is called “decoherence” Typically, a system
gradually loses its coherence as t increases or traveled length x increases.
The interference involves all the cross-product terms ψ1∗ (t)ψ2 (t), ψ2∗ (t)ψ3 (t), ψ3∗ (t)ψ4 (t), and so
on. To deal with a system containing a large number of waves (particles), it is more convenient
to use a statistical tool than listing all the states. The idea is to use probabilities to describe
distributions of states. Such a tool in quantum mechanics is called the density operator or
matrix. Basically, a density matrix contains the information of the probability in each state.
Note 4: Ensemble
An ensemble is a statistical tool to describe a system of many particles. An ensemble
consists of a large number (ideally, infinite) of virtual copies of a particle. Each copy
represents a possible state that a particle can be in. A specific ensemble is specified
19
by assigning the probability in each state. For example, a photon state |photoni is
decomposed as
where |Li (|Ri) is the left(right)-polarized state. In an ensemble, there are many photons.
Let pL and pR be the probabilities of the left-polarized state and the right-polarized state,
respectively, where
pL + pR = 1. (111)
The probabilities pL and pR specify the ensemble where there are infinite particles, and
each has the probability pL (pR ) in the state |Li (|Ri).
However, we can not use the following expression to describe an ensemble,
hhhh (
((((
|ensemblei = p |Li + p |Ri, (112)
hh( h ( (
( ( h L hhh
h R
(((( hh
since this expression is used for one single state. The mathematical tool to describe an
ensemble is the density matrix ρ̂,
where U † U = 1. The matrix element Uij can be obtained explicitly by multiplying hψj 0 | on the
both sides of Eq. (117),
Uij = hψj |ai i. (118)
20
The inverse transforms are
X
|ψi i = Uij† |aj i, (119)
j
X
hψi | = Uj 0 i haj 0 |. (120)
j0
where P is a diagonal matrix whose diagonal elements are pi . In the new bases |aj i, the off-
diagonal element ρjj 0 can be nonzero. Indeed, the off-diagonal element ρjj 0 is related to the
correlation between the two states |aj i and |aj 0 i.
21
Keep in mind that the matrix in Eq. (128) is written in the basis vector defined by
!
1
≡ |Li, (129)
0
!
0
≡ |Ri. (130)
1
Now we consider the new bases |Xi and |Y i (linear polarized states)
1
|Li = √ (|Xi − i|Y i) , (131)
2
1
|Ri = √ (|Xi + i|Y i) . (132)
2
The unitary transformation is
! !
|Li † |Xi
=U , (133)
|Ri |Y i
where
1 −i
√ √
U † = 12 i
2 , (134)
√ √
12 2
√1
√
U = i2 −i
2 . (135)
√ √
2 2
−i T
1
√1
! 1
√ 0 0 √2 √
ρ̂ = i2 −i
2 2 (136)
0 1 √1 i
√ √ √
2 2 2 2
1 −i
!
= 2
i
2
1 . (137)
2 2
Note that the matrix in Eq. (137) is written in the new basis (see Eq. (124)). For this matrix,
the column vectors (1, 0)T and (0, 1)T represent the new basis vectors |a1 i and |a2 i, which
in this case are
!
1
≡ |Xi, (138)
0
!
0
≡ |Y i. (139)
1
22
Exercise 8: Density Matrix
Consider an ensemble of the density matrix
1 3
ρ = |LihL| + |RihR|. (140)
4 4
Calculate the density matrix in the bases |Xi and |Y i.
Although we derive the ensemble average Eq. (145) in the |ψi i bases, the trace of a matrix is
independent of the bases. Thus, Eq. (145) is valid in any basis. This basis-free property is the
advantage of using a trace. One direct application is when A = 1,
X
Tr (ρ) = pi = 1, (146)
i
(a) ρ = ρ†
(b) Tr (ρ) = 1
(c) 0 < Tr ρ2 ≤ 1
Prove that the above properties are true in any set of bases.
23
Note 5: Pure and Mixed Ensemble
We start with the bases |ψi i, where ρ is diagonal. A pure ensemble is specified by pi = 1
of some state |ψi i and all other pj = 0 for j , i. The equivalent condition of a pure
ensemble is
Tr ρ2 = 1, (147)
which applies to a density matrix in any basis. The condition of a mixed ensemble is
Tr ρ2 < 1. (148)
One particle state is always a pure ensemble. One common mistake is to be confused by
the superposition of one particle and the mixed ensemble.3 Consider a one-particle state
(qubit) composed of the superposition of |0i and |1i.
One might think that this state has a density matrix ρ = |α|2 |0ih0| + |β|2 |1ih1|. But, this
is wrong! The correct density matrix is
ρ = |ψihψ| (150)
= (α|0i + β|1i) (α ∗ h0| + β ∗ h1|) (151)
= |α|2 |0ih0| + |β|2 |1ih1| + αβ ∗ |0ih1| + α ∗ β|1ih0| (152)
!
|α|2 αβ ∗
= ∗ , (153)
α β |β|2
where
!
1
≡ |0i, (154)
0
!
0
≡ |1i. (155)
1
It is possible to find the bases where ρ is diagonal, since ρ is a hermitian matrix. The
off-diagonal elements in Eq. (153) describe the correlations between the states |0i and |1i.
An example of a mixed ensemble of qubits is
where both |α|2 and |β|2 are nonzero. In this mixed ensemble, the off-diagonal elements
are zero. This means that there is no correlation between the states |0i and |1i.
24
If the number of bases is N , the most random mixed ensemble is
N
1X
ρMR = |ψi ihψi | (158)
N
i=1
1
= 1N ×N , (159)
N
where 1N ×N is the N -by-N identity. The off-diagonal elements of the ensemble ρMR are
always zero, i.e., there is not any correlation between the basis states. It can be shown
that an entangled system must be in a mixed ensemble. The extent of entanglement of
an ensemble is somewhat related to von Neumann entropy.
(a)
!
0.5 0
ρ=
0 0.5
(b)
cos2 θ
!
cos θ sin θ
ρ=
cos θ sin θ sin2 θ
(c)
!
cos2 θ cos θ sin θeiφ
ρ=
cos θ sin θe −iφ sin2 θ
(d)
1
!
cos2 θ 2 cos θ sin θ
ρ= 1 2
2 cos θ sin θ sin θ
3 In many places, mixed states are called instead of mixed ensemble, although the latter is properer.
25
7.2 Dynamics of Density Operators
First, the density operator is not an observable, so we can not use the Heisenberg’s picture to
obtain its dynamics. Let’s begin with a density matrix in the diagonal form,
X
ρ(t) = pi |ψi (t)ihψi (t)|, (160)
i
where the dynamics of the states can be obtianed with Schrödinger Picture
∂
i~ |ψ (t)i = H|ψi (t)i, (161)
∂t i
∂
−i~ hψi (t)| = hψi (t)|H. (162)
∂t
Using Eqs. (160), (161) and (162), we obtain
∂ρ(t) i
= [ρ(t), H] . (163)
∂t ~
This equation is known as the von Neumann equation
or quantum Liouville equation. Equa-
tion (163) describes a closed system where Tr ρ is a constant in time. This means that the
2
coherence of the system is not changed. How could a system have dissipation and decoherence?
When a system is open to the environment, the interaction between the system and the environ-
ment leads to dissipation and decoherence. The idea is to write H = Hsys + Henv and to derive
a equation only about the reduced density matrix
∂ρ(t)sys ih i
= ρ(t)sys , Hsys + environment terms, (164)
∂t ~
where the reduced density matrix is obtained by the partial trace
There is not a unique answer how to write the environment terms since that depends on what
kind of environment it is and the interaction. The discussions of the environment terms belong
to the subject “Open Quantum Systems”, which is beyond the scope of the note. We will adopt
the phenomenological methods later.
26
Example 4: Dynamics of a Two Level System
Let the unperturbed Hamiltonian be
!
Ec 0
H= , (166)
0 Ev
and write the density matrix in this basis
!
ρcc ρcv
ρ= . (167)
ρvc ρvv
Using the von Neumann equation, Eq. (163), we can obtain four first-order differential
equations. Two of them are redundant because ρcc + ρvv = 1 and ρcv = ρvc
∗ . We need only
two equations
∂
ρ = 0, (168)
∂t cc
∂ i
ρcv = ρcv (Ev − Ec ) , (169)
∂t ~
with the solutions
ρcc (t) = ρcc (0), (170)
ρvv (t) = ρvv (0), (171)
ρcv (t) = ρcv (0)e−iωcv t , (172)
(173)
with ωcv = Ec −E
~ . The populations ρcc and ρvv are unchanged in an unperturbed system.
v
The off-diagonal element ρcv has a constant amplitude and a linearly-growing phase in
time. This means that the coherence of the system is unchanged. In a realistic situation,
the system will be dephased. A phenomenological way to add the dephasing is to add
−γρcv in Eq. (169),
∂ i
ρcv = ρcv (Ev − Ec ) − γρcv , (174)
∂t ~
with the solution
ρcv (t) = ρcv (0)e−iωcv t−γt , (175)
(176)
and γ is called the dephasing rate. The time T2 = γ1 is the characteristic time of dephas-
ing. If the system is open, the probability of the top level ρcc can decay. This can be
described as
∂
ρ = −γr ρcc , (177)
∂t cc
where γr is the relaxation rate. The time T1 = 1
γr is the characteristic time of relaxation.
27
Exercise 12: Quantum Liouville Equation
Derive Eqs. (168) and (169).
References
[1] R. Shankar, Principles of Quantum Mechanics, 1994
[3] Michael A. Nielsen and Isaac L. Chuang, Quantum Computation and Quantum Information,
2000
28