QM Lecture Notes
QM Lecture Notes
QM Lecture Notes
Technion
Preface
3.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4. Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.1 Time Evolution Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.2 Time Independent Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.3 Example - Spin 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.4 Connection to Classical Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 73
4.5 Symmetric Ordering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.7 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 623
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
Q = (q1 , q2 , · · · , qN ) . (1.1)
Consider the case where the vector of coordinates takes the value Q1 at time
t1 and the value Q2 at a later time t2 > t1 , namely
Q (t1 ) = Q1 , (1.2)
Q (t2 ) = Q2 . (1.3)
The action S associated with the evolution of the system from time t1 to
time t2 is defined by
t2
S= dt L , (1.4)
t1
L = L Q, Q̇; t , (1.5)
where
and where overdot denotes time derivative. The time evolution of Q, in turn,
depends on the trajectory taken by the system from point Q1 at time t1
Chapter 1. Hamilton’s Formalism of Classical Physics
Q
Q2
Q1
t
t1 t2
Fig. 1.1. A trajectory taken by the system from point Q1 at time t1 to point Q2
at time t2 .
to point Q2 at time t2 (see Fig. 1.1). For a given trajectory Γ the time
dependency is denoted as
Q (t) = QΓ (t) . (1.7)
d ∂L ∂L
= , (1.8)
dt ∂ q̇n ∂qn
where n = 1, 2, · · · , N.
δS = dt δL
t1
t2 N N
∂L ∂L
= dt δqn + δ q̇n
n=1
∂qn n=1
∂ q̇n
t1
t2 N N
∂L ∂L d
= dt δqn + δqn .
n=1
∂qn n=1
∂ q̇n dt
t1
(1.10)
Integrating the second term by parts leads to
t2 N
∂L d ∂L
δS = dt − δqn
n=1
∂qn dt ∂ q̇n
t1
N t2
∂L
+ δqn .
n=1
∂ q̇n t1
(1.11)
The last term vanishes since
δS = 0 . (1.13)
This has to be satisfied for any δQ, therefore the following must hold
d ∂L ∂L
= . (1.14)
dt ∂ q̇n ∂qn
Q
Q2
Γ’
Q1
t
t1 t2
Fig. 1.2. The classical trajectory QΓ (t) and the trajectory QΓ ′ (t).
In what follows we will assume for simplicity that the kinetic energy T of
the system can be expressed as a function of the velocities Q̇ only (namely,
it does not explicitly depend on the coordinates Q). The components of the
generalized force Fn , where n = 1, 2, · · · , N , are derived from the potential
energy U of the system as follows
∂U d ∂U
Fn = − + . (1.15)
∂qn dt ∂ q̇n
When the potential energy can be expressed as a function of the coordinates
Q only (namely, when it is independent on the velocities Q̇), the system is
said to be conservative. For that case, the Lagrangian can be expressed in
terms of T and U as
L=T −U . (1.16)
Example 1.2.1. Consider a point particle having mass m moving in a one-
dimensional potential U (x). The Lagrangian is given by
mẋ2
L=T −U = − U (x) . (1.17)
2
From the Euler-Lagrange equation
d ∂L ∂L
= , (1.18)
dt ∂ ẋ ∂x
one finds that
∂U
mẍ = − . (1.19)
∂x
1.3 Hamiltonian
The set of Euler-Lagrange equations contains N second order differential
equations. In this section we derive an alternative and equivalent set of equa-
tions of motion, known as Hamilton-Jacobi equations, that contains twice the
number of equations, namely 2N, however, of first, instead of second, order.
Definition 1.3.1. The variable canonically conjugate to qn is defined by
∂L
pn = . (1.20)
∂ q̇n
Definition 1.3.2. The Hamiltonian of a physical system is a function of
the vector of coordinates Q, the vector of canonical conjugate variables P =
(p1 , p2 , · · · , pN ) and time, namely
H = H (Q, P ; t) , (1.21)
is defined by
N
H= pn q̇n − L , (1.22)
n=1
N
∂L
= (q̇n dpn − ṗn dqn ) − dt .
n=1
∂t
(1.25)
H =T +U , (1.32)
where T is the kinetic energy of the system and where U is the potential
energy.
N
H= pl q̇l − L
l=1
∂T
= q̇l − (T − U )
∂ q̇l
l
∂ q̇ ∂ q̇m
n
= αnm q̇m + q̇n q̇l − T + U
∂ q̇l ∂ q̇l
l,n,m
δnl δ ml
T
= T +U .
(1.34)
The Poisson’s brackets are employed for writing an equation of motion for a
general physical quantity of interest, as the following theorem shows.
N
dF ∂F ∂F ∂F
= q̇n + ṗn +
dt n=1
∂qn ∂pn ∂t
N
∂F ∂H ∂F ∂H ∂F
= − +
n=1
∂qn ∂pn ∂pn ∂qn ∂t
∂F
= {F, H} + .
∂t
(1.39)
Corollary 1.4.1. If F does not explicitly depend on time, namely if ∂F/∂t =
0, and if {F, H} = 0, then F is a constant of the motion, namely
dF
=0. (1.40)
dt
1.5 Problems
1. Consider a particle having charge q and mass m in electromagnetic field
characterized by the scalar potential ϕ and the vector potential A. The
electric field E and the magnetic field B are given by
1 ∂A
E = −∇ϕ − , (1.41)
c ∂t
and
B=∇×A. (1.42)
Let r = (x, y, z) be the Cartesian coordinates of the particle.
a) Verify that the Lagrangian of the system can be chosen to be given
by
1 2 q
L= mṙ − qϕ + A · ṙ , (1.43)
2 c
by showing that the corresponding Euler-Lagrange equations are
equivalent to Newton’s 2nd law (i.e., F = mr̈).
b) Show that the Hamilton-Jacobi equations are equivalent to Newton’s
2nd law.
c) Gauge transformation — The electromagnetic field is invariant un-
der the gauge transformation of the scalar and vector potentials
A → A + ∇χ , (1.44)
1 ∂χ
ϕ → ϕ− (1.45)
c ∂t
where χ = χ (r, t) is an arbitrary smooth and continuous function
of r and t. What effect does this gauge transformation have on the
Lagrangian and Hamiltonian? Is the motion affected?
L C
1.6 Solutions
d ∂L ∂L
= , (1.55)
dt ∂ ẋ ∂x
where
d ∂L q ∂Ax ∂Ax ∂Ax ∂Ax
= mẍ + + ẋ + ẏ + ż , (1.56)
dt ∂ ẋ c ∂t ∂x ∂y ∂z
and
∂L ∂ϕ q ∂Ax ∂Ay ∂Az
= −q + ẋ + ẏ + ż , (1.57)
∂x ∂x c ∂x ∂x ∂x
thus
∂ϕ q ∂Ax
mẍ = −q −
∂x c ∂t
qEx
q ∂Ay ∂Ax ∂Ax ∂Az
+ ẏ − −ż − ,
c ∂x ∂y ∂z ∂x
(∇ ×A)z (∇ ×A)y
(ṙ×(∇×A))x
(1.58)
or
q
mẍ = qEx + (ṙ × B)x . (1.59)
c
Similar equations are obtained for ÿ and z̈ in the same way. These 3
equations can be written in a vector form as
1
mr̈ = q E + ṙ × B . (1.60)
c
H = p · ṙ − L
1 q
= ṙ · p − mṙ − A + qϕ
2 c
1 2
= mṙ + qϕ
2
2
p− qc A
= + qϕ .
2m
(1.62)
The Hamilton-Jacobi equation for the coordinate x is given by
∂H
ẋ = , (1.63)
∂px
thus
px − qc Ax
ẋ = , (1.64)
m
or
q
px = mẋ+ Ax . (1.65)
c
The Hamilton-Jacobi equation for the canonically conjugate variable
px is given by
∂H
ṗx = − , (1.66)
∂x
where
q ∂Ax ∂Ax ∂Ax q ∂Ax
ṗx = mẍ+ ẋ + ẏ + ż + , (1.67)
c ∂x ∂y ∂z c ∂t
and
∂H q px − qc Ax ∂Ax py − qc Ay ∂Ay pz − qc Az ∂Az ∂ϕ
− = + + −q
∂x c m ∂x m ∂x m ∂x ∂x
q ∂Ax ∂Ay ∂Az ∂ϕ
= ẋ + ẏ + ż −q ,
c ∂x ∂x ∂x ∂x
(1.68)
thus
c) Clearly, the fields E and B, which are given by Eqs. (1.41) and (1.42)
respectively, are unchanged since
∂χ ∂ (∇χ)
∇ − =0, (1.70)
∂t ∂t
and
∇ × (∇χ) = 0 . (1.71)
Thus, even though both L and H are modified, the motion, which
depends on E and B only, is unaffected.
2. The kinetic energy in this case T = Lq̇ 2 /2 is the energy stored in the
inductor, and the potential energy U = q 2 /2C is the energy stored in the
capacitor.
a) The Lagrangian is given by
Lq̇ 2 q2
L=T −U = − . (1.72)
2 2C
The Euler-Lagrange equation for the coordinate q is given by
d ∂L ∂L
= , (1.73)
dt ∂ q̇ ∂q
thus
q
Lq̈ + =0. (1.74)
C
This equation expresses the requirement that the voltage across the
capacitor is the same as the one across the inductor.
b) The canonical conjugate momentum is given by
∂L
p= = Lq̇ , (1.75)
∂ q̇
and the Hamiltonian is given by
p2 q2
H = pq̇ − L = + . (1.76)
2L 2C
Hamilton-Jacobi equations read
p
q̇ = (1.77)
L
q
ṗ = − , (1.78)
C
thus
q
Lq̈ + =0. (1.79)
C
d ∂L ∂L d ∂L ∂xb
− =
dt ∂ q̇a ∂qa dt ∂ ẋb ∂qa
∂L ∂xb ∂L ∂ ẋb
− −
∂xb ∂qa ∂ ẋb ∂qa
d ∂L ∂L ∂xb
= −
dt ∂ ẋb ∂xb ∂qa
d ∂xb ∂ ẋb ∂L
+ − .
dt ∂qa ∂qa ∂ ẋb
(1.86)
As can be seen from Eq. (1.82), the second term vanishes since
∂ ẋb ∂ 2 xb ∂ 2 xb d ∂xb
= q̇c + = ,
∂qa ∂qa ∂qc ∂t∂qa dt ∂qa
thus
d ∂L ∂L d ∂L ∂L ∂xb
− = − . (1.87)
dt ∂ q̇a ∂qa dt ∂ ẋb ∂xb ∂qa
p2
H= + V (r) . (1.89)
2m
Using
{xi , pj } = δ ij , (1.90)
Lz = xpy − ypx , (1.91)
one finds that
p2 , Lz = p2x , Lz + p2y , Lz + p2z , Lz
= p2x , xpy − p2y , ypx
= −2px py + 2py px
=0,
(1.92)
and
r2 , Lz = x2 , Lz + y 2 , Lz + z 2 , Lz
= −y x2 , px + y 2 , py x
=0.
(1.93)
Thus f r2 , Lz = 0 for arbitrary smooth function f r2 , and con-
sequently {H, Lz } = 0. In a similar way one can show that {H, Lx } =
{H, Ly } = 0, and therefore H, L2 = 0.
β |α ∈ C , (2.2)
∗
β |α = α |β , (2.3)
α| (c1 |β 1 + c2 |β 2 ) = c1 α |β 1 + c2 α |β 2 , where c1 , c2 ∈ C , (2.4)
α |α ∈ R and α |α ≥ 0. Equality holds iff |α = 0 . (2.5)
Note that the asterisk in Eq. (2.3) denotes complex conjugate. Below we list
some important definitions and comments regarding inner product:
• The real number α |α is called the norm of the vector |α ∈ F.
• A normalized vector has a unity norm, namely α |α = 1.
• Every nonzero vector 0 = |α ∈ F can be normalized using the transfor-
mation
|α
|α → . (2.6)
α |α
φm |φn = δ nm . (2.7)
|α = cn |φn , (2.8)
n
where cn ∈ C.
• By evaluating the inner product φm |α , where |α is given by Eq. (2.8)
one finds with the help of Eq. (2.7) and property (2.4) of inner products
that
! "
φm |α = φm cn |φn = cn φm |φn = cm . (2.9)
n n
=δnm
2.2 Operators
Operators, as the definition below states, are function from F to F:
Definition 2.2.1. An operator A : F → F on a vector space maps vectors
onto vectors, namely A |α ∈ F for every |α ∈ F.
X |α = Y |α . (2.11)
• Operators can be added, and the addition is both, commutative and asso-
ciative, namely
X +Y = Y +X , (2.12)
X + (Y + Z) = (X + Y ) + Z . (2.13)
• An operator A : F → F is said to be linear if
A (c1 |γ 1 + c2 |γ 2 ) = c1 A |γ 1 + c2 A |γ 2 (2.14)
XY |α = X (Y |α ) (2.15)
for any |α ∈ F.
• Operator multiplication is associative
X (Y Z) = (XY ) Z = XY Z . (2.16)
XY = Y X . (2.17)
β |α = β| |α . (2.18)
bra ket
Aαβ = |α β| . (2.19)
The outer product Aαβ is clearly an operator since for any |γ ∈ F the object
Aαβ |γ is a ket-vector
Since the above identity holds for any |α ∈ F one concludes that the quantity
in brackets is the identity operator, which is denoted as 1, namely
1= |φn φn | . (2.23)
n
This result, which is called the closure relation, implies that any complete
orthonormal basis can be used to express the identity operator.
β| (c1 |γ 1 + c2 |γ 2 ) = c1 β |γ 1 + c2 β |γ 2 . (2.24)
α| ⇔ |α , (2.26)
Proof. The dual of |β is the bra-vector β|, whereas the dual of β| is found
using Eqs. (2.29) and (2.23), thus
∗
|β DD = β |φn |φn
n
= φn |β
= |φn φn |β
n
= |φn φn | |β
n
=1
= |β .
(2.30)
Claim. FDD = F for any F ∈ F ∗ , where FDD is the dual of the dual of F .
|α = cm |φm . (2.32)
m
Using the above expression for FDD and the linearity property one finds that
FDD |α = cm F (|φn ) φn |φm
n,m
δmn
= cn F (|φn )
n
! "
=F cn |φn
n
= F (|α ) ,
(2.33)
therefore, FDD = F .
• The inner product between the bra-vector β| and the ket-vector |α can
be written as
β |α = β| 1 |α
= β |φn φn | α
n
φ1 |α
= β |φ1 β |φ2 · · · φ2 |α .
..
.
(2.34)
Thus, the inner product can be viewed as a product between the row vector
˙ β |φ1
β| = β |φ2 · · · , (2.35)
which is the matrix representation of the bra-vector β|, and the column
vector
φ1 |α
|α =˙ φ2 |α , (2.36)
..
.
which is the matrix representation of the ket-vector |α . Obviously, both
representations are basis dependent.
• Multiplying the relation |γ = X |α from the left by the basis bra-vector
φm | and employing again the closure relation (2.23) yields
φm |γ = φm | X |α = φm | X1 |α = φm | X |φn φn |α , (2.37)
n
or in matrix form
φ1 |γ φ1 | X |φ1 φ1 | X |φ2 · · · φ1 |α
φ2 |γ φ2 | X |φ1 φ2 | X |φ2 · · ·
= φ2 |α . (2.38)
.. .. .. ..
. . . .
In view of this expression, the matrix representation of the linear operator
X is given by
φ1 | X |φ1 φ1 | X |φ2 · · ·
X= ˙ φ2 | X |φ1 φ2 | X |φ2 · · · . (2.39)
.. ..
. .
Alternatively, the last result can be written as
Xnm = φn | X |φm , (2.40)
where Xnm is the element in row n and column m of the matrix represen-
tation of the operator X.
• Such matrix representation of linear operators can be useful also for mul-
tiplying linear operators. The matrix elements of the product Z = XY are
given by
2.6 Observables
Measurable physical variables are represented in quantum mechanics by Her-
mitian operators.
α| X † ⇔ X |α . (2.43)
†
Claim. X † =X
XY |α = X (Y |α ) ⇔ α| Y † X † = α| Y † X † , (2.46)
thus
(XY )† = Y † X † . (2.47)
Claim. If X = |β α| then X † = |α β|
X |γ = (|β α|) |γ = |β ( α |γ ) ⇔ ( α |γ )∗ β| = γ |α β| = γ| X † ,
(2.48)
where X † = |α β|.
gn = dim Fn . (2.51)
a2 | A = a∗2 a2 | , (2.54)
The first part of the theorem is proven by employing the last result (2.57)
for the case where |a1 = |a2 . Since |a1 is assumed to be a nonzero ket-
vector one concludes that a1 = a∗1 , namely a1 is real. Since this is true for
any eigenvalue of A, one can rewrite Eq. (2.57) as
The second part of the theorem is proven by considering the case where
a1 = a2 , for which the above result (2.58) can hold only if a2 |a1 = 0.
Namely eigenvectors corresponding to different eigenvalues are orthogonal.
Moreover, it is easy to show using the orthonormality relation (2.59) that the
following holds
Pn Pm = Pm Pn = Pn δ nm . (2.63)
For linear vector spaces of finite dimensionality, it can be shown that the
set {|an,i }n,i forms a complete orthonormal basis of eigenvectors of a given
Hermitian operator A. The generalization of this result for the case of ar-
bitrary dimensionality is nontrivial, since generally such a set needs not be
complete. On the other hand, it can be shown that if a given Hermitian oper-
ator A satisfies some conditions (e.g., A needs to be completely continuous)
then completeness is guarantied. For all Hermitian operators of interest for
this course we will assume that all such conditions are satisfied. That is, for
any such Hermitian operator A there exists a set of ket vectors {|an,i }, such
that:
1. The set is orthonormal, namely
where an ∈ R.
3. The set is complete, namely closure relation [see also Eq. (2.23)] is satis-
fied
gn
1= |an,i an,i | = Pn , (2.66)
n i=1 n
where
gn
Pn = |an,i an,i | (2.67)
i=1
that is
A= an Pn . (2.69)
n
The last result is very useful when dealing with a function f (A) of the
operator A. The meaning of a function of an operator can be understood in
terms of the Taylor expansion of the function
f (x) = fm xm , (2.70)
m
where
1 dm f
fm = . (2.71)
m! dxm
With the help of Eqs. (2.63) and (2.69) one finds that
f (A) = fm Am
m
! "m
= fm an Pn
m n
= fm am
n Pn
m n
= fm am
n Pn ,
n m
f (an )
(2.72)
thus
f (am ) = δ mn . (2.76)
Even when the state vector |α is given, quantum mechanics does not gener-
ally provide a deterministic answer to the question: what will be the outcome
of the measurement. Instead it predicts that:
1. The possible results of the measurement are the eigenvalues {an } of the
operator A.
2. The probability pn to measure the eigenvalue an is given by
gn
pn = α| Pn |α = | an,i |α |2 . (2.81)
i=1
Pn |α
|α → . (2.82)
α| Pn |α
We also note that a direct consequence of the collapse postulate is that two
subsequent measurements of the same observable performed one immediately
after the other will yield the same result. It is also important to note that the
above ’standard textbook description’ of the measurement process is highly
controversial, especially, the collapse postulate. However, a thorough discus-
sion of this issue is beyond the scope of this course.
Quantum mechanics cannot generally predict the outcome of a specific
measurement of an observable A, however it can predict the average, namely
the expectation value, which is denoted as A . The expectation value is easily
calculated with the help of Eq. (2.69)
A = an pn = an α| Pn |α = α| A |α . (2.84)
n n
On the other hand, as was first shown by Dirac, the gyromagnetic ratio
for the case of spin angular momentum takes twice this value
2µB
µspin = S. (2.90)
Note that we follow here the convention of using the letter L for orbital
angular momentum and the letter S for spin angular momentum.
The Stern-Gerlach apparatus allows measuring any component of the
magnetic moment vector. Alternatively, in view of relation (2.90), it can be
said that any component of the spin angular momentum S can be measured.
The experiment shows that the only two possible results of such a measure-
ment are + /2 and − /2. As we have seen above, one can construct a com-
plete orthonormal basis to the vector space made of eigenvectors of any given
observable. Choosing the observable Sz = S · ẑ for this purpose we construct
a basis made of two vectors {|+; ẑ , |−; ẑ }. Both vectors are eigenvectors of
Sz
|+ +| + |− −| = 1 . (2.97)
|α = |+ + |α + |− − |α . (2.98)
The closure relation (2.97) and Eqs. (2.91) and (2.92) yield
Exercise 2.8.1. Given that the state vector of a spin 1/2 is |+; ẑ calculate
(a) the expectation values Sx and Sz (b) the probability to obtain a value
of + /2 in a measurement of Sx .
01 1
Sx = +| Sx |+ = 10 =0, (2.111)
2 10 0
1 0 1
Sz = +| Sz |+ = 10 = . (2.112)
2 0 −1 0 2
(b) First, the eigenvectors of the operator Sx are found by solving the equa-
tion Sx |α = λ |α , which is done by diagonalization of the matrix represen-
tation of Sx . The relation Sx |α = λ |α for the two eigenvectors is written
in a matrix form as
! " ! "
01 √1 √1
2 = 2 , (2.113)
2 10 √1 2 √12
2
! " ! "
01 √1 √1
2 =− 2 . (2.114)
2 10 − √12 2 − √12
That is, in ket notation
p+ = α| P+;x̂ |α , (2.118)
where the projection P+;x̂ is evaluated with the help of Eq. (2.79)
Sx − − 2
P+;x̂ = , (2.119)
2 − −2
and therefore
1 1 1
2 2 1
p+ = 1 0 1 1 = . (2.121)
2 2 0 2
U= |bn an | , (2.124)
n
transforms each of the basis vector |an to the corresponding basis vector |bn
and those of U † by
an | U † |am = bn |am .
|α = |an an |α , (2.128)
n
which can be represented as a column vector in the basis {|an }. The nth
element of such a column vector is an |α . The operator U can be employed
for finding the corresponding column vector representation of the same ket-
vector |α in the other basis {|bn }
2.10 Trace
= bl |bk bk | X |bl
k,l
δkl
= bk | X |bk .
k
(2.132)
Tr (XY ) = Tr (Y X) . (2.133)
Proof. With the help of the closure relation (2.23) one finds that
Tr (XY ) = an | XY |an
n
= an | X |am am | Y |an
n,m
= am | Y |an an | X |am
n,m
= am | Y X |am
m
= Tr (Y X) .
(2.134)
where
gn
Pn = |an,i an,i | . (2.148)
i=1
Now consider another observable B, which is assumed to commute with
A, namely [A, B] = 0.
Claim. The operator B has a block diagonal matrix in the basis {|an,i },
namely am,j | B |an,i = 0 for n = m.
Proof. Multiplying Eq. (2.145) from the left by am,j | B yields
am,j | BA |an,i = an am,j | B |an,i . (2.149)
On the other hand, since [A, B] = 0 one has
am,j | BA |an,i = am,j | AB |an,i = am am,j | B |an,i , (2.150)
thus
(an − am ) am,j | B |an,i = 0 . (2.151)
For a given n, the gn × gn matrix an,i′ | B |an,i is Hermitian, namely
an,i′ | B |an,i = an,i | B |an,i′ ∗ . Thus, there exists a unitary transformation
Un , which maps Fn onto Fn , and which diagonalizes the block of B in the
subspace Fn . Since Fn is an eigensubspace of A, the block matrix of A in the
new basis remains diagonal (with the eigenvalue an ). Thus, we conclude that
a complete and orthonormal basis of common eigenvectors of both operators
A and B exists. For such a basis, which is denoted as {|n, m }, the following
holds
A |n, m = an |n, m , (2.152)
B |n, m = bm |n, m . (2.153)
where
A = α| A |α , (2.155)
) 2*
A = α| A2 |α . (2.156)
thus
1 1) *
∆A∆B = [∆A, ∆B] + [∆A, ∆B]+ , (2.168)
2 2
∈I ∈R
and consequently
1 1 ) * 2
| ∆A∆B |2 = | [∆A, ∆B] |2 + [∆A, ∆B]+ . (2.169)
4 4
Finally, with the help of the identity [∆A, ∆B] = [A, B] one finds that
' (' ( 1
(∆A)2 (∆B)2 ≥ | [A, B] |2 . (2.170)
4
2.14 Problems
| u |v | ≤ u |u v |v , (2.171)
| u| A |v | ≤ u| A |u v| A |v . (2.173)
A = |α α| − |β β| . (2.182)
H = H0 + W . (2.183)
The eigenstates of H0 are the six states |ϕn , with the same eigenvalue
E0 . The operator W is described by
W |ϕ1 = −a |ϕ2 − a |ϕ6 ,
W |ϕ2 = −a |ϕ3 − a |ϕ1 ,
..
.
W |ϕ6 = −a |ϕ1 − a |ϕ5 .
(2.184)
Find the eigenvalues and eigen vectors of H. Clue: Consider a solution of
the form
6
|k = eikn |ϕn . (2.185)
n=1
2.15 Solutions
1. Let
|γ = |u + λ |v , (2.186)
u |u + λ u |v + λ∗ v |u + |λ|2 v |v ≥ 0 . (2.187)
By choosing
v |u
λ=− , (2.188)
v |v
one has
2
v |u u |v v |u
u |u − u |v − v |u + v |v ≥ 0 , (2.189)
v |v v |v v |v
thus
| u |v | ≤ u |u v |v . (2.190)
one has F = F † .
4. In general, (|β α|)† = |α β|, thus clearly the operator A is Hermitian.
Moreover it is positive-definite since for every |u the following holds
u| A |u = u |a a |u = | a |u |2 ≥ 0 . (2.194)
5. Let
v| A |u
|γ = |u − |v .
v| A |v
| u| A |v | ≤ u| A |u v| A |v . (2.196)
Note that this result allows easy proof of the following: Under the same
conditions (namely, A is a Hermitian positive-definite operator) Tr (A) =
0 if and only if A = 0.
6. The expansion is given by
−1
(A − λB)−1 = A 1 − λA−1 B
−1
= 1 − λA−1 B A−1
2 3
= 1 + λA−1 B + λA−1 B + λA−1 B + · · · A−1 .
(2.197)
7. By definition:
d A (λ + ǫ) B (λ + ǫ) − A (λ) B (λ)
(AB) = lim
dλ ǫ→0 ǫ
(A (λ + ǫ) − A (λ)) B (λ) A (λ + ǫ) (B (λ + ǫ) − B (λ))
= lim + lim
ǫ→0 ǫ ǫ→0 ǫ
dA dB
= B+A .
dλ dλ
(2.198)
8. Taking the derivative of both sides of the identity 1 = AA−1 on has
dA −1 dA−1
0= A +A , (2.199)
dλ dλ
thus
d dA −1
A−1 = −A−1 A . (2.200)
dλ dλ
|n n| = 1 . (2.201)
n
In this basis
! "
Tr (|u v|) = n |u v |n = v| |n n| |u = v |u . (2.202)
n n
†
10. The operator A† A is Hermitian since A† A = A† A, and positive-
definite since the norm of A |u is nonnegative for every |u , thus one
has u| A† A |u ≥ 0. Moreover, using a complete orthonormal basis {|n }
one has
Tr A† A = n| A† A |n
n
= n| A† |m m| A |n
n,m
= | m| A |n |2 .
n,m
(2.203)
11. Let {|b′ } be a complete orthonormal basis made of eigenvectors of B
(i.e., B |b′ = b′ |b′ ). Using this basis for evaluating the trace one has
12. Let f (s) = esL Ae−sL , where s is real. Using Taylor expansion one has
1 df 1 d2 f
f (1) = f (0) + + +··· , (2.205)
1! ds s=0 2! ds2 s=0
thus
1 df 1 d2 f
eL Ae−L = A + + +··· , (2.206)
1! ds s=0 2! ds2 s=0
where
df
= LesL Ae−sL − esL Ae−sL L = [L, f (s)] , (2.207)
ds
d2 f df
= L, = [L, [L, f (s)]] , (2.208)
ds2 ds
therefore
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · . (2.209)
2! 3!
13. The identity clearly holds for the case m = 1. Moreover, assuming it
holds for m, namely assuming that
one has
, m+1 -
A , B = A [Am , B] + [A, B] Am
= mAm [A, B] + [A, B] Am .
(2.211)
m
It is easy to show that if [[A, B] , A] = 0 then [[A, B] , A ] = 0, thus one
concludes that
, m+1 -
A , B = (m + 1) Am [A, B] . (2.212)
14. Define the function f (s) = esA esB , where s is real. The following holds
df
= AesA esB + esA BesB
ds
= A + esA Be−sA esA esB
s2
f (s) = e(A+B)s e[A,B] 2 . (2.215)
15. Define
, -
f (β) ≡ A, e−βH , (2.217)
β
−βH
g (β) ≡ e eλH [H, A] e−λH dλ . (2.218)
0
= n| X |m m| Y |n
n,m
= m| Y |n n| X |m
n,m
= m| Y X |m
m
= Tr (Y X) .
(2.221)
Note that using this result it is easy to show that Tr (U + XU ) = Tr (X)
, provided that U is a unitary operator.
17. Clearly A is Hermitian, namely A† = A, thus the two eigenvalues λ1
and λ2 are expected to be real. Since the trace of an operator is basis
independent, the following must hold
Tr (A) = λ1 + λ2 , (2.222)
and
On the other hand, with the help of Eq. (2.177) one finds that
and
Tr A2 = Tr (|α α |α α|) + Tr (|β β |β β|) − Tr (|α α |β β|) − Tr (|β β |α α|)
= 2 − α |β Tr (|α β|) − β |α Tr (|β α|)
= 2 1 − | α |β |2 ,
(2.225)
thus
.
2
λ± = ± 1 − | α |β | . (2.226)
|β = a |α + c |γ , (2.227)
α| A |α α| A |γ |c|2 −ac∗
A=
˙ = ≡ Â . (2.228)
γ| A |α γ| A |γ −a∗ c − |c|2
Thus,
Tr  = 0 , (2.229)
and
H |k = Em |k , (2.232)
where
6
|k = eikn |ϕn . (2.233)
n=1
thus
6 ( (
H |k = E0 |k − a eikn ϕ(n−1)′ + ϕ(n+1)′ = E |k , (2.234)
n=1
e6ik = 1 , (2.235)
or
mπ
km = , (2.236)
3
where m = 1, 2, · · · , 6. The corresponding eigen vectors are denoted as
6
|km = eikm n |ϕn , (2.237)
n=1
H |km = Em |k , (2.239)
where
Em = E0 − 2a cos km . (2.240)
[x, p] = i . (3.9)
i∆x p
J (∆x ) = exp − , (3.12)
where ∆x ∈ R.
xJ (∆x ) |x′ = ([x, J (∆x )] + J (∆x ) x) |x′ = (x′ + ∆x ) J (∆x ) |x′ , (3.18)
ψα (x′ ) = x′ |α . (3.20)
x′ | x |α = x′ x′ |α = x′ ψα (x′ ) , (3.21)
A (x) = an xn . (3.22)
n
ip∆x i∆x
J (−∆x ) = exp =1+ p + O (∆x )2 , (3.25)
thus
i∆x 2
x′ | J (−∆x ) |α = ψα (x′ ) + x′ | p |α + O (∆x ) . (3.26)
On the other hand, according to Eq. (3.24) also the following holds
ψα (x′ + ∆x ) − ψα (x′ )
x′ | p |α = −i + O (∆x ) . (3.28)
∆x
Thus, in the limit ∆x → 0 one has
dψα
x′ | p |α = −i . (3.29)
dx′
To mathematically understand the last result, consider the differential
operator
d
J˜ (−∆x ) = exp ∆x
dx
2
d 1 d
= 1 + ∆x + ∆x +··· .
dx 2! dx
(3.30)
In view of the Taylor expansion of an arbitrary function f (x)
df (∆x )2 d2 f
f (x0 + ∆x ) = f (x0 ) + ∆x + +···
dx 2! dx2
d
= exp ∆x f
dx x=x0
= J˜ (−∆x ) f ,
x=x0
(3.31)
one can argue that the operator J˜ (−∆x ) generates translation.
As we have pointed out above, the spectrum (i.e., the set of all eigen-
values) of x is continuous. On the other hand, in the discussion in chapter
2 only the case of an observable having discrete spectrum has been consid-
ered. Rigorous mathematical treatment of the case of continuous spectrum
is nontrivial mainly because typically the eigenvectors in such a case cannot
be normalized. However, under some conditions one can generalize some of
the results given in chapter 2 for the case of an observable having a continu-
ous spectrum. These generalization is demonstrated below for the case of the
position operator x:
1. The closure relation (2.23) is written in terms of the eigenvectors |x′ as
∞
β |α = ′
dx β |x ′ ′
x |α = dx′ ψ ∗β (x′ ) ψα (x′ ) . (3.34)
−∞ −∞
thus, as expected 12 = 1.
p |p′ = p′ |p . (3.44)
φα (p′ ) = p′ |α . (3.47)
β |α = ′
dp β |p ′ ′
p |α = dp′ φ∗β (p′ ) φα (p′ ) . (3.50)
−∞ −∞
What is the relation between the position wavefunction ψα (x′ ) and its mo-
mentum counterpart φα (p′ )?
1 ip′ x′
x′ |p′ = √ exp . (3.52)
2π
2πδ(x′ −x′′ )
(3.57)
Thus, by choosing N to be real one finds that
1 ip′ x′
x′ |p′ = √ exp . (3.58)
2π
The last result together with Eqs. (3.32) and (3.45) yield
/∞ ′ ip′ x′
∞ dp e φα (p′ )
−∞
ψα (x′ ) = x′ |α = dp′ x′ |p′ p′ |α = √ , (3.59)
2π
−∞
/∞ ip′ x′
∞ dx′ e− ψα (x′ )
−∞
φα (p′ ) = p′ |α = dx′ p′ |x′ ′
x |α = √ . (3.60)
2π
−∞
That is, transformations relating ψα (x′ ) and φα (p′ ) are the direct and inverse
Fourier transformations.
one has
*
qx |r′ = qx′ qx′ , qy′ , qz′ , (3.62)
*
qy |r′ = qy′ qx′ , qy′ , qz′ , (3.63)
*
qz |r′ = qz′ qx′ , qy′ , qz′ . (3.64)
The closure relation is written as
∞ ∞ ∞
one has
*
px |p′ = p′x p′x , p′y , p′z , (3.68)
*
py |p′ = p′y p′x , p′y , p′z , (3.69)
*
pz |p′ = p′z p′x , p′y , p′z . (3.70)
The closure relation is written as
∞ ∞ ∞
1 ip′ · r′
r′ |p′ = exp . (3.75)
(2π )3/2
3.4 Problems
1. Show that
dA
[p, A (x)] = −i , (3.76)
dx
dB
[x, B (p)] = i , (3.77)
dp
where A (x) is a differentiable function of x and B (p) is a differentiable
function of p.
2. Show that the mean value of x in a state described by the wavefunction
ψ (x), namely
+∞
1 x′2
ψα (x′ ) = √ exp ikx′ − 2 . (3.81)
π1/4 d 2d
Calculate
' (' (
a) (∆x)2 (∆p)2
b) p′ |α
6. Show that
1 ip′ · (r′ − r′′ )
3 d3 p′ exp = δ (r′ − r′′ ) . (3.84)
(2π )
O = p + Kx , (3.85)
where p is the momentum operator (i.e. the minimum value of the quan-
tity α| (p − O)2 |α is obtained when the operator O is chosen to be
pα ).
a) Calculate the matrix elements x′ | pα |x′′ of the operator pα in the
position representation.
b) The operator P is the difference between the ’true’ momentum op-
erator and pα
P = p − pα . (3.87)
' (
Calculate the variance (∆P)2 with respect to the state |α
' (
(∆P)2 = α| P 2 |α − α| P |α
2
. (3.88)
and where
' (
(∆p)2 = α| p2 |α − α| p |α 2
. (3.91)
3.5 Solutions
1. The commutator [x, p] = i is a constant, thus the relation (2.179) can
be employed
dxm
[p, xm ] = −i mxm−1 = −i , (3.92)
dx
m
dp
[x, pm ] = i mpm−1 = i . (3.93)
dp
This holds for any m, thus, for any differentiable function A (x) of x and
for any differentiable function B (p) of p one has
dA
[p, A (x)] = −i , (3.94)
dx
dB
[x, B (p)] = i . (3.95)
dp
2. The following holds
+∞
F (a) = dxψ∗ (x + a) x2 ψ (x + a)
−∞
+∞
2
= dx′ ψ∗ (x′ ) (x′ − a) ψ (x′)
−∞
' (
= (x − a)2
) *
= x2 − 2a x + a2 .
(3.96)
The requirement
dF
=0 (3.97)
da
+∞
) *
2
x = dx′ ψ∗α (x′ ) x′2 ψα (x′ )
−∞
+∞
1 x′2
= dx′ exp − x′2
π1/2 d d2
−∞
d3 π1/2
1
=
π1/2 d 2
d2
= ,
2
(3.100)
+∞
dψα
p = −i dx′ ψ∗α (x′ )
dx′
−∞
+∞
i x′2 x′
=− dx′ exp − ik −
π1/2 d d2 d2
−∞
i
=− ikdπ1/2
π1/2 d
= k,
(3.101)
+∞
) 2* d2 ψα
p = (−i )2 dx′ ψ∗α (x′ )
dx′2
−∞
+∞ !! ""
2
2 1 ′ x′2 x′ 1
= (−i ) dx exp − 2 ik − 2 − 2
π1/2 d d d d
−∞
1 1 √ 2d4 k2 + d2
= (−i )2 1/2 − d π
π d 2 d4
! "
2 1
= ( k) 1+ 2 ,
2 (dk)
(3.102)
a) thus
! ! " "
' (' ( d2 1 2
(∆x)2 (∆p)2 = ( k) 2
1+ 2 − ( k) 2
= .
2 2 (dk) 4
(3.103)
p |α = dx′ |x′ x′ | p |α
−∞
∞
d
= −i dx′ |x′ x′ |α ,
dx′
−∞
(3.106)
thus, since |α is an arbitrary ket vector, the following holds
∞
d
p = −i dx′ |x′ x′ | . (3.107)
dx′
−∞
6. With the help of Eqs. (3.66), (3.71) and (3.75) one finds that
δ (r′ − r′′ ) = r′ |r′′
= d3 p′ r′ |p′ p′ |r′′
dg i dg i dg
eg(x) pe−g(x) = p + i + g (x) , + g (x) , g (x) , +· · · .
dx 2! dx 3! dx
(3.111)
Kx2
g (x) = (3.112)
2i
yields
UpU † = p + Kx = O , (3.113)
and thus
α| (p − pα )2 |α
∞
2
d logψ α
= dx′ ρ (x′ ) i + φα (x′ ) .
dx′
−∞
(3.126)
where
2
ρ (x′ ) = |ψα (x′ )| . (3.127)
d log ψ
ψ∗
α
x′ | pα |x′′ = α
δ (x′ − x′′ ) . (3.129)
2i dx′
Note: Comparing this result with the expression for the current den-
sity J associated with the state |α [see Eq. (4.223)] yields the fol-
lowing relation
dψ
J= Im ψ∗α α′
m dx
′
ρ (x ) d logψα d logψ∗α
= −
m 2i dx′ dx′
′
ρ (x )
= φ (x′ ) .
m α
(3.130)
b) As can be seen from Eqs. (3.123) and (3.128) the following holds
∞
d 1 d log ψ
ψ
α
∗
P =i dx′ |x′ − ′ + α
x′ | , (3.131)
dx 2 dx′
−∞
hence ∞
dψα 1 dψα dψ∗α
α| P |α = i dx′ −ψ∗α + ψ∗α − ψ
dx′ 2 dx′ dx′ α
−∞
∞
i dψα dψ∗α
=− dx′ ψ∗α + ψ
2 dx′ dx′ α
−∞
∞
i dρ (x′ )
=− dx′
2 dx′
−∞
=0,
(3.132)
thus'[see Eqs.
( (3.126) and (3.128)]
2
(∆P) = α| P 2 |α
∞
2 2
d logψα d logψ∗α
= dx′ ρ (x′ ) +
2 dx′ dx′
−∞
∞
2 2
d logρ (x′ )
= dx′ ρ (x′ ) .
2 dx′
−∞
(3.133)
Note that the result α| P |α = 0 implies that pα and p have the
same expectation value, i.e. α| pα |α = α| p |α . On the other hand,
thus
α| ppα + pα p − 2p2α |α
α|p x′
∞ | x′ |p|α
ψ∗ + ψα d logψα d logψα
2 ′
= dx ρ α
− 2 Im Im
dx′ dx′
−∞
=0,
(3.136)
and 'therefore
( ' (
(∆p)2 = α| P 2 |α + (∆pα )2
≥ α| P 2 |α
∞
2 2
d logρ (x′ )
= dx′ ρ (x′ ) .
2 dx′
−∞
(3.137)
For general real functions f (x′ ) , g (x′ ) : R → R the Schwartz in-
equality (2.171) implies that
∞ 2 ∞ ∞
′ ′ ′ ′ ′ 2 2
dx f (x ) g (x ) ≤ dx (f (x )) dx′ (g (x′ )) . (3.138)
−∞ −∞ −∞
x = dx′ ρ (x′ ) x′
−∞
2
/∞ ′ ′ ′ d logρ(x′ )
∞
2
dx ρ (x ) (x − x ) dx′
′ ′ d logρ (x′ ) −∞
dx ρ (x ) ≥ ' ( ,
dx′ (∆x)2
−∞
(3.141)
where
∞
' (
2 2
(∆x) = dx′ ρ (x′ ) (x′ − x ) (3.142)
−∞
=− dx′ ρ (x′ )
−∞
= −1 .
(3.143)
Combining these results [see Eqs. (3.137) and (3.141)] lead to
' (' ( 2
(∆x)2 (∆p)2 ≥ . (3.144)
2
Claim. The time evolution operator satisfies the Schrödinger equation (4.1).
Proof. Expressing the Schrödinger equation (4.1) in terms of Eq. (4.4)
d
i u (t, t0 ) |α (t0 ) = Hu (t, t0 ) |α (t0 ) , (4.5)
dt
and noting that |α (t0 ) is t independent yield
Chapter 4. Quantum Dynamics
d
i u (t, t0 ) |α (t0 ) = Hu (t, t0 ) |α (t0 ) . (4.6)
dt
du (t, t0 )
i = Hu (t, t0 ) . (4.7)
dt
This result leads to the following conclusion:
iH (t − t0 )
u (t, t0 ) = exp − . (4.9)
The operator u (t, t0 ) takes a relatively simple form in the basis of eigenvectors
of the Hamiltonian H. Denoting these eigenvectors as |an,i , where the index
i is added to account for possible degeneracy, and denoting the corresponding
eigenenergies as En one has
where
iEn (t − t0 )
|α (t) = exp − |α (t0 ) . (4.15)
However, the phase factor multiplying |α (t0 ) has no effect on any mea-
surable physical quantity of the system, that is, the system’s properties are
time independent. This is why the eigenvectors of the Hamiltonian are called
stationary states.
U = −µ · B . (4.16)
is the Bohr’s magneton (note that the electron charge is taken to be negative
e < 0). Based on these relations we hypothesize that the Hamiltonian of a
spin 1/2 in a magnetic field B is given by
e
H=− S·B. (4.19)
me c
Assume the case where
B = Bẑ , (4.20)
where B is a constant. For this case the Hamiltonian is given by
H = ωSz , (4.21)
where
|e| B
ω= (4.22)
me c
is the so-called Larmor frequency. In terms of the eigenvectors of the operator
Sz
Sz |± = ± |± , (4.23)
2
where the compact notation |± stands for the states |±; ẑ , one has
ω
H |± = ± |± , (4.24)
2
namely the states |± are eigenstates of the Hamiltonian. Equation (4.13) for
the present case reads
iωt iωt
u (t, 0) = e− 2 |+ +| + e 2 |− −| . (4.25)
Exercise 4.3.1. Consider spin 1/2 in magnetic field given by B = Bẑ, where
B is a constant. Given that |α (0) = |+; x̂ at time t = 0 calculate (a) the
probability p± (t) to measure Sx = ± /2 at time t; (b) the expectation value
Sx (t) at time t.
Solution 4.3.1. Recall that [see Eq. (2.102)]
1
|±; x̂ = √ (|+ ± |− ) (4.26)
2
(a) Using Eq. (4.25) one finds
p± (t) = | ±; x̂| u (t, 0) |α (0) |2
2
1 iωt iωt
= ( +| ± −|) e− 2 |+ +| + e 2 |− −| (|+ + |− )
2
2
1 − iωt iωt
= e 2 ±e 2 ,
2
(4.27)
thus
ωt
p+ (t) = cos2 , (4.28)
2
ωt
p− (t) = sin2 . (4.29)
2
(b)Using the results for p+ and p− one has
Sx = (p+ − p− )
2
ωt ωt
= cos2 − sin2
2 2 2
= cos (ωt) .
2
(4.30)
dA(H) du† du ∂A
= Au + u† A + u† u
dt dt dt ∂t
1 ∂A
= −u† HAu + u† AHu + u† u
i ∂t
1 ∂A
= −u† Huu† Au + u† Auu† Hu + u† u
i ∂t
1 ∂A(H)
= −H(H) A(H) + A(H) H(H) + .
i ∂t
(4.36)
Thus, we have found that
We see that the Poisson’s brackets in the classical equation of motion (4.31)
for the classical variable A(c) are replaced by a commutation relation in the
quantum counterpart equation of motion (4.38) for the expectation value A
1
{, } → [, ] . (4.39)
i
Note that for the case where the Hamiltonian is time independent, namely
for the case where the time evolution operator is given by Eq. (4.9), u com-
mutes with H, namely [u, H] = 0, and consequently
H(H) = u† Hu = H . (4.40)
1
{, } → [, ] , (4.41)
i
namely
1 2
x(c) , p(c) = 1 → [x, p] = i . (4.42)
implies that
1
e ξ(x −x) dξ = δ x(c) − x
i (c)
, (4.48)
2π
1 i
η(p(c) −p)
e dη = δ p(c) − p . (4.49)
2π
At first glance these relations may lead to the (wrong) conclusion that the
term Υ equals to δ x(c) − x δ p(c) − p , however, this is incorrect since x
and p are non-commuting operators.
4.6 Problems
1. Consider spin 1/2 in magnetic field given by B = Bẑ, where B is a
constant. At time t = 0 the system is in the state |+; x̂ . Calculate Sx ,
Sy and Sz as a function of time t.
2. The dynamics of a given system is governed by the Hamiltonian H, which
is assumed to be time independent. The state' ( system |ψ0 and
of the
2
the variance of the Hamiltonian operator (∆H) at time t = 0 are
given. The observable P = |ψ0 ψ0 | is measured at time t. Calculate the
expectation
' ( value P to second order in t and express the result in terms
2
of (∆H) .
3. Consider a point particle having mass m moving in one dimension under
the influence of the potential V (x). Let |ψn be a normalized eigenvector
of the Hamiltonian of the system with eigenvalue En . Show that the cor-
responding wavefunction ψn (x′ ) in the coordinate representation satisfies
the following equation
2
d2 ψn (x′ )
− + V (x′ ) ψ n (x′ ) = En ψn (x′ ) . (4.50)
2m dx′2
4. Consider the Hamiltonian operator
p2
H= + V (r) , (4.51)
2m
where r = (x, y, z) is the vector of position operators, p = (px , py , pz )
is the vector of canonical conjugate operators, and the mass m is a con-
stant. Let |ψn be a normalizable eigenvector of the Hamiltonian H with
eigenvalue En . Show that
ψn | p |ψn = 0 . (4.52)
5. Show that in the p representation the Schrödinger equation
d |α
i = H |α , (4.53)
dt
where H is the Hamiltonian
p2
H= + V (r) , (4.54)
2m
can be transformed into the integro-differential equation
d p2
i φα = φ + dp′ U (p − p′ ) φα , (4.55)
dt 2m α
where φα = φα (p′ , t) = p′ |α is the momentum wave function and
where
i
U (p) = (2π )−3 dr V (r) exp − p · r . (4.56)
d2 ψi (xi ) 2m
+ 2 [Ei − Vi (xi )] ψi (xi ) = 0 , (4.58)
dx2i
d2 ψ (x) 2m
+ 2 (E − V (x)) ψ (x) = 0 (4.59)
dx2
if the potential energy is an even function of x , namely V (x) = V (−x).
11. Show that the first derivative of the time-independent wavefunction is
continuous even at points where V (x) has a finite discontinuity.
12. A particle having mass m is confined by a one dimensional potential given
by
+
−W if |x| ≤ a
Vs (x) = , (4.60)
0 if |x| > a
where a > 0 and W > 0 are real constants. Show that the particle has
at least one bound state (i.e., a state having energy E < 0).
13. Consider a particle having mass m confined in a potential well given by
+
0 if 0 ≤ x ≤ a
V (x) = . (4.61)
∞ if x < 0 or x > a
where δ (x) is the delta function. The value of the parameter α suddenly
changes from α1 at times t < 0 to the value α2 at times t > 0. Both
α1 and α2 are positive real numbers. Given that the particle was in the
ground state at times t < 0, what is the probability p that the particle
will remain bounded at t > 0?
15. Consider a point particle having mass m in a one dimensional potential
given by
where δ (x) is the delta function, and where α > 0. Let |γ 0 be the
ground state and let E0 be the energy of the ground state. The particle
is prepared in the state
ip0 x
|g (p0 ) = exp |γ 0 , (4.65)
where δ (x) is the delta function and α is a constant. Let E0 be the energy
of the ground state. Under what conditions E0 < 0?
17. The same as the previous exercise, however for the potential
+
∞ x<0
V (x) = , (4.67)
−αδ (x − x0 ) x ≥ 0
where the sum is taken over all energy eigen-states of the particle (where
H |k = Ek |k ), and x is the x component of the position vector operator
r (the Thomas-Reiche-Kuhn sum rule).
19. A particle having mass m is confined in a one dimensional potential well
given by
+
0 0<x<a
V (x) = .
∞ else
a) At time t = 0 the position was measured and the result was x = a/2.
The resolution of the position measurement is ∆x , where ∆x ≪ a.
After time τ 1 the energy was measured. Calculate the probability pn
to measure that the energy of the system is En , where En are the
eigenenergies of the particle in the well, and where n = 1, 2, · · · .
b) Assume that the result of the measurement in the previous section
was E2 . At a later time τ 2 > τ 1 the momentum p of the particle
was measured. Calculate the expectation value p .
20. A particle having mass m is in the ground state of an infinite potential
well of width a, which is given by
+
0 0<x<a
V1 (x) = . (4.70)
∞ else
At time t = 0 the potential suddenly changes and becomes
+
0 0 < x < 2a
V2 (x) = , (4.71)
∞ else
namely the width suddenly becomes 2a. (a) Find the probability p to
find the particle in the ground state of the new well. (b) Calculate the
expectation value of the energy H before and after the change in the
potential. ' (
21. Calculate the uncertainties in position (∆x)2 and in momentum
' (
(∆p)2 of the energy eigenstates of a particle having mass m, which
is confined in a one dimensional potential well given by
+
0 0<x<a
V (x) = . (4.72)
∞ else
ρ = ψψ∗ (4.74)
qρ
J= Im (ψ∗ ∇ψ) − A (4.75)
m mc
is the current density, and A is the electromagnetic vector potential.
23. A particle having mass m moves in one dimension under the influence of
the potential V (x′ ). In the range |x′ | > a the potential V (x′ ) vanishes.
Consider a solution to the time independent Schrödinger equation, whose
wavefunction ψ (x′ ) in the range |x′ | > a is taken to be given by
+ ′ ′
′ A1 eikx + B1 e−ikx x′ < −a
ψ (x ) = ′ ′ , (4.76)
A2 eikx + B2 e−ikx x′ > a
H |ψ = E |ψ . (4.79)
|ψ1 = P |ψ , (4.81)
where
H11 = P HP , (4.83)
H22 = QHQ , (4.84)
H12 = P HQ , (4.85)
H21 = QHP , (4.86)
and where
Q=1−P . (4.87)
iHt
u = exp − . (4.89)
N
where UI is given by
N
UI = U † (Uu)N U N . (4.92)
Derive a Schrödinger equation for the operator UI having the form [see
Eq. (4.7)]
dUI
i = Heff UI , (4.93)
dt
and find an expression for the effective Hamiltonian Heff valid in the limit
N → ∞.
4.7 Solutions
1. The operators Sx , Sy and Sz are given by Eqs. (2.102), (2.103) and (2.99)
respectively. The Hamiltonian is given by Eq. (4.21). Using Eqs. (4.38)
and (2.136) one has
d Sx ω
= [Sx , Sz ] = −ω Sy , (4.94)
dt i
d Sy ω
= [Sy , Sz ] = ω Sx , (4.95)
dt i
d Sz ω
= [Sz , Sz ] = 0 , (4.96)
dt i
where
|e| B
ω= . (4.97)
me c
At time t = 0 the system is in state
1
|+; x̂ = √ (|+ + |− ) , (4.98)
2
thus
yields
dφα (p′ )
i = dp′′ p′ | H |p′′ φα (p′′ ) . (4.117)
dt
The following hold
and
i (p′ − p′′ ) · r′
= (2π )−3 dr′ exp − V (r′ )
= U (p′ − p′′ ) ,
(4.119)
thus the momentum wave functionφα (p′ ) satisfies the following equation
dφα p′2
i = φ + dp′′ U (p′ − p′′ ) φα . (4.120)
dt 2m α
6. The Hamiltonian is given by
p2
H= + V (r) . (4.121)
2m
Using Eq. (4.38) one has
d x 1 1 ), -* px
= [x, H] = x, p2x = , (4.122)
dt i i 2m m
and
d px 1
= [px , V (r)] , (4.123)
dt i
or with the help of Eq. (3.76)
5 6
d px ∂V
=− . (4.124)
dt ∂x
This together with Eq. (4.122) yield
5 6
d2 x ∂V
m =− . (4.125)
dt2 ∂x
Similar equations are obtained for y and z , which together yield Eq.
(4.57).
7. Substituting a solution having the form
ψ (r) = ψ1 (x1 ) ψ2 (x2 ) ψ3 (x3 ) (4.126)
into the time-independent Schrödinger equation, which is given by
2m
∇2 ψ (r) + 2
[E − V (r)] ψ (r) = 0 , (4.127)
In the sum, the ith term (i ∈ {1, 2, 3}) depends only on xi , thus each
term must be a constant
1 d2 ψi (xi ) 2m 2m
2 − 2 Vi (xi ) = − 2 Ei , (4.129)
ψ i (xi ) dxi
where E1 + E2 + E3 = E.
ψ1 = eα ψ 2 , (4.138)
d2 ψ 2m
+ 2 (E − V (x)) ψ = 0 , (4.139)
dx2
which the eigen-wave-functions satisfy, is real. Therefore given that ψ (x)
is a solution with a given eigenenergy E, then also ψ ∗ (x) is a solution
d2 ψn d2 ψn+1 2m
ψn+1 − ψ n + 2 (En − En+1 ) ψn ψ n+1 = 0 , (4.142)
dx2 dx2
or
d dψn dψ 2m
ψn+1 − ψn n+1 + 2
[En − En+1 ] ψn ψn+1 = 0 . (4.143)
dx dx dx
Without lost of generality, assume that ψn (x) > 0 in the range (x1 , x2 ).
Since ψn (x) is expected to be continuous, the following must hold
dψn
>0, (4.145)
dx x=x1
dψn
<0. (4.146)
dx x=x2
As can be clearly seen from Eq. (4.144), the assumption that ψn+1 (x) > 0
in the entire range (x1 , x2 ) leads to contradiction. Similarly, the possibil-
ity that ψn+1 (x) < 0 in the entire range (x1 , x2 ) is excluded. Therefore,
ψn+1 must have at least one zero in this range.
10. Clearly if ψ (x) is an eigen function with energy E, also ψ (−x) is an
eigen function with the same energy. Consider two cases: (i) The level E
is non-degenerate. In this case ψ (x) = cψ (−x), where c is a constant.
Normalization requires that |c|2 = 1. Moreover, since the wavefunctions
can be chosen to be real, the following holds: ψ (x) = ±ψ (−x). (ii) The
level E is degenerate. In this case every superposition of ψ (x) and ψ (−x)
can be written as a superposition of an odd eigen function ψodd (x) and
an even one ψ even (x), which are defined by
d2 ψ (x) 2m
+ 2 (E − V (x)) ψ (x) = 0 . (4.149)
dx2
Assume V (x) has a finite discontinuity at x = x0 . Integrating the
Schrödinger equation in the interval (x0 − ε, x0 + ε) yields
x0 +ε
x0 +ε
dψ (x) 2m
= 2
(V (x) − E) ψ (x) = 0 . (4.150)
dx x0 −ε
x0 −ε
where
√
−2mE
γ= , (4.152)
and
2m (W + E)
k= . (4.153)
Requiring that both ψ (x) and dψ (x) /dx are continuous at x = a yields
and
or in a matrix form
A 0
C = , (4.156)
B 0
where
Note, however, that according to Eq. (4.158) tan K > 0. Thus, Eq. (4.158)
is equivalent to the set of equations
K
|cos K| = , (4.162)
K0
tan K > 0 . (4.163)
This set has at least one solution (this can be seen by plotting the func-
tions |cos K| and K/K0 ).
13. Final answers: (a) |a1 |2 + |a2 |2 . (b)
7 ! 3
8 "2
2 2 8 3
π 9 2 2
∆E = |an | n4 − |an | n2 . (4.164)
2ma2 n=1 n=1
d2 2m
+ 2 E ψ (x) = 0 . (4.165)
dx2
The boundary conditions at x = 0 are
ψ 0+ = ψ 0− , (4.166)
+ −
dψ (0 ) dψ (0 ) 2
− = − ψ (0) , (4.167)
dx dx a0
where
2
a0 = . (4.168)
mα
The parameter κ is real for E < 0. This even wavefunction satisfies the
Schrödinger equation for x = 0 and the boundary condition (4.166). The
condition (4.167) leads to a single solution for the energy of the ground
state
mα2
E=− . (4.171)
2 2
Thus the normalized wavefunction of the ground state ψ0 (x) is given by
0
mα mα
ψ0,α (x) = 2
exp − 2 |x| . (4.172)
∞ 2
mα ip0 x 2mα
= 2
exp exp − 2
|x| dx
−∞
1
= 2 .
2 p2
1+ 0
4m2 α2
(4.175)
16. The Schrödinger equation for the wavefunction ψ (x) is given by
d2 2m
2
+ 2 (E − V ) ψ (x) = 0 . (4.176)
dx
The boundary conditions imposed upon ψ (x) by the potential are [see
Eq. (4.150)]
ψ (±a) = 0 , (4.177)
ψ 0+ = ψ 0− , (4.178)
dψ (0+ ) dψ (0− ) 2
− = − ψ (0) , (4.179)
dx dx a0
where
2
a0 = . (4.180)
mα
Due to symmetry V (x) = V (−x) the solutions are expected to have
definite symmetry (even ψ (x) = ψ (−x) or odd ψ (x) = −ψ (−x)). For
the ground state, which is expected to have even symmetry, we consider
a wavefunction having the form
+
A sinh (κ (x − a)) x > 0
ψ (x) = , (4.181)
−A sinh (κ (x + a)) x < 0
where A is a normalization constants and where
√
−2mE0
κ= . (4.182)
The parameter κ is real for E0 < 0. This even wavefunction satisfies Eq.
(4.176) for x = 0 and the boundary conditions (4.177) and (4.178). The
condition (4.179) reads
κa0 = tanh (κa) . (4.183)
Nontrivial (κ = 0) real solution exists only when a > a0 , thus E0 < 0 iff
2
a > a0 = . (4.184)
mα
17. For the present case the boundary conditions imposed upon ψ (x) by the
potential are [see Eq. (4.150)]
ψ (0) = 0 , (4.185)
ψ x+
0 = ψ x0
−
, (4.186)
dψ x+ 0 dψ x−0 2
− = − ψ (x0 ) , (4.187)
dx dx a0
where
2
a0 = . (4.188)
mα
Consider a solution having the form
+
A sinh (κx) + B cosh (κx) 0 ≤ x < x0
ψ (x) = , (4.189)
e−κ(x−x0 ) x > x0
where
√
−2mE0
κ= . (4.190)
dx(H) 1 3 (H) 4
= x ,H , (4.197)
dt i
therefore
dx(H) 1 i (Ek − El )
k| |l = k| x(H) H − Hx(H) |l = k| x(H) |l .
dt i
(4.198)
Integrating yields
i (Ek − El ) t
k| x(H) (t) |l = k| x(H) (t = 0) |l exp . (4.199)
1 3 (H) 4 px
(H)
dx(H)
= x ,H = , (4.201)
dt i m
therefore
3 4
(Ek − El ) | k| x |l |2 = l| x(H) , p(H)
x |l
2im
k
= i
2im
2
= .
2m
(4.202)
19. The wavefunctions of the normalized eigenstates are given by
0
2 nπx
ψn (x) = sin , (4.203)
a a
and the corresponding eigenenergies are
π2 2 n2
En = . (4.204)
2ma2
a) The wavefunction after the measurement is a normalized wavepacket
centered at x = a/2 and having a width ∆x
+ 1
√ x − a2 ≤ ∆x
2
ψ (x) = ∆x . (4.205)
0 else
Thus in the limit ∆x ≪ a
a 2
∆x 2 nπ
pn = dxψ∗n (x) ψ (x) ≃2 sin . (4.206)
0 a 2
Namely, pn = 0 for all even n, and the probability of all energies with
odd n is equal.
b) Generally, for every bound state in one dimension p = 0 [see Eq.
(4.52)].
20. For a well of width a the wavefunctions of the normalized eigenstates are
given by
0
(a) 2 nπx
ψn (x) = sin , (4.207)
a a
and the corresponding eigenenergies are
2 2 2
π n
En(a) = . (4.208)
2ma2
(a) The probability is given by
a 2
(a) (2a) 32
p= dxψ1 (x) ψ1 (x) = . (4.209)
0 9π 2
(a)
(b) For times t < 0 it is given that H = E1 . Immediately after the
change (t = 0+ ) the wavefunction remains unchanged. A direct evaluation
(a)
of H using the new Hamiltonian yields the same result H = E1 as for
t < 0. At later times t > 0 the expectation value H remains unchanged
due to energy conservation.
21. The wavefunctions of the normalized eigenstates are given by [see Eq.
(4.204)]
0
2 nπx
ψn (x) = sin , (4.210)
a a
and the corresponding eigenenergies are [see Eq. (4.203)]
π2 2 n2
En = , (4.211)
2ma2
ψ = ψ (x′ ) = x′ |α . (4.218)
dρ 1 q 2 q 2
i = ψ∗ −i ∇− A ψ − ψ i ∇− A ψ∗ , (4.219)
dt 2m c c
where
ρ = ψψ∗ (4.220)
dρ
+ ∇J = 0 , (4.222)
dt
where
qρ
J= Im (ψ∗ ∇ψ) − A. (4.223)
m mc
23. The current density J [see Eq. (4.75)] that is associated with the wave-
′ ′
function ψ (x′ ) = Aeikx + Be−ikx is given by
∂
J= Im ψ∗ ψ
m ∂x′
′ ′ ′ ′
= Im ik A∗ e−ikx + B ∗ eikx Aeikx − Be−ikx
m
′ ′
= Im ik |A|2 − |B|2 + AB ∗ e2ikx − A∗ Be−2ikx
m
k
= |A|2 − |B|2 .
m
(4.224)
Thus for a solution to the time independent Schrödinger equation, for
which the current density ρ = ψψ∗ is time independent, the continuity
equation (4.73) yields the relation
24. In this problem the potential is piecewise constant. At the points where
the piecewise constant potential abruptly changes the solution has to
satisfy the requirements that both ψ (x) and dψ/dx′ [see Eq. (4.150)] are
Al Ar
=M , (4.230)
Bl Br
where
−1
eikl x0 e−ikl x0 eikr x0 e−ikr x0
M= ikl x0
kl e −kl e−ikl x0 kr eikr x0
−kr e−ikr x0
! "
1 1 + kkrl 1 − kr
kl
= Φ (kl x0 ) Φ (−kr x0 ) ,
2 1 − kkrl 1 + kr
kl
(4.231)
and where the matrix Φ (θ) is defined by
e−iθ 0
Φ (θ) = . (4.232)
0 eiθ
The above general result (4.230) is employed below for the given piecewise
constant potential (4.77). Assume first that the wave function on the right
′
side of the barrier in the region x′ > a/2 is given by ψ (x′ ) = eikx , and
′
on the left side of the barrier in the region x < −a/2 it is given by
′ ′
ψ (x′ ) = Aeikx + Be−ikx . With the help of Eq. (4.230) one finds that
1 ka κ κ
A 1+ k 1− k
= Φ − κ κ Φ (κa)
B 4 2 1− k 1+ k
k k ka
1+ κ 1− κ 1
× k k Φ −
1− κ 1+ κ 2 0
eika cos κa − 2i κk + κk sin κa
= 1 κ k ,
2 i k − κ sin κa
(4.233)
where
2 2
k
=E, (4.234)
2m
2 2
κ
= E − Ub , (4.235)
2m
and thus
1 e−ika
t= = , (4.236)
A cos κa − 2i κk + κ
k sin κa
1
B i κ−k e −ika
sin κa
r= = 2 k iκ k κ
. (4.237)
A cos κa − 2 κ + k sin κa
Note that, as is expected from current conservation [see Eq. (4.225)], the
following holds |t|2 + |r|2 = 1.
25. Using Eq. (4.45) one has
1
p(c) x(c) e [ξ(x −x)+η(p −p)] dξdηdx(c) dp(c) .
i (c) (c)
A (x, p) = 2
(2π )
(4.238)
one has
i
ξx − i ηp i
(ξx+ηp) − 2 12 ξη[x,p]
e− e = e− e , (4.240)
thus
1 (c) (c) i ξη
p(c) x(c) e (ξx +ηp ) e 2 e− ξx e− ηp dξdηdx(c) dp(c)
i i i
A (x, p) =
(2π )2
1 (c) η (c)
p(c) x(c) e [(ξ(x + 2 )+ηp )] e− ξx e− ηp dξdηdx(c) dp(c) .
i i i
=
(2π )2
(4.241)
Changing the integration variable
(4.247)
Integration by parts yields
∂p(c)
A (x, p) = px − δ p(c) − p dp(c)
2i ∂p(c)
= px −
2i
[x, p]
= px +
2
xp + px
= .
2
(4.248)
(c) (c)
26. Below we derive an expression for the variable A x , p in terms
of the matrix elements of the operator A (x, p) in the basis of position
eigenvectors |x′ . To that end
' ( we begin by evaluating the matrix element
′ x′′ ′ x′′
x − 2 A (x, p) x + 2 using Eqs. (4.241), (3.19) and (4.245)
5 6
′ x′′ ′ x′′
x − A (x, p) x +
2 2
1 η
A x(c) , p(c) e [(ξ(x + 2 )+ηp )]
i (c) (c)
= 2
(2π )
5 6
x′′ − i ξx − i ηp ′ x′′
× x′ − e e x + dξdηdx(c) dp(c)
2 2
1 (c) (c) i
′′
[(ξ(x(c) + η2 )+ηp(c) )] e− i ξ x′ − x2
= A x , p e
(2π )2
5 6
x′′ ′ x′′
× x′ − x + + η dξdηdx(c) dp(c)
2 2
1 ′
A x(c) , p(c) e− x p dx(c) dp(c) e [ξ(x −x )] dξ
i ′′ (c) i (c)
= 2
(2π )
1 i ′′ (c)
= A x(c) , p(c) e− x p dx(c) dp(c) δ x(c) − x′
2π
1 i ′′ (c)
= A x′ , p(c) e− x p dp(c) .
2π
i
x′′ p′
Applying the inverse Fourier transform, i.e. multiplying by e and
integrating over x′′ yields
5 6
′ x′′ ′ x′′ i ′′ ′
x − A (x, p) x + e x p dx′′
2 2
1 i ′′ ′
A x′ , p(c) dp(c) e x (p −p ) dx′′ ,
(c)
=
2π
(4.249)
thus with the help of Eq. (4.246) one finds the desired inversion of Eq.
(4.45) is given by
i
x′′ (p′ −p′′ )
= p′′ | A (x, p) |p′′ e dx′′ dp′′
= 2π p′ | A (x, p) |p′ .
(4.252)
27. By expressing |ψ as
|ψ = |ψ1 + |ψ2 ,
In the limit N → ∞ the effective Hamiltonian Heff becomes [in this limit
u → 1, see Eq. (4.89)]
N
1 k−1
Heff = U† HU k−1 . (4.259)
N
k=1
U= Pn eiθn , (4.260)
n
where Pn are projection operators, eiθn are eigenvalues and θ n are distinct
real numbers. Using this notation Eq. (4.259) becomes
N
1
Heff = Pn′′ HPn′ ei(θn′ −θn′′ )(k−1) . (4.261)
n′ ,n′′
N
k=1
one finds that in the limit N → ∞ the effective Hamiltonian Heff (4.261)
becomes
Heff = Pn′ HPn′ . (4.263)
n′
α = α0 e−iωt , (5.4)
p2 mω 2 x2
H= + . (5.7)
2m 2
In quantum mechanics the variables x and p are regarded as operators satis-
fying the following commutation relations [see Eq. (3.9)]
[x, p] = xp − px = i . (5.8)
Chapter 5. The Harmonic Oscillator
5.1 Eigenstates
N = a† a, (5.14)
1
H= ω N+ . (5.16)
2
N |n = n |n . (5.17)
H |n = En |n , (5.18)
1
En = ω n + . (5.19)
2
|n + 1 = (n + 1)−1/2 a† |n (5.20)
|n − 1 = n−1/2 a |n (5.21)
and
N a |n = ([N, a] + aN ) |n = (n − 1) a |n . (5.25)
and
n − 1 |n − 1 = n−1 n| a† a |n = 1 . (5.27)
Claim. The spectrum (i.e. the set of eigenvalues) of N are the nonnegative
integers {0, 1, 2, · · · }.
Proof. First, note that since the operator N is positive-definite the eigenval-
ues are necessarily non negative
n = n| a† a |n ≥ 0 . (5.30)
Furthermore, using Eq. (5.29) one can express the state |n in terms of the
ground state |0 as
n
a†
|n = √ |0 . (5.32)
n!
As can be easily seen from Eqs. (5.11), (5.12), (5.28) and (5.29), all energy
eigenstates |n have vanishing position and momentum expectation values
n| x |n = 0 , (5.33)
n| p |n = 0 . (5.34)
Clearly these states don’t oscillate in phase space as classical harmonic os-
cillators do. Can one find quantum states having dynamics that resembles
classical harmonic oscillators?
|α = D (α) |0 , (5.35)
where
In the set of problems at the end of this chapter the following results are
obtained:
• The displacement operator is unitary D† (α) D (α) = D (α) D† (α) = 1.
• The coherent state |α is an eigenvector of the operator a with an eigenvalue
α, namely
a |α = α |α . (5.37)
2
H α = α| H |α = ω |α| + 1/2 , (5.43)
In view of Eqs. (5.43), (5.45) (5.48) and (5.49), we see from this results that
H α , ∆Hα , ∆xα and ∆pα are all time independent. On the other hand, as
can be seen from Eqs. (5.46) and (5.47) the following holds
0
2
x α = α| x |α = Re α0 e−iωt , (5.54)
mω
√
p α = α| p |α = 2 mω Im α0 e−iωt . (5.55)
These results show that indeed, x α and p α have oscillatory time depen-
dence identical to the dynamics of the position and momentum of a classical
harmonic oscillator [compare with Eqs. (5.5) and (5.6)].
5.3 Problems
3. Show that
α(2XY −αX 2 −αY 2 )
∞ α n exp 1−α2
2 Hn (X) Hn (Y )
= √ , (5.58)
n=0
n! 1 − α2
x0 = ψ0 | x |ψ0 , (5.60)
p0 = ψ0 | p |ψ0 , (5.61)
(xp)0 = ψ0 | xp |ψ0 , (5.62)
(∆x)20 = ψ0 | (x − x0 )2 |ψ0 , (5.63)
(∆p)20 2
= ψ0 | (p − p0 ) |ψ0 . (5.64)
6. Consider a harmonic oscillator of angular frequency ω and mass m.
a) Express the Heisenberg picture x(H) (t) and p(H) (t) in terms x(H) (0)
and p(H) (0). , - , -
b) Calculate the following commutators p(H) (t1 ) , x(H) (t2 ) , p(H) (t1 ) , p(H) (t2 )
, -
and x(H) (t1 ) , x(H) (t2 ) .
7. Consider a particle having mass m confined by a one dimensional poten-
tial V (x), which is given by
+ mω2
2
V (x) = 2 x x>0 , (5.65)
∞ x≤0
where ω is a constant.
a) Calculate the eigenenergies of the) system.
*
b) Calculate the expectation values x2 of all energy eigenstates of the
particle.
8. Calculate the possible energy values of a particle in the potential given
by
mω2 2
V (x) = x + αx . (5.66)
2
9. A particle is in the ground state of harmonic oscillator with potential
energy
mω2 2
V (x) = x . (5.67)
2
Find the probability p to find the particle in the classically forbidden
region.
10. Consider a particle having mass m in a potential V given by
+ mω2 z2 a
− 2 ≤ x ≤ a2 and − a2 ≤ y ≤ a2
V (x, y, z) = 2 , (5.68)
∞ else
where ω and a are positive real constants. Find the eigenenergies of the
system.
11. Consider a harmonic oscillator having angular resonance frequency ω 0 .
At time t = 0 the system’s state is given by
1
|α (t = 0) = √ (|0 + |1 ) , (5.69)
2
where the states |0 and |1 are the ground and first excited states, re-
spectively, of the oscillator. Calculate as a function of time t the following
quantities:
a) x
b) )p *
c) x2
d) ∆x∆p
12. Harmonic oscillator having angular resonance frequency ω is in state
1
|ψ (t = 0) = √ (|0 + |n ) (5.70)
2
at time t = 0, where |0 is the ground state and |n is the eigenstate
with eigenenergy ω (n + 1/2) (n is a non zero integer). Calculate the
expectation value x for time t ≥ 0.
13. Consider a harmonic oscillator having mass m and angular resonance
frequency ω . At time t = 0 the system’s state is given by |ψ(0) = c0 |0 +
c1 |1 , where |n are the eigenstates with energies En = ω (n +.1/2).
1
Given that H = ω, |ψ(0) is normalized, and x (t = 0) = 2 mω ,
calculate x (t) at times t > 0.
14. Show that
|α|2 † ∗ |α|2 ∗ †
D (α) = e− 2 eαa e−α a
=e 2 e−α a eαa . (5.71)
17. Show that the coherent state |α is an eigenvector of the operator a with
an eigenvalue α, namely
a |α = α |α . (5.73)
p2 1
H= + mω 2 x2 . (5.76)
2m 2
The system at time t is in a normalized state |α , which is an eigenvector
of the annihilation operator a, thus
a |α = α |α , (5.77)
ψ| a† a† aa |ψ
g(2) = 2 . (5.86)
ψ| a† a |ψ
|ψ = S † (r) |0 , (5.87)
where the operator S (r) is given by Eq. (5.83) and where r is a real
number.
31. The state |r from the previous exercise, which is called a squeezed state,
can be alternatively defined as a normalized state that satisfies the rela-
tion
Q (r) |r = 0 , (5.88)
r is a real number, and a and a† are the annihilation and creation opera-
tors respectively. Based on the above definition calculate the expectation
values r| x |r of the position operator x, the expectation value r| p |r
of the momentum operator p, the variance (∆x)2 of x and the variance
(∆p)2 of p with respect to the state |r .
32. Consider one dimensional motion of a particle having mass m. The Hamil-
tonian is given by
H = ω 0 a† a + ω1 a† a† aa , (5.90)
where
0
mω 0 ip
a= x+ , (5.91)
2 mω0
H = ǫN , (5.92)
where the real non-negative parameter ǫ has units of energy, and where
the operator N is given by
N = b† b . (5.93)
|α (t = 0) = |+ , (5.98)
x ′2 √ x′ † a†2
exp − 2x2 + 2 x0 a − 2
′ 0
|x = 1/2
|0 , (5.105)
π1/4 x0
a + a† x
X= √ = , (5.110)
2 x0
where a is the annihilation operator, x is the position operator and
0
x0 = . (5.111)
mω
Show that
d dF
: F (X) : =: : . (5.112)
dX dX
41. Calculate the matrix elements n2 | S |n1 , where the operator S is given
by
∞ k
eλ − 1
S= a†k ak , (5.113)
k!
k=0
where a and a† are the annihilation and creation operators, both ω and
ω 1 are positive, and where k is integer. At initial time t = 0 the state
of the system is an eigenstate of the operator a with eigenvalue α, i.e.
|ψ (t = 0) = |α c , where a |α c = α |α c .
a) Find a general expression for the state of the system |ψ (t) at time
t > 0.
b) Evaluate |ψ (t) at time t = 2π/ω 1 .
c) Evaluate |ψ (t) at time t = π/ω1 .
d) Evaluate |ψ (t) at time t = π/2ω 1 for the case where k is even.
43. Consider two normalized coherent states |α and |β , where α, β ∈ C.
The operator A is defined as
A = |α α| − |β β| . (5.115)
5.4 Solutions
1. The Hamiltonian is given by
p2 mω2 x2
H= + .
2m 2
thus
! "
∞ 2
2 x
|A0 | exp − dx = 1 . (5.122)
−∞ x0
√
πx0
All other wavefunctions are found using Eqs. (5.32) and (5.117)
n
1 d
ψn (x′ ) = √ x′ − x20 ψ0 (x′ )
(2x0 )n/2 n! dx′
! "
n 2
1 1 ′ d 1 x′
= √ x − x20 exp − .
π 1/4 n n+1/2
2 n! x0 dx′ 2 x0
(5.124)
Using the notation
dn
Hn (X) = exp 2Xt − t2 . (5.132)
dtn t=0
2
The identity 2Xt − t2 = X 2 − (X − t) yields
dn
Hn (X) = exp X 2 exp − (X − t)2 . (5.133)
dtn t=0
d d
− exp X 2 exp −X 2 g = 2X − g, (5.136)
dX dX
and
X2 d X2 d
exp X− exp − g = 2X − g, (5.137)
2 dX 2 dX
thus
n
X2 d X2
Hn (X) = exp X− exp − . (5.138)
2 dX 2
according to which the following holds (for the case a = 1, b = 2iX and
c = 0)
∞
2 1
exp −X =√ exp −x2 + 2iXx dx , (5.140)
π
−∞
b) Using the expressions for x(H) (t) and p(H) (t) and Eq. (5.8) one finds
that 3 4
p(H) (t1 ) , x(H) (t2 )
3 4
= − (cos (ωt1 ) cos (ωt2 ) + sin (ωt1 ) sin (ωt2 )) x(H) (0) , p(H) (0)
= −i cos (ω (t1 − t2 )) ,
3 4 (5.157)
p(H) (t1 ) , p(H) (t2 )
3 4
= mω (cos (ωt1 ) sin (ωt2 ) − sin (ωt1 ) cos (ωt2 )) x(H) (0) , p(H) (0)
= −i mω sin (ω (t1 − t2 )) ,
(5.158)
and 3 4
x(H) (t1 ) , x(H) (t2 )
1 3 4
= (cos (ωt1 ) sin (ωt2 ) − sin (ωt1 ) cos (ωt2 )) x(H) (0) , p(H) (0)
mω
i
=− sin (ω (t1 − t2 )) .
mω
(5.159)
7. Due to the infinite barrier for x ≤ 0 the wavefunction must vanish at
x = 0. This condition is satisfied by the wavefunction of all number
states |n with odd value of n (the states |n are eigenstates of the ’regu-
lar’ harmonic oscillator with potential V (x) = mω 2 /2 x2 ). These wave-
functions obviously satisfy the Schrödinger equation for x > 0.
a) Thus the possible energy values are Ek = ω (2k + 3/2) where k =
0, 1, 2, · · · .
b) The corresponding normalized wavefunctions are given by
+√
2ψ2k+1 (x) x > 0
ψ̃k (x) = , (5.160)
0 x≤0
= 2k + 1| x2 |2k + 1 ,
(5.161)
thus with the help of Eq. (5.144) one finds that
) 2* 3
x k= 2k + . (5.162)
mω 2
8. The potential can be written as
mω2 α 2 α2
V (x) = x+ − . (5.163)
2 mω 2 2mω 2
This describes a harmonic oscillator centered at x0 = −α/mω 2 having
angular resonance frequency ω. The last constant term represents energy
shift. Thus, the eigenenergies are given by
where n = 0, 1, 2, · · · .
9. In the classically forbidden region V (x) > E0 = ω/2, namely |x| > x0
where
0
x0 = . (5.165)
mω
Using Eq. (5.123) one finds
∞
2
p=2 |ψ0 (x)| dx
x0
! "
∞ 2
2 x
= exp − dx
π1/2 x0 x0 x0
= 1 − erf (1)
= 0.157 .
(5.166)
10. The answer is [see Eqs. (4.204) and (5.19)]
π2 2
n2x + n2y 1
Enx ,ny ,nz = 2
+ ω nz + , (5.167)
2ma 2
= .
mω 0
(5.176)
d) Similarly
) 2* m ω0 2
p =− α (t)| −a + a† |α (t)
2
m ω0 2 , -
=− α (t)| a2 + a† − a, a† − 2a† a |α (t)
2
= m ω0 ,
(5.177)
thus ;
0
cos2 (ω 0 t) sin2 (ω0 t)
∆x∆p = 1− 1−
2 2
0
1
= 2 + sin2 (2ω 0 t) .
2 4
(5.178)
1 iE0 t iEn t
|ψ (t) = √ exp − |0 + exp − |n , (5.179)
2
where
1
En = ω n + , (5.180)
2
thus, using
0
x= a + a† , (5.181)
2mω
and
√
a |n = n |n − 1 , (5.182)
√
a† |n = n + 1 |n + 1 , (5.183)
one finds that x (t) = 0 if n > 1, and for n = 1
0
x (t) = ψ (t)| a + a† |ψ (t)
2mω
0
= cos (ωt) .
2mω
(5.184)
13. Since H = ω and |ψ(0) is normalized one has
2 2 1
|c0 | = |c1 | = , (5.185)
2
thus |ψ(0) can be written as
0
1
|ψ(0) = |0 + eiθ |1 , (5.186)
2
where θ is real. Given that at time t = 0
0
1
x (t = 0) = , (5.187)
2 mω
one finds using the identities
0
x= a + a† , (5.188)
2mω
√
a |n = n |n − 1 , (5.189)
√
a† |n = n + 1 |n + 1 , (5.190)
that
√
2
cos θ = . (5.191)
2
Using this result one can evaluate p (t = 0), where
0
m ω
p=i −a + a† , (5.192)
2
thus
0 0 √
m ω m ω 2
p (t = 0) = sin θ = ± = ±mω x (t = 0) . (5.193)
2 2 2
Using these results together with Eq. (5.155) yields
0
1
x (t) = (cos (ωt) ± sin (ωt))
2 mω
0
π
= cos ωt ∓ .
2mω 4
(5.194)
14. According to identity (2.180), which states that
1 1
eA+B = eA eB e− 2 [A,B] = eB eA e 2 [A,B] , (5.195)
provided that
thus
16. Using Eqs. (5.35), (5.28) and (5.29) one finds that
|α|2 † ∗ |α|2 †
|α = e− 2 eαa e−α a
|0 = e− 2 eαa |0
∞
|α|2 αn
= e− 2 √ |n .
n=0 n!
(5.200)
17. Using Eqs. (5.42) and (5.28) one has
∞
|α|2 αn
a |α = e− 2 √ a |n
n=0 n!
∞
−
|α|2 αn−1
= αe 2 |n − 1
n=1 (n − 1)!
= α |α .
(5.201)
18. Using Eqs. (5.36), (5.9) and (5.10) one has
0 0
mω 1
D (α) = exp (α − α∗ ) x − i (α + α∗ ) p , (5.202)
2 2 mω
thus with the help of Eqs. (2.180) and (5.8) the desired result is obtained
0
mω α − α∗
D (α) = exp √ x
2
i α + α∗ α∗2 − α2
× exp − √ √ p exp .
m ω 2 4
(5.203)
D (α) D† (α) = 1
using
a |α = α |α , (5.213)
and
√
a |n = n |n − 1 , (5.214)
one finds
∞ ∞
√
cn n |n − 1 = α cn |n , (5.215)
n=0 n=0
thus
α
cn+1 = √ cn , (5.216)
n+1
therefore
∞
αn
|α = A √ |n . (5.217)
n=0 n!
thus
|α|2 αn
cn = e− 2 √ . (5.220)
n!
Note that this result is identical to Eq. (5.42), thus |α is a coherent
state. The possible results of the measurement are
1
En = ω n + , (5.221)
2
i p αx i x αp α∗2 − α2
D (α) = exp exp − exp . (5.225)
4
i x αp
exp − |x′ = |x′ + x α ,
thus
i p αx i x αp α∗2 − α2
x′ |α = x′ | exp exp − exp |0
4
′
α∗2 − α2 i p αx
= exp exp x′ − x α |0 .
4
(5.226)
where
0
∆xα = , (5.228)
2mω
thus
2
x′ − x α
∗2 2 ′ exp − 2∆xα
α −α i p αx
x′ |α = exp exp √
4 (2π)1/4 ∆xα
2
α∗2 − α2 mω 1/4 x− x α x
= exp exp − +i p α .
4 π 2∆xα
(5.229)
23. Using Eqs. (5.36) and (2.180) this relation is easily obtained.
24. With the help of Eq. (5.42) one has
1 1 1 2
|α α| d2 α = |n m| √ e−|α| αn α∗m d2 α .
π π n,m n!m!
(5.230)
2πδnm
∞
2 2
= |n n| dρρ2n+1 e−ρ
n
n!
0
1
= |n n| Γ (n + 1)
n
n!
=n!
= |n n|
n
=1.
(5.231)
x′ = x + β . (5.235)
a) Thus, using Eqs. (5.155) and (5.156) together with the relations
x′(H) (t) = x(H) (t) + β , (5.236)
p(H) (t) = p′(H) (t) , (5.237)
one finds
sin (ωt) (H)
x(H) (t) = x(H) (0) + β cos (ωt) + p (0) − β , (5.238)
mω
p(H) (t) = p(H) (0) cos (ωt) − mω sin (ωt) x(H) (0) + β . (5.239)
b) For '
this case (at time t = 0
x(H) (0) = 0 , (5.240)
' (
p(H) (0) = 0 , (5.241)
thus
' (
x(H) (t) = β (cos (ωt) − 1) . (5.242)
e−|α0 | α2n
2 n
2 0 1 mω∆2x mω∆2x
Pn = | n|ψ (t) | = = exp − .
n! n! 2 2
(5.250)
xp = i 0| aa† − a† a |0 = i . (5.258)
2 2
The Hamiltonian for times t > 0 is given by
p2
H= + gx . (5.259)
2m
Using the Heisenberg equation of motion for the operators x and x2 one
finds
dx(H) 1 , -
= x(H) , H , (5.260)
dt i
dp(H) 1 , -
= p(H) , H , (5.261)
dt i
dx2(H) 1 3 2 4
= x(H) , H , (5.262)
dt i
or using [x, p] = i
dx(H) p(H)
= , (5.263)
dt m
dp(H)
= −g , (5.264)
dt
dx2(H) 1 1
= x p + p(H) x(H) = 2x(H) p(H) − i , (5.265)
dt m (H) (H) m
thus
i t 2 t
x2(H) (t) = x2(H) (0) − + x (t′ ) p(H) (t′ ) dt′
m m 0 (H)
i t 2 t p(H) (0) t′ gt′2 , -
= x2(H) (0) − + x(H) (0) + − p(H) (0) − gt′ dt′
m m 0 m 2m
i t
= x2(H) (0) −
! m "
2 t p2(H) (0) t′ gt′2 ′ p(H) (0) gt′2 g 2 t′3
+ x(H) (0) p(H) (0) + − p (0) − x(H) (0) gt − + dt′
m 0 m 2m (H) m 2m
i t
= x2(H) (0) −
! m "
2 p2(H) (0) t2 p(H) (0) gt3 x(H) (0) gt2 p(H) (0) gt3 g2 t4
+ x(H) (0) p(H) (0) t + − − − + .
m 2m 6m 2 3m 8m
(5.268)
Using the initial conditions Eqs. (5.251), (5.252), (5.253), (5.254) and
(5.258) one finds
gt2
x (t) = − , (5.269)
2m
2g 2 t4
x (t) = , (5.270)
4m2
p (t) = −gt , (5.271)
) 2 * 2 2 4
i t 2 i t ωt g t
x (t) = − + + + , (5.272)
2mω m m 2 4 8m
and
' ( ) * ωt2
(∆x)2 (t) = x2 (t) − x (t) 2
= + = 1 + ω 2 t2 .
2mω 2m 2mω
(5.273)
r 2 r4 r3
T = 1+ + +··· a− r+ +··· a† + · · · , (5.282)
2! 4! 3!
a) Thus
T = Aa + Ba† , (5.283)
where
A = cosh r , (5.284)
B = − sinh r . (5.285)
b) Using the0relations
x= a + a† , (5.286)
2mω
0
m ω
p=i −a + a† . (5.287)
2
one finds 0
r| x |r = 0| S (r) a + a† S † (r) |0
2mω
0
= 0| T |0 + 0| T † |0
2mω
=0,
0 (5.288)
m ω
r| p |r = i 0| S (r) −a + a† S † (r) |0
2
0
= − 0| T |0 + 0| T † |0
2mω
=0.
(5.289)
c) Note that S (r) is unitary, namely S † (r) S (r) = 1, since the operator
2
a2 − a† is anti Hermitian. Thus
r| x2 |r = 0| S (r) a + a† a + a† S † (r) |0
2mω
= 0| S (r) a + a† S † (r) S (r) a + a† S † (r) |0
2mω
2
= 0| T + T † |0
2mω
(A + B)2 2
= 0| a + a† |0
2mω
(cosh r − sinh r)2
=
2mω
e−2r
= ,
2mω
(5.290)
and
m ω 2
r| p2 |r = 0| S (r) a − a† S † (r) |0
2
m ω 2
= 0| T − T † |0
2
m ω (A − B)2 2
= 0| a − a† |0
2
m ω (cosh r + sinh r)2
=
2
m ωe2r
= .
2
(5.291)
Thus
e−2r
(∆x)2 = , (5.292)
2mω
m ωe2r
(∆p)2 = , (5.293)
2
(∆x) (∆p) = . (5.294)
2
30. Using the relation [see Eq. (5.283)]
one obtains
one obtains
e−2r 2
r| x2 |r = r| Q (r) + Q† (r) |r
2mω
e−2r
= r| Q (r) Q† (r) |r
2mω
e−2r
= ,
2mω
(5.302)
and similarly
mωe2r 2
r| p2 |r = − r| Q (r) − Q† (r) |r
2
mωe2r
= r| Q (r) Q† (r) |r
2
mωe2r
= ,
2
(5.303)
thus
2 e−2r
(∆x) = , (5.304)
2mω
2 mωe2r
(∆p) = . (5.305)
2
32. Using the commutation relation
, †-
a, a = 1 , (5.306)
one finds
H = ω0N + ω1 N 2 − N , (5.307)
where
N = a† a (5.308)
is the number operator.
a) The eigenvectors of N
N |n = n |n , (5.309)
(where n = 0, 1, · · · ) are also eigenvectors of H and the following
holds
H |n = En |n , (5.310)
where
, -
En = ω0 n + ω1 n2 − n . (5.311)
Note that
En+1 − En
= ω 0 + 2ω 1 n , (5.312)
i. 0|x|0 = 0
ii. '0|p|0 = 0 (
iii. 0| (∆x)2 |0 = 2mω0
' (
iv. 0| (∆p)2 |0 = m ω0
2
33. The proof of the clue is:
2
N = b† bb† b = b† 1 − b† b b = N . (5.319)
N |n = n |n . (5.320)
Using the clue one finds that n2 = n, thus the possible values of n
are 0 (ground state) and 1 (excited state). Thus, the eigenvalues of
H are 0 and ǫ.
b) To verify the statement in the clue we calculate
N b† |0 = b† bb† |0 = b† (1 − N ) |0 = b† |0 , (5.321)
|1 = b† |0 . (5.322)
0|1 = 0 . (5.324)
(5.328)
Thus
2
iHt
p0 (t) = −| exp − |+
2
1 iǫt
= ( 0| − 1|) 1 + N −1 + exp − (|0 + |1 )
4
2
1 iǫt
= 1 − exp −
4
ǫt
= sin2 .
2
(5.329)
34. The closure relation (5.31) can be written as
∞
1= |n m| δ n,m . (5.330)
n,m=0
which is obtained using the Taylor power expansion series of the function
ς m , one finds that
|0 0| = : Z : . (5.333)
: f g : =: gf : , (5.334)
: f gh : =: f hg : , (5.335)
and
: f (: g : ) : = : f g : . (5.336)
Thus
d
1 = exp a† : Z : exp (aς)
dς ς=0
d †
= : exp a Z exp (aς) :
dς ς=0
d
= : exp a† exp (aς) Z :
dς ς=0
d n m
a† dς (aς)
= : Z :
n,m
n! m!
ς=0
d n
† n dς m am
= : a ς Z:
n,m
n! m!
ς=0
δn,m
†
= : exp a a Z :
= : exp a† a ( : Z : ) : ,
(5.337)
and therefore
where the initial time t0 will be taken below to be −∞. The Heisenberg
operator a† (t) is found from the Hermitian conjugate of the last result.
Let Pn (t) be the Heisenberg representation of the projector |n n|. The
probability pn (t) to find the oscillator in the number state |n at time t
is given by
thus
+
− |n n odd
P |n = , (5.349)
|n n even
P = eiπN . (5.350)
37. Using Eqs. (5.31), (5.32) and (5.338) together with the relation
a† a |n = n |n , (5.351)
yields
∞
†
eλa a
= eλn |n n|
n=0
∞ n
a† an
= eλn √ |0 0| √
n=0 n! n!
∞
eλn † n
= a : exp −a† a : an
n=0
n!
∞
eλn n
= : a† exp −a† a an :
n=0
n!
∞
eλn † n
=: a a exp −a† a :
n=0
n!
=: exp eλ a† a exp −a† a : ,
(5.352)
thus
† , -
eλa a
=: exp eλ − 1 a† a : . (5.353)
where
n
a†
|n = √ |0 , (5.355)
n!
thus with the help of Eq. (5.126) and the generating function of Hermite
polynomials (5.56) one finds that (note that x′ |n is real)
′2
x x′ n
exp − 2x2
∞ Hn x0 a†
′ 0
|x = 1/2
√ √ |0
π 1/4 x0 n=0 2n n! n!
x ′2 √ ′
a†2
exp − 2x2 + 2 xx0 a† − 2
0
= 1/2
|0 .
π1/4 x0
(5.356)
39. Using the relation x |x′ = x′ |x′ and Eq. (3.32) one finds that
∞
′2
2
exp kx = dx′ ekx |x′ x′ | . (5.357)
−∞
where
a + a† x
X= √ = , (5.359)
2 x0
and where
x′
X′ = . (5.360)
x0
Thus
∞
1 ′ 2
−X ) +KX ′2
exp kx 2
=√ dX ′ : e−(X :
π
−∞
∞
1 ′2
+2X ′ X−X 2
=√ dX ′ : e−(1−K)X : ,
π
−∞
(5.361)
where K = kx20 . With the help of the identity (5.139) this becomes
Kx2 1 K x2
exp =√ : exp : . (5.362)
x20 1−K 1 − K x20
1 κ x2 x2
√ exp =: exp κ : . (5.364)
1+κ 1 + κ x20 x20
40. The orthogonality between number states yields according to Eq. (5.126)
(5.367)
(5.369)
d 1 dHn (X)
: Xn : = n
dX 2 dX
Hn−1 (X)
=n
2n−1
= n : X n−1 : ,
(5.372)
thus
d d n
: X n : =: X : . (5.373)
dX dX
Thus, for a general smooth function F (X) of the operator X the following
holds
d dF
: F (X) : =: : . (5.374)
dX dX
41. The following holds [see Eqs. (5.28) and (5.29)]
∞ k
eλ − 1
S |n1 = a†k ak |n1
k!
k=0
n1 k
eλ − 1 n1 !
= |n1 ,
k! (n1 − k)!
k=0
(5.375)
thus, with the help of the binomial theorem one finds that
hence
Alternatively, the same result can be easily obtained with the help of
Eq.(5.104), according to which
†
S = eλa a
. (5.378)
iHt † k
= e−iω1 (a a) t e−iωa at ,
†
u (t) = exp − (5.380)
thus
|ψ (t) = u (t) |ψ (t = 0)
∞
† k |α| 2
αn
= e−iω1 (a a) t e−iωa at e− 2
†
√ |n
n=0 n!
∞ n
† k |α| 2 αe−iωt
= e−iω1 (a a) t e− 2 √ |n
n=0 n!
∞ −iωt n
|α|2 αe
= e− 2 √ e−iφn |n ,
n=0 n!
(5.381)
where
φn = ω 1 tnk . (5.382)
e−iφn = 1 , (5.383)
and therefore
6 (
2π 2πiω
ψ = αe− ω1 . (5.384)
ω1 c
one has
and therefore n
6 − πiω
π |α|2
∞ αe ω1
ψ = e− 2 √ (−1)n |n
ω1 n=0 n!
(
− πiω
= −αe ω 1 .
c
(5.387)
d) At time t = π/2ω 1 the phase factor φn is given by φn = (π/2) nk .
For the case where k is even one has
+
k 0 n is even
mod n , 4 = , (5.388)
1 n is odd
thus
+
1 n is even
e−iφn = , (5.389)
−i n is odd
and therefore
πiω n
π
6 2
∞ αe− 2ω1
− |α|
ψ =e 2 √ e−iφn |n . (5.390)
2ω 1 n=0 n!
This state can be expressed as a superposition of two coherent states
6 ( (
π 1 iπ πiω iπ πiω
ψ = √ e− 4 αe− 2ω1 + e 4 −αe− 2ω1 . (5.391)
2ω 1 2 c c
and
Tr A2 = λ2n . (5.393)
n
On the other hand, with the help of Eq. (2.177) one finds that
Tr (A) = Tr (|α α|) − Tr (|β β|) = 0 , (5.394)
and
Tr A2 = Tr (|α α |α α|) + Tr (|β β |β β|)
− Tr (|α α |β β|) − Tr (|β β |α α|)
= 2 − α |β Tr (|α β|) − β |α Tr (|β α|)
= 2 1 − | α |β |2 .
(5.395)
Clearly, A cannot have more than two nonzero eigenvalues, since the
dimensionality of the subspace spanned by the vectors {|α , |β } is at
most 2, and therefore A has three eigenvalues 0, λ+ and λ− , where [see
Eq. (5.233)]
.
λ± = ± 1 − | α |β |2 = ± 1 − e−|α−β| .
2
(5.396)
Below we will see that one can define a unitary operator that generates ro-
tations.
Exercise 6.1.1. Show that
x x
†
Dẑ (φ) y Dẑ (φ) = Rẑ y ,
(6.8)
z z
where
iφLz
Dẑ (φ) = exp − , (6.9)
and where
cos φ − sin φ 0
Rẑ = sin φ cos φ 0 . (6.10)
0 0 1
The matrix Rẑ [see Eq. (6.10)] represents a geometrical rotation around
the z axis with angle φ. Therefore, in view of the above result, we refer to the
operator Dẑ (φ) as the generator of rotation around the z axis with angle φ.
It is straightforward to generalize the above results and to show that rotation
around an arbitrary unit vector n̂ axis with angle φ is given by
iφL · n̂
Dn̂ (φ) = exp − . (6.20)
On the other hand, as can be seen from Eq. (6.5), different components of L do
not commute and therefore rotation operators Dn̂ (φ) with different rotations
axes n̂ need not commute. Both the above results, which were obtained from
commutation relations between quantum operators, demonstrate two well
known geometrical facts: (i) different linear translations commute, whereas
(ii) generally, different rotations do not commute.
|a, b + 1 ≡ −1
[a − b (b + 1)]−1/2 J+ |a, b (6.31)
where
J+ = Jx + iJy , (6.32)
|a, b − 1 ≡ −1
[a − b (b − 1)]−1/2 J− |a, b (6.35)
where
J− = Jx − iJy , (6.36)
Similarly
where
thus
†
J− J− = J+ J−
= (Jx + iJy ) (Jx − iJy )
= Jx2 + Jy2 + i [Jy , Jx ]
= J2 − Jz2 + Jz ,
(6.44)
one finds that
†
a, b| J+ J+ |a, b = a, b| J2 |a, b − a, b| Jz (Jz + ) |a, b
= 2 [a − b (b + 1)] ,
(6.45)
and
†
a, b| J− J− |a, b = a, b| J2 |a, b − a, b| Jz (Jz − ) |a, b
= 2 [a − b (b − 1)] .
(6.46)
Thus the states |a, b + 1 and |a, b − 1 are both normalized.
What are the possible values of b? Recall that we have shown in chapter 5
for the case of harmonic oscillator that the eigenvalues of the number operator
N must be nonnegative since the operator N is positive-definite. Below we
employ a similar approach to show that:
Claim. b2 ≤ a
Proof. Both Jx2 and Jy2 are positive-definite, therefore
Thus, also
†
J+ J+ |a, bmax = 0 (6.49)
or
In a similar way with the help of Eq. (6.44) one can show that there exists a
minimum value bmin for which
or
The vector space of a spin 1/2 system is the subspace spanned by the ket-
vectors |j = 1/2, m = −1/2 and |j = 1/2, m = 1/2 . In this subspace the
spin angular momentum is labeled using the letter S, as we have discussed
above. The matrix representation of some operators of interest in this basis
can be easily found with the help of Eqs. (6.63), (6.64), (6.65) and (6.66):
3 2 10
S2 =
˙ , (6.69)
4 01
1 0
Sz =
˙ ≡ σz , (6.70)
2 0 −1 2
01
S+ =
˙ , (6.71)
00
00
S− =
˙ . (6.72)
10
The above results for S+ and S− together with the identities
S+ + S−
Sx = , (6.73)
2
S+ − S−
Sy = , (6.74)
2i
can be used to find the matrix representation of Sx and Sy
01
Sx =
˙ ≡ σx , (6.75)
2 10 2
0 −i
Sy =
˙ ≡ σy . (6.76)
2 i 0 2
The matrices σ x , σy and σz are called Pauli’s matrices, and are related to
the corresponding spin angular momentum operators by the relation
Sk =
˙ σk . (6.77)
2
r′ | r |α = r′ r′ |α , (6.88)
r′ | p |α = ∇ r′ |α , (6.89)
i
[see Eqs. (3.21) and (3.29)] one finds that
∂ ∂
r′ | Lx |α = y −z ψα (r′ ) , (6.90)
i ∂z ∂y
∂ ∂
r′ | Ly |α = z −x ψα (r′ ) , (6.91)
i ∂x ∂z
∂ ∂
r′ | Lz |α = x −y ψα (r′ ) , (6.92)
i ∂y ∂x
where
ψα (r′ ) = r′ |α . (6.93)
∂
r′ | Lz |α = −i ψ (r′ ) . (6.98)
∂φ α
2. Using Eqs. (6.90) and (6.91) together with the relation L+ = Lx + iLy
one has
i ′ i ′
r | L+ |α = r | Lx + iLy |α
∂ ∂ ∂ ∂
= y −z + iz − ix ψα (r′ )
∂z ∂y ∂x ∂z
∂ ∂ ∂
= z i − − i (x + iy) ψα (r′ )
∂x ∂y ∂z
∂ ∂ ∂
= z i − − ir sin θeiφ ψα (r′ ) .
∂x ∂y ∂z
(6.99)
Thus, by using the identity
∂ ∂x ∂ ∂y ∂ ∂z ∂
= + +
∂θ ∂θ ∂x ∂θ ∂y ∂θ ∂z
∂ ∂ ∂
= r cos θ cos φ + sin φ − r sin θ ,
∂x ∂y ∂z
(6.100)
or
∂ ∂ ∂ ∂
r sin θ = r cos θ cos φ + sin φ − , (6.101)
∂z ∂x ∂y ∂θ
m = −l, −l + 1, · · · , l − 1, l . (6.111)
However, as is shown by the claim below, only integer values are allowed for
the case of orbital angular momentum. In view of this result, one may argue
that the existence of spin, which corresponds to half integer values, is in fact
predicted by the commutation relations (6.22) only.
e2πim = 1 , (6.120)
The spherical harmonics Ylm (θ, φ) can be obtained by solving Eqs. (6.116)
and (6.117). However, we will employ an alternative approach, in which in
the first step we find the spherical harmonics Yll (θ, φ) by solving the equation
L+ |l, l = 0 , (6.121)
which is of first order [contrary to Eq. (6.116), which is of the second order].
Using the identity (6.85), which is given by
∂ ∂
r′ | L+ |α = −i eiφ i − cot θ r′ |α , (6.122)
∂θ ∂φ
one has
∂
− l cot θ Fll (θ) = 0 . (6.123)
∂θ
∂ ∂
r′ | L± |α = −i exp (±iφ) ±i − cot θ r′ |α , (6.127)
∂θ ∂φ
∂
e−iφ − − m cot θ Ylm (θ, φ) = l (l + 1) − m (m − 1)Ylm−1 (θ, φ) ,
∂θ
(6.128)
which allows finding Ylm (θ, φ) for all possible values of m provided that
Yll (θ, φ) is given. The normalized spherical harmonics are found using this
method to be given by
;
(−1)l 2l + 1 (l + m)! imφ dl−m
Ylm (θ, φ) = l e (sin θ)−m l−m
(sin θ)2l .
2 l! 4π (l − m)! d (cos θ)
(6.129)
6.6 Problems
1. Let Rı̂ (where i ∈ {x, y, z} ) be the 3×3 rotation matrices (as defined in
the lecture). Show that for infinitesimal angle φ the following holds
where
[Rx̂ (φ) , Rŷ (φ)] = Rx̂ (φ) Rŷ (φ) − Rŷ (φ) Rx̂ (φ) . (6.134)
2. Show that
iJz φ iJz φ
exp Jx exp − = Jx cos φ − Jy sin φ . (6.135)
01 0 −i 1 0
σx = , σy = , σz = . (6.136)
10 i 0 0 −1
a) Show that
(σ · a) (σ · b) = a · b + iσ · (a × b) , (6.137)
where a and b are vector operators which commute with σ , but not
necessarily commute with each other.
b) Show that
iσ · n̂φ φ φ
exp − = 1 cos − iσ · n̂ sin , (6.138)
2 2 2
1 + iα (σ · n̂)
U= , (6.139)
1 − iα (σ · n̂)
where
σ = σ x x̂ + σ y ŷ + σz ẑ (6.140)
n̂ = nx x̂ + ny ŷ + nz ẑ (6.141)
is a unit vector, i.e. n̂· n̂ = 1, and nx , ny , nz and α are all real parameters.
Note that generally for a matrix or an operator A1 ≡ A−1 .
a) show that U is unitary.
b) Show that
dU 2i (σ · n̂)
= U. (6.142)
dα 1 + α2
c) Calculate U by solving the differential equation in the previous sec-
tion.
i.e. two oscillatory magnetic field pulses, both having duration of τ p , are
applied, and the dwell time between these pulses is τ 0 . The normalized
pulse duration αp is defined to be
αp = ω 1 τ p , (6.153)
α0 = ∆ωτ 0 , (6.154)
∆ω = ω 0 − ω ,
and where
|e| B0
ω0 = , (6.156)
me c
|e| B1
ω1 = . (6.157)
me c
At time t = 0 the state is assumed to be given by
|α (t = 0) = |+; ẑ . (6.158)
Calculate the probability P++ (t) to find the system in the state |+; ẑ
at time t > 2τ p + τ 0 . Assume that the normalized detuning is small, i.e.
|δ| ≪ 1, and expand P++ (t) to lowest nonvanishing order in δ for the
case where the normalized pulse duration is taken to be given by
π
αp = . (6.159)
2
12. Find the time evolution of the state vector of a spin 1/2 particle in
a magnetic field along the z direction with time dependent magnitude
B (t) = B (t) ẑ.
13. A magnetic field given by B = B cos (ωt) ẑ, where B is a constant, is
applied to a spin 1/2. At time t = 0 the spin is in state |ψ (t) , which
satisfies
Sx |ψ (t = 0) = |ψ (t = 0) , (6.160)
2
Calculate the expectation value Sz at time t ≥ 0.
14. Consider a spin 1/2 particle. The time dependent Hamiltonian is given
by
4ωSz
H=− , (6.161)
1 + (ωt)2
|ψ = α |+ + β |− ,
H = ωSx , (6.164)
23. Find the condition under which the Hamiltonian of a charged particle in
a magnetic field
1 q 2
H= p− A . (6.170)
2m c
can be written as
1 2 q q2
H= p − p·A+ A2 . (6.171)
2m mc 2mc2
24. Consider a point particle having mass m and charge q moving under the
influence of electric field E and magnetic field B, which are related to
the scalar potential ϕ and to the vector potential A by
1 ∂A
E = −∇ϕ − , (6.172)
c ∂t
and
B=∇×A. (6.173)
26. Find the energy spectrum of a charged particle having mass m and charge
q moving in uniform and time-independent magnetic field B = Bẑ and
electric field E = Ex̂.
27. Consider a particle having mass m and charge e moving in xy plane under
the influence of the potential U (y) = 12 mω 20 y 2 . A uniform and time-
independent magnetic field given by B = Bẑ is applied perpendicularly
to the xy plane. Calculate the eigenenergies of the particle.
28. Consider a particle with charge q and mass µ confined to move on a circle
of radius a in the xy plane, but is otherwise free. A uniform and time
independent magnetic field B is applied in the z direction.
a) Find the eigenenergies.
b) Calculate the current Jm for each of the eigenstates of the system.
29. The Hamiltonian of a non isotropic rigid rotator is given by
L2x L2y L2
H= + + z , (6.177)
2Ixy 2Ixy 2Iz
iLz φ iLz φ
Az (φ) = ψ 0 | exp H exp − |ψ0 . (6.188)
iLx φ iLx φ
Ax (φ) = ψ0 | exp H exp − |ψ0 . (6.189)
where both ω and η are real constants, S1 and S2 are the angular mo-
mentum vector operators of the first and second spin respectively, i.e.
S1 = (S1x , S1y , S1z ) and S2 = (S2x , S2y , S2z ). At time t = 0 the first par-
ticle is in an eigenstate of the operator S1z with eigenvalue + /2 and the
second one is in an eigenstate of the operator S2z with eigenvalue − /2.
Calculate the expectation values S1z (t) and S2z (t) at time t > 0.
38. Consider a system in a common eigenvector |j, m of the angular momen-
tum operators J2 and Jz . A measurement of the operator Jn̂ = n̂ · J is
being performed, where n̂ = (cos ϕ sin θ, sin ϕ sin θ, cos θ) ' is a unit
( vector.
2
Calculate the expectation value Jn̂ and the variance (∆Jn̂ ) .
39. Consider a harmonic oscillator having angular resonance frequency ω and
mass m. The operator S (ξ, ϕ) is defined by [compare with Eq. (5.83)]
is the annihilation operator [see Eq. (5.9)]. Show that S (ξ, ϕ) can be
factorized according to
eiϕ †2
S (ξ, ϕ) = exp a tanh ξ
2
log (cosh ξ)
× exp − aa† + a† a
2
e−iϕ 2
× exp − a tanh ξ .
2
(6.198)
6.7 Solutions
1. By cyclic permutation of
cos φ − sin φ 0
Rẑ = sin φ cos φ 0 , (6.201)
0 0 1
one has
1 0 0
Rx̂ = 0 cos φ − sin φ , (6.202)
0 sin φ cos φ
cos φ 0 − sin φ
Rŷ = 0 1 0 . (6.203)
sin φ 0 cos φ
On one hand
1 − [Rx̂ (φ) , Rŷ (φ)]
1 −1 + cos2 φ sin φ − sin φ cos φ
= 1 − cos2 φ 1 sin φ cos φ − sin φ
sin φ − sin φ cos φ sin φ cos φ − sin φ 1
2
1 −φ 0
= φ2 1 0 + O φ3 .
0 0 1
(6.204)
On the other hand
cos φ2 − sin φ2 0 1 −φ2 0
2
Rẑ φ = sin φ2 cos φ2
0 = φ2 1 0 + O φ3 , (6.205)
0 0 1 0 0 1
thus
one has
iJz φ iJz φ
exp Jx exp −
2
iφ 1 iφ
= Jx + [Jz , Jx ] + [Jz , [Jz , Jx ]]
2!
3
1 iφ
+ [Jz , [Jz , [Jz , Jx ]]] + · · ·
3!
1 2 1
= Jx 1 − φ + · · · − Jy φ − φ3 + · · ·
2! 3!
Jx cos φ − Jy sin φ .
(6.209)
3. The components of the Pauli matrix vector σ are given by:
01 0 −i 1 0
σx = , σy = , σz = . (6.210)
10 i 0 0 −1
(σ · n̂)2 = 1 , (6.214)
thus with the help of the Taylor expansion of the functions cos (x)
and sin (x) one finds
thus
cos θ sin θe−iϕ
σ · n̂ = . (6.217)
sin θeiϕ − cos θ
λ+ + λ− = Tr (σ · n̂) = 0 , (6.218)
and
thus
λ± = ±1 . (6.220)
θ θ
cos2 − sin2 = cos θ . (6.225)
2 2 2 2
6. In general, note that all smooth functions of the matrix (σ · n̂) commute,
a fact that greatly simplifies the calculations.
a) The following holds
1
= 1 + iα (σ · n̂) + [iα (σ · n̂)]2 + · · · , (6.226)
1 − iα (σ · n̂)
thus
†
1
= 1 − iα (σ · n̂) + [(−i) α (σ · n̂)]2 + · · ·
1 − iα (σ · n̂)
1
= ,
1 + iα (σ · n̂)
(6.227)
therefore
1 + iα (σ · n̂) 1 − iα (σ · n̂)
UU† = =1, (6.228)
1 − iα (σ · n̂) 1 + iα (σ · n̂)
and similarly U † U = 1.
b) Exploiting again the fact that all smooth functions of the matrix
(σ · n̂) commute and using Eq. (6.214) one has
dU [1 − iα (σ · n̂)] (σ · n̂) + [1 + iα (σ · n̂)] (σ · n̂)
=i
dα [1 − iα (σ · n̂)]2
2 (σ · n̂)
=i
[1 − iα (σ · n̂)]2
2 (σ · n̂) 1 + iα (σ · n̂)
=i
[1 − iα (σ · n̂)] [1 + iα (σ · n̂)] 1 − iα (σ · n̂)
2i (σ · n̂)
= U.
1 + α2
(6.229)
c) By integration one has
α
dα′
U = U0 exp 2i (σ · n̂)
0 1 + α′2
U0 exp 2i (σ · n̂) tan−1 α ,
(6.230)
where U0 is a the matrix U at α = 0. With the help of Eq. (6.138)
one thus finds that
, -
U = U0 1 cos 2 tan−1 α + iσ · n̂ sin 2 tan−1 α , (6.231)
Using the identities
1 − α2
cos 2 tan−1 α = , (6.232)
1 + α2
2α
sin 2 tan−1 α = , (6.233)
1 + α2
1 − α2 2α
U= 2
+ iσ · n̂ . (6.234)
1+α 1 + α2
7. With the help of Eq. (4.38) one finds that
d A1
= −ω A2 , (6.235)
dt
d A2
= ω A1 , (6.236)
dt
or in a matrix form
d A1 A1
= −iωσ , (6.237)
dt A2 A2
0 −i
σ= . (6.238)
i 0
A1 (t) A1 (t = 0)
= exp (−iωσ) . (6.239)
A2 (t) A2 (t = 0)
A |α = λ |α , (6.244)
−λ z ∗
0 = det , (6.246)
z −λ
thus λ = ± |z|.
9. In terms of Pauli matrices
H=E
˙ a σ0 + ∆σx + Ed σz , (6.247)
where
EL + ER EL − ER
Ea = , Ed = , (6.248)
2 2
and
10 01 1 0
σ0 = , σx = , σz = . (6.249)
01 10 0 −1
iσ · n̂φ φ φ
exp − = cos − iσ · n̂ sin , (6.250)
2 2 2
iEa σ 0 t i (∆σ x + Ed σz ) t
= exp − exp −
! "
iEa t iσ · n̂ ∆2 + Ed2 t
= exp − exp − ,
(6.251)
where
(∆, 0, Ed )
σ · n̂ = σ · , (6.252)
∆2 + Ed2
thus
! "
iEa t ∆2 + Ed2 t ∆σ x + Ed σz t ∆2 + Ed2
u (t) = exp − cos −i sin .
∆2 + Ed2
(6.253)
The probability pR (t) is thus given by
2
pR (t) = | R| u (t) |ψ (t = 0) |
= | R| u (t) |L |2
.
EL −ER 2
∆2 2
t ∆2 + 2
= sin .
EL −ER 2
∆2 + 2
(6.254)
10. The Hamiltonian is given by
H = ω 0 Sz + ω1 (cos (ωt) Sx + sin (ωt) Sy ) , (6.255)
where
|e| B0
ω0 = , (6.256)
me c
|e| B1
ω1 = . (6.257)
me c
The matrix representation in the basis {|+ , |− } (where |+ = |+; ẑ
and |− = |−; ẑ ) is found using Eqs. (6.70), (6.75) and (6.76)
ω0 ω 1 exp (−iωt)
H=
˙ . (6.258)
2 ω 1 exp (iωt) −ω 0
The Schrödinger equation is given by
d
i |α = H |α . (6.259)
dt
It is convenient to express the general solution as
iωt iωt
|α (t) = b+ (t) exp − |+ + b− (t) exp |− . (6.260)
2 2
Substituting into the Schrödinger equation yields
iωt
d e− 2 0 b+
i iωt
dt 0 e 2 b−
iωt
1 ω 0 ω 1 e−iωt e− 2 0 b+
= ,
ω1 eiωt −ω0
iωt
2 0 e 2 b−
(6.261)
or
iωt iωt
iω −e− 2 0 b+ e− 2 0 ḃ+
iωt + iωt
2 0 e 2 b− 0 e 2 ḃ−
iωt
−i ω 0 ω1 e−iωt e− 2 0 b+
= .
2 ω 1 eiωt −ω 0 iωt
0 e 2 b−
(6.262)
By multiplying from the left by
iωt
e 2 0
− iωt
0 e 2
one has
iω −1 0 b+ ḃ+ −i ω0 ω1 b+
+ = , (6.263)
2 0 1 b− ḃ− 2 ω 1 −ω0 b−
or
ḃ+ Ω b+
i = , (6.264)
ḃ− 2 b−
where
∆ω ω 1
Ω= = ∆ωσ z + ω 1 σ x , (6.265)
ω 1 −∆ω
and
∆ω = ω 0 − ω . (6.266)
At time t = 0
b+ (0) 1
= . (6.267)
b− (0) 0
(6.268)
where
.
ω 21 + (∆ω)2 t
θ= . (6.269)
2
iωt iωt
|α (t) = b+ (t) exp − |+ + b− (t) exp |− . (6.271)
2 2
For time periods where g (t) is constant the time evolution is governed
by Eq. (6.268). Thus at time t = 2τ p + τ 0 one has
b+ (2τ p + τ 0 ) 1
= Mp M0 Mp , (6.272)
b− (2τ p + τ 0 ) 0
where
cos θp − iδ
√sin θ2p i sin θp
−√ 2
1+δ 1+δ
Mp = i sin θ p iδ sin θp
, (6.273)
−√ 2
cos θp + √
1+δ 1+δ 2
α0
e−i 2 0
M0 = α0 , (6.274)
0 ei 2
and where
1 + δ 2 αp
θp = . (6.275)
2
Thus the probability P++ (2τ p + τ 0 ) is given by
P++ (2τ p + τ 0 ) = |b+ (2τ p + τ 0 )|2
! "2 2
iδ sin θp eiα0 sin2 θ p
= cos θp − −
1 + δ2 1 + δ2
2
iδ sin (2θ p ) 1 − eiα0 sin2 θp
= cos (2θp ) − + .
1 + δ2 1 + δ2
(6.276)
For the case where αp = π/2 one has to lowest nonvanishing order in δ
1 − cos α0
P++ (2τ p + τ 0 ) = + δ sin α0 + O δ 2 . (6.277)
2
Note that the probability P++ (t) is unchanged for t > 2τ p + τ 0 .
d |α
i = H |α , (6.278)
dt
where
H = ωSz , (6.279)
and
|e| B (t)
ω (t) = . (6.280)
me c
In the basis of the eigenvectors of Sz one has
|α = c+ |+ + c− |− , (6.281)
and
where
Sz |± = ± |± , (6.283)
2
thus one gets 2 decoupled equations
iω
ċ+ = − c+ , (6.284)
2
iω
ċ− = c− . (6.285)
2
The solution is given by
t
i
c± (t) = c± (0) exp ∓ ω (t′ ) dt′
2 0
t
i |e|
= c± (0) exp ∓ B (t′ ) dt′ .
2me c 0
(6.286)
13. At time t = 0
1
|ψ (t = 0) = √ (|+ + |− ) . (6.287)
2
Using the result of the previous problem and the notation
eB
ωc = , (6.288)
mc
one finds
t
t
1 iω c iω c
|ψ (t) = √ exp − cos (ωt′ ) dt′ |+ + exp cos (ωt′ ) dt′ |−
2 2 2
0 0
1 iω c sin ωt iωc sin ωt
= √ exp − |+ + exp |− ,
2 2ω 2ω
(6.289)
thus
(6.293)
Setting an initial condition u (t = 0) = 1 yields
4iSz
u (t) = exp tan−1 (ωt) . (6.294)
θ ϕ θ ϕ
|+; S · n̂ = cos e−i 2 |+ + sin ei 2 |− . (6.298)
2 2
The operator Sz is written as
θ θ
+; S · n̂| Sz |+; S · n̂ = cos2 − sin 2 = cos θ . (6.300)
2 2 2 2
|ψ = |+; S · n̂ , (6.302)
θ+ iϕ θ+ iϕ+
|+; S · n̂ = cos exp − + |+ +sin exp |− , (6.303)
2 2 2 2
θ+ α
ctg = , (6.304)
2 β
and
|ψ = |−; S · n̂ , (6.306)
where
θ− iϕ θ− iϕ−
|−; S · n̂ = − sin exp − − |+ +cos exp |− , (6.307)
2 2 2 2
one finds
θ− α
tan = , (6.308)
2 β
ϕ− = arg (β) − arg (α) + π . (6.309)
θ ϕ θ ϕ
|+; S · û = cos e−i 2 |+ + sin ei 2 |− , (6.311)
2 2
θ ϕ θ ϕ
|−; S · û = − sin e−i 2 |+ + cos ei 2 |− , (6.312)
2 2
Thus in the time interval 0 < t < τ the state vector is given by
iγB0 t iγB0 t
|α = |+; S · û +; S · û |+ exp + |−; S · û −; S · û |+ exp −
2 2
θ ϕ iγB0 t θ ϕ iγB0 t
= |+; S · û cos ei 2 exp − |−; S · û sin ei 2 exp −
2 2 2 2
θ iγB0 t θ iγB0 t
= eiϕ cos2 exp + sin2 exp − |+
2 2 2 2
θ θ iγB0 t iγB0 t
+ sin cos exp − exp − |−
2 2 2 2
1 + cos θ iγB0 t 1 − cos θ iγB0 t
= eiϕ exp + exp − |+
2 2 2 2
γB0 t
+i sin θ sin |−
2
γB0 t γB0 t γB0 t
= eiϕ cos + i cos θ sin |+ + i sin θ sin |− .
2 2 2
(6.313)
Thus for t > τ
γB0 τ
P− (t) = sin2 θ sin2 . (6.314)
2
iHt iγB0 t
u (t) = exp − = exp (σ · û) . (6.317)
2
(6.318)
thus for t > τ
2
1 γB0 τ
P− (t) = 0 1 u (t) = sin2 θ sin2 . (6.319)
0 2
iHt iωt
u (t) = exp − ˙ exp −
= (x̂ · σ) . (6.322)
2
2 2 θ
P (+, +|+) = | +|+; û | | +|+; û | = cos4 , (6.331a)
2
θ
P (+, −|+) = | +|−; û |2 | +|−; û |2 = sin 4 , (6.331b)
2
2 2 θ
P (−, −|−) = | −|−; û | | −|−; û | = cos4 , (6.331c)
2
2 2 θ
P (−, +|−) = | −|+; û | | −|+; û | = sin 4 , (6.331d)
2
thus independently on what was the result of the first measurement one
has
θ θ 1
psame = cos4 + sin 4 = 1 − sin2 θ . (6.332)
2 2 2
20. The Hamiltonian is given by
H = −µ · B . (6.333)
Using Eq. (4.38) for µz one has
d µz 1
= [µz , H]
dt i
γ2
=− Bx [Sz , Sx ] + By [Sz , Sy ]
i
= γ 2 By Sx − Bx Sy
= γ (µ × B) · ẑ .
(6.334)
Similar expressions are obtained for µx and µy that together can be
written in a vector form as
d
µ (t) = γ µ (t) × B (t) . (6.335)
dt
21. Using Eq. (6.137), which is given by
(σ · a) (σ · b) = a · b + iσ · (a × b) , (6.336)
one has
3 q 42 q 2
σ· p− A = p− A + iσ · ((p − qA) × (p − qA))
c c
q 2 q
= p− A − i σ · (A × p + p × A) .
c c
(6.337)
The z component of the term (A × p + p × A) can be expressed as
(A × p + p × A) · ẑ = Ax py − Ay px + px Ay − py Ax
= [Ax , py ] − [Ay , px ] ,
(6.338)
dAx dAy
(A × p + p × A) · ẑ = i − = −i (∇ × A) · ẑ . (6.339)
dy dx
and
p·A=A·p, (6.344)
or
∇·A=0 . (6.347)
r′ | H |α = E r′ |α . (6.349)
r′ |α = ψ (r′ ) (6.350)
for the wavefunction together with Eqs. (3.23) and (3.29) one has
1 q 2
−i ∇ − A + qϕ ψ (r′ ) = Eψ (r′ ) . (6.351)
2m c
25. The Hamiltonian is given by
2 2
p− ec A p2 + p2z py − eBx
H= = x + c
2m 2m 2m
p2 1 cpy 2 p2
= x + mω 2c x − + z ,
2m 2 eB 2m
(6.352)
where
eB
ωc = . (6.353)
mc
Using the clue
one finds that the time independent Schrödinger equation for the wave
function χ (x) is thus given by
2
p̂2x 1 c ky 2 2
kz
+ mω 2c x − χ (x) = E − χ (x) , (6.355)
2m 2 eB 2m
2 2
p̂2x 1 1 ky + 2 kz2
+ mω 2c (x − x̃0 )2 − mω 2c x̃20 + ϕ (x) = Eϕ (x) ,
2m 2 2 2m
(6.363)
mc2 q ky
x̃0 = qE + B . (6.364)
q2 B2 mc
where
p̂ = −i ∇ .
where p̂x = −i ∂/∂x and p̂y = −i ∂/∂y. By substituting the trial wave-
function
or
p̂2y 2 2
k 1 eB k
+ + mω2c0 y 2 − y χ (y) = Eχ (y) , (6.370)
2m 2m 2 mc
1 2 q q2
H= p − p·A+ A2 . (6.373)
2µ µc 2µc2
2 2
1 ∂ ∂ψ 1 ∂2ψ ∂2ψ i qB ∂ψ q2 ρB
− ρ + + + + ψ = Eψ .
2µ ρ ∂ρ ∂ρ ρ2 ∂φ2 ∂z 2 2µc ∂φ 2µc2 2
(6.374)
Φ = Bπa2 , (6.378)
where
2 l (l + 1) 1 1
El,m = + − m2 . (6.389)
2Ixy 2Iz 2Ixy
b)
1
010 2
1 √1 1 1 0 1 √1 =
Lx = √ 2 2 2 2
. (6.398)
2 010 1
2
c)
√1
010 − 2
Lx = √ − √12 0 √1
2
1 0 1 0 = 0 . (6.399)
2 010 √1
2
d)
exp (−iφ) 0 0
iφLz
Dẑ (φ) = exp − ˙
= 0 1 0 . (6.400)
0 0 exp (iφ)
e) In general
i (dφ) L · n̂ i (dφ) L · n̂
Dn̂ (dφ) = exp − =1− + O (dφ)2 ,
(6.401)
thus
1 − i(dφ)
√
2
0
i(dφ)
˙
Dx̂ (dφ) = − 2
√ 1 − i(dφ)
√ + O (dφ)
2
2
. (6.402)
0 − i(dφ)
√
2
1
31. Using
Lz = xpy − ypx , (6.403)
0
x= ax + a†x , (6.404)
2mω
0
y= ay + a†y , (6.405)
2mω
0
m ω
px = i −ax + a†x , (6.406)
2
0
m ω
py = i −ay + a†y , (6.407)
2
one finds
i , -
Lz = ax + a†x −ay + a†y − ay + a†y −ax + a†x
2
= i ax a†y − a†x ay .
(6.408)
a) Thus
b) Using
, the†commutation
- relations
ax , ax = 1 , (6.410)
, -
ay , a†y = 1 , (6.411)
one finds
Lx = 0 . (6.419)
where
2
l (l + 1) m2 2 2
I1
El,m = + = l (l + 1) − m2 + m2 . (6.426)
2I1 2Ie 2I1 I2
iLz φ iLz φ
exp H exp − =H, (6.427)
L+ + L−
Lx = . (6.429)
2
In general
L+ |l, m = l (l + 1) − m (m + 1) |l, m + 1 , (6.430)
L− |l, m = l (l + 1) − m (m − 1) |l, m − 1 , (6.431)
thus
and consequently
iLx φ
exp − |ψ0 = |ψ0 , (6.433)
thus
Any operator of the first particle commutes with any operator of the
second one thus
S2 = S21 + S22 + 2S1 · S2
= S21 + S22 + 2 (S1x S2x + S1y S2y + S1z S2z ) .
(6.446)
In terms of the operators S1± and S2± , which are related to S1x , S2x ,
S1y and S2y by
With
, 2 the- help
, Eqs. (6.24) and (6.41) one finds that -
S , Sz = S21 + S22 + S1+ S2− + S1− S2+ + 2S1z S2z , S1z + S2z
= [S1+ S2− + S1− S2+ , S1z + S2z ]
= [S1+ , S1z ] S2− + [S1− , S1z ] S2+ + S1+ [S2− , S2z ] + S1− [S2+ , S2z ]
= (−S1+ S2− + S1− S2+ + S1+ S2− − S1− S2+ ) ,
(6.450)
thus
, 2 -
S , Sz = 0 . (6.451)
and
|+, + 100 0 |+, +
|+, − 0 0 0 0 |+, −
Sz
|−, + = 0 0 0 0 |−, +
.
(6.453)
|−, − 0 0 0 −1 |−, −
and
(2/ )2 (S1 · û1 ) (S2 · û2 )
= − sin θ1 sin θ 2 cos (ϕ1 − ϕ2 − δ) − cos θ1 cos θ2 .
(6.465)
The above result (6.465) can be rewritten as
where
Sn± = Snx ± iSny , (6.469)
one finds that the Hamiltonian (6.195) can be rewritten as
ω S1+ S2− + S1− S2+
H= + (1 + η) S1z S2z . (6.470)
2
Let |η 1 , η2 be a normalized common eigenvectors of the operator S1z
and S2z with eigenvalues η1 ( /2) and η2 ( /2), respectively, where η1 ∈
{+, −} and η2 ∈ {+, −}. The following holds [see Eqs. (6.70), (6.71),
(6.72) and (6.470)]
|+, + |+, +
|+, − |+, −
H
|−, + = H |−, + ,
(6.471)
|−, − |−, −
where
1+η 0 0 0
ω 0 −1 − η 2 0 .
H= (6.472)
4 0 2 −1 − η 0
0 0 0 1+η
The 4 × 4 matrix H can be diagonalized using the transformation
E1,1 0 0 0
0 E1,0 0 0
U −1 HU = 0
, (6.473)
0 E0,0 0
0 0 0 E1,−1
where the unitary matrix U is given by
1 0 0 0
0 √1 √1 0
U = 2 2
0 √1 − √1 0 , (6.474)
2 2
0 0 0 1
and where the eigenenergies are given by
(1 + η) ω
E1,1 = , (6.475)
4
(1 − η) ω
E1,0 = , (6.476)
4
(−3 − η) ω
E0,0 = , (6.477)
4
(1 + η) ω
E1,−1 = . (6.478)
4
where the states |S, M are the common eigenvectors of the operators
S2 = (S1 + S2 )2 and Sz = S1z + S2z given by Eqs. (6.454), (6.455),
(6.456) and (6.457). The initial state at time t = 0 can be expressed as
|0, 0 + |1, 0
|ψ (t = 0) = |+, − = √ , (6.480)
2
and thus for a general time t one has [see Eq. (4.14)]
iE0,0 t iE1,0 t
e− |0, 0 + e− |1, 0
|ψ (t) = √ . (6.481)
2
The following holds
sin2 θ (J+ J− + J− J+ )
j, m| Jn̂2 |j, m = j, m| + cos2 θJz2 |j, m
4
sin2 θ J2 − Jz2
= j, m| + cos2 θJz2 |j, m
2
sin2 θ j (j + 1) − m2
= 2 + cos2 θm2 ,
2
(6.489)
thus the expectation value is given by Jn̂ = m cos θ and the variance
is given by
' ( + 1) − m2
2 2 j (j
(∆Jn̂ ) = sin2 θ . (6.490)
2
39. Define the vector of operators Σ = (Σx , Σy , Σz ), where
a†2 − a2
Σx = , (6.491)
2
a†2 + a2
Σy = −i , (6.492)
2
† †
aa + a a
Σz = . (6.493)
2
Using Eq. (5.13), which is given by
, †-
a, a = 1 , (6.494)
where
1 a†2
Σ+ = (Σx + iΣy ) = , (6.500)
2 2
1 a2
Σ− = (Σx − iΣy ) = − . (6.501)
2 2
where
1 01
σ+ = (σ x + iσy ) = , (6.503)
2 00
1 00
σ− = (σ x − iσy ) = , (6.504)
2 10
is given by
s (ξ, ϕ)
= exp eiϕ tanh ξσ+
× exp (− log (cosh ξ) σ z )
× exp e−iϕ tanh ξσ − .
(6.507)
The above expression for s (ξ, ϕ) yields a similar identity for the operator
S (ξ, ϕ)
eiϕ †2
S (ξ, ϕ) = exp a tanh ξ
2
log (cosh ξ)
× exp − aa† + a† a
2
e−iϕ 2
× exp − a tanh ξ .
2
(6.508)
40. With the help of Eqs. (5.338) and (5.105) one finds that
∞
1
Q (µ) = √ dx′
x0 πµ
−∞
x′2 √ x′ † a†2
× exp − 2 + 2 a −
2µ2 x0 µx0 2
×: exp −a† a :
x′2 √ x′ a2
× exp − + 2 a − .
2x20 x0 2
(6.509)
−
1+ 12 x′2
µ
+
2 a+ a
µ
x′
−
(a+a† )2
2x2 x0 2
×: e 0 : ,
(6.510)
or with the help of the identity (5.139) one finds that
† 2
0 a+ a
µ
−
(a+a† )2
2µ 1+ 12 2
Q (µ) = :e µ :
1 + µ2
0 2
2µ 1−µ2 a†2 −a2
− (1−µ) aa†
= 2
: e 1+µ2 2 1+µ2 : .
1+µ
(6.511)
µ = e−ξ , (6.512)
Using also
0
1 1
= e− 2 log(cosh ξ)
cosh ξ
and
1 aa† + a† a
a† a + =
2 2
one has
log(cosh ξ)
Q (µ) = e
tanh ξ †2
2 a
e− 2 (aa† +a† a) e− tanh
2
ξ 2
a
. (6.518)
r= x2 + y 2 + z 2 (7.1)
from the origin. The Hamiltonian is given by
p2
H= + V (r) . (7.2)
2m
Exercise 7.0.1. Show that
[H, Lz ] = 0 , (7.3)
, -
H, L2 = 0 . (7.4)
Solution 7.0.1. Using
[xi , pj ] = i δ ij , (7.5)
Lz = xpy − ypx , (7.6)
one has
, 2 - , - , - , -
p , Lz = p2x , Lz + p2y , Lz + p2z , Lz
, - , -
= p2x , xpy − p2y , ypx
= i (−2px py + 2py px )
=0,
(7.7)
and
, - , - , - , -
r2 , Lz = x2 , Lz + y 2 , Lz + z 2 , Lz
, - , -
= −y x2 , px + y 2 , py x
=0.
(7.8)
Thus Lz commutes with any smooth function of r2 , and consequently
[H, Lz ] = 0. In
, a similar
- way one can show that [H, Lx ] = [H, Ly ] = 0,
and therefore H, L2 = 0.
Chapter 7. Central Potential
and
H, L2 = 0 . (7.10)
L2 = r2 p2 − (r · p)2 + i r · p . (7.13)
or
one has
i
L2z = x2 p2y + y 2 p2x − xpx ypy − ypy xpx + (xpx + ypy )
2
1 2 2
+ x px − xpx xpx + y 2 p2y − ypy ypy .
2
(7.17)
By cyclic permutation one obtains similar expression for L2x and for L2y . Com-
bining these expressions lead to
L2 = L2x + L2y + L2z
i 1 2 2
= y 2 p2z + z 2 p2y − ypy zpz − zpz ypy + (ypy + zpz ) + y py − ypy ypy + z 2 p2z − zpz zpz
2 2
i 1 2 2
+z 2 p2x + x2 p2z − zpz xpx − xpx zpz + (zpz + xpx ) + z pz − zpz zpz + x2 p2x − xpx xpx
2 2
i 1 2 2
+x2 p2y + y 2 p2x − xpx ypy − ypy xpx + (xpx + ypy ) + x px − xpx xpx + y 2 p2y − ypy ypy
2 2
= x2 + y2 + z 2 p2x + p2y + p2z − (xpx + ypy + zpz )2 + i (xpx + ypy + zpz )
= r2 p2 − (r · p)2 + i r · p .
(7.18)
Exercise 7.1.2. Show that
1 ∂2 ′ ′ 1
r′ | p2 |α = − 2
r r |α − r′ | L2 |α . (7.19)
r′ ∂r′2 2 r′2
r′ | p |α = ∇ r′ |α , (7.22)
i
one finds that
r′ | L2 |α = r′ | r2 p2 |α − r′ | (r · p)2 |α + i r′ | r · p |α . (7.23)
The following hold
∂
r′ | r · p |α = −i r′ · ∇ r′ |α = −i r′ r′ |α , (7.24)
∂r′
2
∂
r′ | (r · p)2 |α = − 2
r′ r′ |α
∂r′
2 ∂2 ∂
=− r′2 ′2 + r′ ′ r′ |α ,
∂r ∂r
(7.25)
r′ | r2 p2 |α = r′2 r′ | p2 |α , (7.26)
thus
∂2 2 ∂ 1
r′ | p2 |α = − 2
+ r′ |α − r′ | L2 |α , (7.27)
∂r′2 r′ ∂r′ 2 r′2
or
1 ∂2 ′ ′ 1
r′ | p2 |α = − 2
r r |α − r′ | L2 |α . (7.28)
r′ ∂r′2 2 r′2
r′ | H |α = E r′ |α , (7.29)
where the Hamiltonian H is given by Eq. (7.2), can thus be written using the
above results as
− 2 1 ∂2 ′ ′ 1
r′ | H |α = r r |α − r′ | L2 |α +V (r′ ) r′ |α . (7.30)
2m r′ ∂r′2 2 r ′2
r′ | L2 |α = 2
l (l + 1) ϕ (r′ ) . (7.32)
− 2 1 d2 1
2
rR (r) − 2 l (l + 1) R (r) + V (r) R (r) = ER (r) . (7.33)
2m r dr r
The above equation, which is called the radial equation, depends on the quan-
tum number l, however, it is independent on the quantum number m. The
different solutions for a given l are labeled using the index k
− 2 1 d2 1
2
rRkl − 2 l (l + 1) Rkl + V Rkl = ERkl . (7.34)
2m r dr r
It is convenient to introduce the function ukl (r), which is related to Rkl (r)
by the following relation
1
Rkl (r) = ukl (r) . (7.35)
r
− 2 d2
+ Veff (r) ukl (r) = Ekl ukl (r) , (7.36)
2m dr2
The wave functions ϕklm (r) represent common eigenstates of the operators
H, Lz and L2 , which are denoted as |klm and which satisfy the following
relations
and
H |klm = Ekl |klm , (7.42)
L2 |klm = l (l + 1) 2 |klm , (7.43)
Lz |klm = m |klm . (7.44)
The following claim reveals an important property of the radial wavefunc-
tion near the origin (r = 0):
Claim. If the potential energy V (r) does not diverge more rapidly than 1/r
near the origin then
Proof. Consider the case where near the origin u (r) has a dominant power
term having the form rs (namely, all other terms are of order higher than s,
and thus become negligibly small for sufficiently small r). Substituting into
Eq. (7.36) and keeping only the dominant terms (of lowest order in r) lead
to
− 2 l (l + 1) 2
s (s − 1) rs−2 + rs−2 = 0 , (7.46)
2m 2m
thus s = −l or s = l + 1. However, the solution s = −l for l ≥ 1 must
be rejected since for this case the normalization condition (7.39) cannot be
satisfied as the integral diverges near r = 0. Moreover, also for l = 0 the
solution s = −l must be rejected. For this case ϕ (r) ≃ 1/r near the origin,
however, such a solution contradicts Eq. (7.30), which can be written as
2
− ∇2 ϕ (r) + V (r) ϕ (r) = Eϕ (r) . (7.47)
2m
since
1
∇2 = −4πδ (r) . (7.48)
r
We thus conclude that only the solution s = l + 1 is acceptable, and conse-
quently limr→0 u (r) = 0.
Exercise 7.3.1. Consider two point particles having mass m1 and m2 re-
spectively. The potential energy V (r) depends only on the relative coordi-
nate r = r1 − r2 . Show that the Hamiltonian of the system in the center of
mass frame is given by
p2
H= + V (r) , (7.49)
2µ
where the reduced mass µ is given by
m1 m2
µ= . (7.50)
m1 + m2
M = m1 + m2 , (7.55)
Note that the Euler Lagrange equation for the coordinate r0 yields that r̈0 = 0
(since the potential is independent on r0 ). In the center of mass frame r0 = 0.
The momentum canonically conjugate to r is given by
∂L
p= . (7.57)
∂ ṙ
Thus the Hamiltonian is given by
p2
H = p · ṙ−L = + V (r) . (7.58)
2µ
For the case of hydrogen atom the potential between the electron having
charge −e and the proton having charge e is given by
e2
V (r) = − . (7.59)
r
Since the proton’s mass mp is significantly larger than the electron’s mass
me (mp ≃ 1800me ) the reduced mass is very close to me
me mp
µ= ≃ me . (7.60)
me + mp
10
8
6
4
2
0 1 2 3 4 5
-2
-4
-6
-8
-10
We seek solutions of Eq. (7.67) that represent bound states, for which Ekl
is negative, and thus λkl is a nonvanishing real positive. In the limit ρ → ∞
the potential Vl (ρ) → 0, and thus it becomes negligibly small in comparison
with λkl [see Eq. (7.67)]. Therefore, in this limit the solution is expected to
be asymptotically proportional to e±λkl ρ . To ensure that the solution is nor-
malizable the exponentially diverging solution e+λkl ρ is excluded. Moreover,
as we have seen above, for small ρ the solution is expected to be proportional
to ρl+1 . Due to these considerations we express ukl (r) as
Substituting into Eq. (7.67) yields an equation for the function y (ρ)
(7.72)
thus
cq 2 [λkl (q + l) − 1]
= . (7.73)
cq−1 q (q + 2l + 1)
We argue below that for physically acceptable solutions y (ρ) must be a poly-
nomial function [i.e. the series (7.71) needs to be finite]. To see this note that
for large q Eq. (7.73) implies that
cq 2λkl
lim = . (7.74)
q→∞ cq−1 q
Similar recursion relation holds for the coefficients of the power series expan-
sion of the function e2λkl ρ
∞
=
e2λkl ρ = c̃q ρq , (7.75)
q=0
where
n=k+l , (7.83)
have the same energy. The index n is called the principle quantum number.
Due to this degeneracy, which is sometimes called accidental degeneracy, it
is common to label the states with the indices n, l and m, instead of k, l and
m. In such labeling the eigenenergies are given by
EI
En = − , (7.84)
n2
where
n = 1, 2, · · · . (7.85)
For a given n the quantum number l can take any of the possible values
l = 0, 1, 2, · · · , n − 1 , (7.86)
and the quantum number m can take any of the possible values
m = −l, −l + 1, · · · , l − 1, l . (7.87)
Note that the factor of 2 is due to spin. The normalized radial wave functions
of the states with n = 1 and n = 2 are found to be given by
3/2
1
R10 (r) = 2 e−r/a0 , (7.89)
a0
3/2
r 1 r
R20 (r) = 2− e− 2a0 , (7.90)
a0 2a0
3/2
1 r r
R21 (r) = √ e− 2a0 . (7.91)
2a0 3a0
The wavefunction ϕn,l,m (r) of an eigenstate with quantum numbers n, l and
m is given by
∞ 1 2π
′ ∗
′ ′ ′ 2
n l m |nlm = dr r Rn′ l′ Rnl d (cos θ) dφ Ylm
′ Ylm = δ n,n′ δ l,l′ δ m,m′ .
0 −1 0
(7.93)
While the index n labels the shell number, the index l labels the sub-shell.
In spectroscopy it is common to label different sub-shells with letters:
l=0 s
l=1p
l=2d
l=3 f
l=4g
7.4 Problems
1 1 2r
ϕ=e + exp − , (7.95)
a0 r a0
|α (t = 0) = A (|2, 1, −1 + |2, 1, 1 ) ,
where r = x2 + y 2 + z 2 .
b) Employ the relation (7.100) for the case of a point particle of mass
m moving in three dimensions under the influence of the potential
V (r), and for the observable
A= r·p+p·r, (7.101)
r′
V (r′ ) = V0 log , (7.103)
r0
where V0 and r0 are positive constants. Calculate the kinetic energy ex-
pectation values Tn of the bounded energy eigenstates of the system.
14. The radial equation for the hydrogen atom (7.61) represents the time
independent Schrödinger equation for a point particle of mass µ moving
in one dimension along the r axis whose Hamiltonian is given by
p2r r
Hl = + Veff , (7.104)
2µ a0
2 l (l + 1)
Veff (ρ) = EI − + , (7.105)
ρ ρ2
a0 ipr l+1 1
al = √ − + . (7.106)
2 r (l + 1) a0
a) Show that
! "
1
Hl = 2EI a†l al − . (7.107)
2 (l + 1)2
3 4
b) Show that the commutation relation al , a†l is given by
3 4 H
l+1 − Hl
al , a†l = . (7.108)
2EI
7.5 Solutions
1. In general the current density is given by Eq. (4.223). For a wavefunction
having the form
ψ (r) = α (r) eiβ(r) , (7.110)
where both α and β are real, one has
J= Im [α (∇α + iα∇β)]
µ
α2
= ∇β
µ
|ψ|2
= ∇β .
µ
(7.111)
a) The wavefunction ϕn,l,m (r) is given by
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) = Rnl (r) Flm (θ) eimφ , (7.112)
where both Rnl and Flm are real, thus
2
ϕn,l,m (r)
Jn,l,m (r) = ∇ (mφ) . (7.113)
µ
In spherical coordinates one has
∂ ∂ ∂
∇ = x̂ + ŷ + ẑ
∂x ∂y ∂z
∂ 1 ∂ 1 ∂
= r̂ + θ̂ + φ̂ ,
∂r r ∂θ r sin θ ∂φ
(7.114)
thus
2
ϕn,l,m (r)
Jn,l,m (r) = m φ̂ . (7.115)
µ r sin θ
L = m ẑ . (7.117)
2 e 2r
ρ = −e ϕ1,0,0 (r) =− 3 exp − . (7.118)
πa0 a0
∇2 ϕ = −4πρ . (7.119)
To verify that the electrostatic potential given by Eq. (7.95) solves this
equation we calculate
1 d2
∇2 ϕ = (rϕ)
r dr2
e d2 r 2r
= + 1 exp −
r dr2 a0 a0
4e exp − a2r0
=
a30
= −4πρ .
(7.120)
Note also that
Thus
∞ ∞
3
r = rf (r) dr = 4a0 x3 exp (−2x) dx = a0 . (7.124)
0 0 2
The most probable value r0 is found from the condition
df 8r0 2r0
0= = 4 exp − (a0 − r0 ) , (7.125)
dr a0 a0
thus
r0 = a0 . (7.126)
c) For this case the probability vanishes due to the orthogonality be-
tween spherical harmonics with different l.
√
5. The normalization constant is chosen to be A = 1/ 2. Since both states
|2, 1, −1 and |2, 1, 1 have the same energy the state |α is stationary.
The following holds
where
√
−2mE
κ= . (7.145)
The boundary condition that is imposed upon u (r) at the origin is u (0) =
0. Since the centrifugal term l (l + 1) 2 /2mr2 is non-negative the ground
state is obtained with l = 0. For that case the solution in the range r ≤ r0
has the form u (r) = sin kr, where k is related to the energy E by
h2 k2
= E + U0 . (7.150)
2m
In the range r > r0 the general solution has the form u (r) = Ae−κr +
Beκr , where
h2 κ2
= −E . (7.151)
2m
A bound state can be obtained provided that E < 0 (to ensure that κ
is real) and B = 0 (to ensure that limr→∞ u (r) = 0; it is assumed that
κ is non-negative). The requirements that both u (r) and du/dr [see Eq.
(4.150)] are continuous at r = r0 yield (for the case B = 0)
sin kr0 = Ae−κr0 , (7.152)
k
cos kr0 = −Ae−κr0 , (7.153)
κ
thus the following must hold
k
tan kr0 = − . (7.154)
κ
Since both k and κ are required to be nonnegative, the above condition
can be satisfied only if tan kr0 ≤ 0, which implies that
0
2m (E + U0 ) π
kr0 = r0 ≥ . (7.155)
h2 2
This together with the requirement that E < 0 yield
0
2mU0 π
2
r0 ≥ , (7.156)
h 2
or
π2 h2
U0 ≥ . (7.157)
8mr02
R (r′ )
ψ (r′ ) = R (r′ ) Yl=0
m=0
θ ′ , φ′ = √ , (7.158)
4π
where the radial function R (r′ ) is an arbitrary normalized function.
10. The Lagrangian is given by
m ṙ21 + ṙ22 1 1 2
L= − mω2 r21 + r22 − mΩ 2 (r1 − r2 ) . (7.159)
2 2 2
In terms of center of mass r0 and relative r coordinates, which are given
by
r1 + r2
r0 = , (7.160)
2
r = r1 − r2 , (7.161)
the Lagrangian is given by
3 4
2 2
m ṙ0 + 12 ṙ + ṙ0 − 12 ṙ
L=
2
2 2
1 1 1 1
− mω 2 r0 + r + r0 − r − mΩ 2 r2
2 2 2 2
m 2ṙ20 + 12 ṙ2 1 1 1
= − mω 2 2r20 + r2 − mΩ 2 r2
2 2 2 2
2 2
M ṙ0 1 µṙ 1
= − M ω2 r20 + − µ ω 2 + 2Ω 2 r2 ,
2 2 2 2
(7.162)
where the total mass M is given by
M = 2m , (7.163)
3
En0x ,n0y ,n0z ,nx ,ny ,nz = ω + n0x + n0y + n0z
2
3
+ ω 2 + 2Ω 2 + nx + ny + nz .
2
(7.165)
11. The normalization condition reads (the coordinate transformation r′ =
r0 ρ is being employed)
1 = α |α
∞ π 2π
2
r′
2 −
= |A| dr′ r′2 e r0
dθ sin θ dφ
0 0 0
4π
∞
2
= 4π |A|2 r03 dρ ρ2 e−ρ
0
e| [A, H] |e = E ( e| A |e − e| A |e ) = 0 . (7.171)
p2
H= + V (r) , (7.172)
2m
and thus the relation (7.100) yields
p2
e| r · p + p · r, |e = − e| [r · p + p · r, V ] |e . (7.173)
2m
p2
2 e| |e = e| (r · ∇V ) |e . (7.175)
2m
13. Let |ψn be a bounded energy eigenstate. With the help of the virial the-
orem (7.102) one finds that the corresponding kinetic energy expectation
values Tn is given by
p2
Tn = ψ n | |ψ
2m n
1
= ψ | (r · ∇V ) |ψn ,
2 n
(7.176)
and thus [see Eq. (7.103)]
p2
Tn = ψ n | |ψ
2m n
V0 ∂ r′
= ψn | r′ ′ log |ψ n
2 ∂r r0
V0
= ψn |ψn
2
V0
= .
2
(7.177)
l+1 l+1
pr , =i , (7.180)
r r2
and thus with the help of Eq. (7.178) one finds that
! "
1 a20 p2r a20 l (l + 1) a0
2EI a†l al − = 2EI + − = Hl .
2 (l + 1)2 2 2 2 r2 r
(7.181)
Hl |E l = E |E l . (7.183)
With the help of Eqs. (7.107) and (7.108) one finds that
Hl+1 al |E l = (Hl+1 − Hl ) al |E l + Hl al |E l
3 4
= 2EI al , a†l al |E l + ([Hl , al ] + al Hl ) |E l
3 3 4 3 4 4
= 2EI al , a†l al + a†l al , al + Eal |E l
3 3 4 3 4 4
= 2EI al , a†l al + a†l , al al + Eal |E l
= Eal |E l ,
(7.184)
thus the state al |E l is an eigenvector of Hl+1 with energy eigenvalue
E. A normalized eigenvector of Hl+1 with energy eigenvalue E, which
is denoted by |E l+1 , is obtained by dividing by the norm of al |E l
(note that |E l is assumed to be normalized)
al |E l
|E l+1 =. , (7.185)
†
l E| al al |E l
Hl ≥ Veff . (7.187)
On the other hand, with the help of Eq. (7.105) it is easy to show
that
EI
Veff (ρ) ≥ − , (7.188)
l (l + 1)
E 1
lmax E| a†l al |E lmax = + =0, (7.189)
2EI 2 (lmax + 1)2
and therefore
EI
E=− , (7.190)
n2
where n = lmax + 1 is a positive integer (recall that the quantum
number l is a nonnegative integer).
0 ≤ wi ≤ 1 , (8.1)
and where
wi = 1 . (8.2)
i
A = Tr (ρA) , (8.8)
where
('
ρ= wi α(i) α(i) (8.9)
i
Proof. Let {|bm } be an orthonormal and complete basis for the vector space
|bm bm | = 1 . (8.10)
m
= wi
i
=1.
(8.12)
Exercise 8.0.3. Show that for any normalized state |β the following holds
0 ≤ β| ρ |β ≤ 1 . (8.13)
On the other hand, according to the Schwartz inequality [see Eq. (2.171)],
which is given by
| u |v | ≤ u |u v |v , (8.15)
one has
' .) *
α(i) |β ≤ β |β α(i) α(i) = 1 . (8.16)
=
Moreover, i wi = 1, thus
' 2
β| ρ |β = wi α(i) |β ≤1. (8.17)
i
and
1 = Tr ρ2 = 2
qm . (8.28)
m
thus
dρ 1
= − [ρ, H] . (8.35)
dt i
This result resembles the equation of motion (4.37) of an observable in the
Heisenberg representation, however, instead of a minus sign on the right hand
side, Eq. (4.37) has a plus sign.
Alternatively, the time evolution of the operator ρ can be expressed in
terms of the time * evolution operator u (t,
* t0 ), which relates the state vector
at time α(i) (t0 ) with its value α(i) (t) at time t [see Eq. (4.4)]
( (
α(i) (t) = u (t, t0 ) α(i) (t0 ) . (8.36)
Z= e−βEi (8.41)
i
e−βH
ρ= . (8.42)
Tr (e−βH )
and
thus
e−βH
ρ= . (8.46)
Tr (e−βH )
As will be demonstrated below [see Eq. (8.491)], the last result for ρ can
also be obtained from the principle of maximum entropy.
8.3 Problems
6. Express the density matrix ρ of a spin 1/2 system in terms of the expec-
tations values σx , σy and σz , where σ x , σy and σz are the Pauli’s
matrices.
7. Let ρ be a density operator that can be expressed in terms of the density
operators ρ1 and ρ2 as
ρ = ηρ1 + (1 − η) ρ2 , (8.50)
where
0<η<1. (8.51)
ρ= d2 α |α α| P (α) , (8.54)
L C
|α = cn |an , (8.57)
n
where f (t) is assumed to have compact support with a peak near the
time of the measurement.
a) Express the vector state of the entire system |Ψ (t) at time t in the
basis of states {|p′ ⊗ |an′ }. This basis spans the Hilbert space of
the entire system (MS and MD). The state |p′ ⊗ |an′ is both, an
eigenvector of A (with eigenvalue an ) and of the momentum p of the
MD (with eigenvalue p′ ).
b) In what follows the final state of the system after the measurement
will be evaluated by taking the limit t → ∞. The outcome of the
measurement of the observable A, which is labeled by A, is deter-
mined by performing a measurement of the momentum variable p of
the MD. The outcome, which is labeled by P, is related to A by
P
A= , (8.61)
pi
where
∞
pi = dt′ f (t′ ) . (8.62)
15. A particle having mass m moves in the xy plane under the influence of
a two dimensional potential V (x, y), which is given by
mω 2 2
V (x, y) = x + y 2 + λmω 2 xy , (8.65)
2
where both ω and λ are real constants. Calculate
) *in thermal equilibrium
at temperature T the expectation values x , x2 .
16. Consider a harmonic oscillator having angular resonance
) frequency ω and
*
mass m. Calculate the correlation function G (t) = x(H) (t) x(H) (0) ,
where x(H) (t) is the Heisenberg representation of the position operator,
for the cases where
a) the oscillator is in its ground state.
b) the oscillator is in thermal equilibrium at temperature T .
17. In general, the Wigner function of a point particle moving in one dimen-
sion is given by
where ρ is the density operator of the system, and where |x′ represents
an eigenvector of the position operator x having eigenvalue x′ , i.e. x |x′ =
x′ |x′ . As can be seen from Eq. (4.250), the Wigner function is the inverse
Weyl transformation of the density operator divided by the factor of 2π.
Consider the case of a point particle having mass m in a potential of a
harmonic oscillator having angular frequency ω. Calculate the Wigner
function W (x′ , p′ ) for the case where the system is in a coherent sate
|α .
18. A particle having mass m is in the ground state of the one-dimensional
2
potential well V1 (x) = (1/2) mω 2 (x − ∆x ) for times t < 0 . At time
t = 0 the potential suddenly changes and becomes V2 (x) = (1/2) mω 2 x2 .
Calculate the Wigner function of the system at times t > 0.
19. Consider a point particle having mass m in a potential of a harmonic
oscillator having angular frequency ω. Calculate the Wigner function
W (x′ , p′ ) for the case where the system is in thermal equilibrium at
temperature T .
20. Consider a point particle having mass m in a potential of a harmonic
oscillator having angular frequency ω. Calculate the Wigner function
W (x′ , p′ ) for the case where the system is in the number state |n = 1 .
21. The Wigner function of a point particle moving in one dimension is
given by Eq. (8.66). Show that the marginal distributions x′ | ρ |x′ and
p′ | ρ |p′ of the position x and momentum p observables, respectively, are
given by
∞
x′ | ρ |x′ = −1
dp′ W (x′ , p′ ) , (8.67)
−∞
∞
p′ | ρ |p′ = −1
dx′ W (x′ , p′ ) . (8.68)
−∞
22. Show that for a pure state the Wigner function is bounded by |W (x′ , p′ )| ≤
1/2π. Note that this bound together with Eqs. (8.67) and (8.68) can be
used to demonstrate the uncertainty principle (3.10).
23. Consider a particle having mass m moving along the x axis under the
influence of the potential V (x). Show that the time evolution of the
Wigner function W (8.66) is governed by
2l
dW =∞
2i ∂ 2l+1 V ∂ 2l+1 W
= {H, W } + 2l+1
, (8.69)
dt l=1 (2l + 1)! ∂ (x′ ) ∂ (p′ )2l+1
a + a† a − a†
X= √ , P = √ , (8.72)
2 i 2
and which satisfy [X, P ] = i [see Eq. (5.13)] and ρ is the density operator.
Show that
∞ 5 6
′ 1 ′ ′′ ′ X ′′ X ′′ ′′
P′
W (X , P ) = dX X − ρ X′ + eiX , (8.73)
2π 2 2
−∞
W (X ′ , P ′ ) = Tr (Υ ρ) , (8.74)
Note that the operator Υ given by Eq. (8.75) is the dimensionless version
of the Weyl kernel (4.46), which defines the Weyl transformation (4.45).
Show that
X ′ + iP ′ X ′ + iP ′
Υ (X ′ , P ′ ) = π −1 D† − √ PD − √ , (8.76)
2 2
where
P= dX ′ |X ′ −X ′ | (8.78)
−∞
a† eiφ + ae−iφ
Xφ = √ , (8.79)
2
where a and a† are annihilation and creation operators [see Eqs. (5.9) and
(5.10)]. Let w Xφ′ be the normalized probability distribution function
of the observable Xφ . The technique of homodyne detection can be used
to measure w Xφ′ for any given value of the phase φ.
a) To generalize Eqs. (8.67) and (8.68) show that the following holds
for any real φ
∞
w Xφ′ = dPφ′ W Xφ′ cos φ − Pφ′ sin φ, Xφ′ sin φ + Pφ′ cos φ .
−∞
(8.80)
b) Show that the Wigner function (8.66) can be extracted from the
measured distributions w Xφ′ for all values of φ.
27. Consider a harmonic oscillator in thermal equilibrium at temperature T ,
whose Hamiltonian is given by
p2 mω2 x2
H= + . (8.81)
2m 2
Calculate the matrix elements x′′ | ρ |x′ of the density operator in the
basis of eigenvectors of the position operator x.
28. Consider a harmonic oscillator having angular resonance frequency ω.
The oscillator is in)thermal* equilibrium at temperature T . Calculate the
expectation value e−iζXφ , where Xφ is given by [see Eq. (8.79)]
a† eiφ + ae−iφ
Xφ = √ , (8.82)
2
a and a† are annihilation and creation operators [see Eqs. (5.9) and (5.10)]
and both
) −iζX * φ and ζ are real. Use your result for the expectation value
e φ
to evaluate the Wigner function of the system.
29. Show that when w Xφ′ is φ independent the following holds
∞
′ 1 ′
W (X , P ) = dζ ζ w̃ (ζ) J0 ζ X ′2 + P ′2 , (8.83)
2π
0
∞
′
w̃ (ζ) = dXφ′ w Xφ′ e−iζXφ . (8.84)
−∞
X′ − ρR X ′ + = X′ − ρ X′ + e 2
,
2 2 2 2
(8.85)
where the dimensionless parameter η characterizes the strength of the
measurement, and where |X ′ represents an eigenvector of the dimen-
sionless position operator X [see Eq. (8.72)] having eigenvalue X ′ , i.e.
X |X ′ = X ′ |X ′ . Express the reduced Wigner function WR (X ′ , P ′ ),
which is given by [see Eq. (8.73)]
∞ 5 6
′ 1′ ′′ ′ X ′′ ′ X ′′ ′′ ′
WR (X , P ) = dX X − ρR X + eiX P , (8.86)
2π 2 2
−∞
where 2 2 2
) r *= x + y + z and where ω is a real constant. Calculate x
and x2 in thermal equilibrium at temperature T .
35. The entropy σ is defined by
σ = − Tr (ρ log ρ) . (8.89)
σ = − Tr (ρ log ρ) . (8.92)
a) Calculate σ.
b) A measurement of Sz is performed. Calculate the entropy after the
measurement.
Consider the case where the density matrix is assumed to satisfy a set of
contrarians, which are expressed as
gl (ρ) = 0 , (8.94)
ρ1 = Tr2 ρ , (8.100)
σ = − Tr (ρ log ρ) . (8.101)
σ1 + σ2 ≥ σ . (8.106)
42. Consider a spin 1/2 in a magnetic field B = Bẑ and in thermal equilib-
rium at temperature T . Calculate the entropy σ, which is defined by
σ = − Tr (ρ log ρ) , (8.107)
1 1 2 (Sx + Sz )
|H H| = 1+ √ , (8.108)
2 2
where 1 is the identity operator, and where Sx and Sz are spin angular
momentum operators. In a measurement of Sz what is the probability
pz+ to obtain the value + /2?
44. Consider a harmonic oscillator of angular frequency ω and mass m in
thermal equilibrium at temperature T . Calculate the entropy σ, which is
defined by
σ = − Tr (ρ log ρ) , (8.109)
8.4 Solutions
= an | (AB)† |an ,
n
(8.110)
and thus, with the help of Eqs. (2.47) and (2.133) and the relations
A† = A and B † = B one finds that
(Tr (AB))∗ = an | B † A† |an
n
= an | AB |an
n
= Tr (AB) ,
(8.111)
and therefore Tr (AB) is real.
2. The Hamiltonian is given by
H = ωSz , (8.112)
where
|e| B
ω= (8.113)
me c
is the Larmor frequency. In the basis of the eigenvectors of Sz
Sz |± = ± |± , (8.114)
2
one has
ω
H |± = ± |± , (8.115)
2
thus
e−Hβ
ρ=
Tr (e−Hβ )
ωβ ωβ
e− 2 |+ +| + e 2 |− −|
= ωβ ωβ ,
e− 2 +e 2
(8.116)
where β = 1/kB T , and therefore with the help of Eqs. (2.102) and (2.103),
which are given by
ωβ
= − tanh ,
2 2
(8.122)
thus
cos θ ωβ
S · û = − tanh . (8.123)
2 2
3. Recall that
1
|±; ŷ = √ (|+ ± i |− ) , (8.124)
2
a) thus
1 1 1 1 −i
ρ=
˙ 1 −i = . (8.125)
2 i 2 i 1
d) and
1 10
ρn =
˙ . (8.127)
2n 01
cos θ2
|ψ (t = 0) =
˙ , (8.128)
sin θ2
thus
cos ωT ωT
2 i sin 2 cos θ2
|ψ (t = T ) =
˙ ωT ωT
i sin 2 cos 2 sin θ2
cos ωT θ ωT θ
2 cos 2 + i sin 2 sin 2
= .
i sin 2 cos 2 + cos 2 sin θ2
ωT θ ωT
(8.133)
a) The probabilities to measured ± /2 are thus given by
ωT θ ωT θ
P+ = cos2 cos2 + sin2 sin2
2 2 2 2
1 + cos (ωT ) cos θ
= , (8.134)
2
and
ωT θ ωT θ
P− = cos2 sin 2 + sin2 cos 2
2 2 2 2
1 − cos (ωT ) cos θ
= . (8.135)
2
b) The density operator is given by
ρ11 = P+ ,
ρ22 = P− ,
ωT θ ωT θ ωT θ ωT θ
ρ21 = cos cos + i sin sin −i sin cos + cos sin
2 2 2 2 2 2 2 2
sin θ i
= − sin ωT cos θ ,
2 2
ρ12 = ρ∗21 .
5. The Hamiltonian is given by
H = −ωSz , (8.136)
where
eB
ω= , (8.137)
me c
thus, the density operator is given by
ω
1 exp 2kB T 0
,
ρ=˙ (8.138)
Z 0 exp − 2kBωT
where
ω ω
Z = exp + exp − . (8.139)
2kB T 2kB T
a) Using
iHt iHt
Sz (t) = exp Sz (0) exp − = Sz (0) , (8.140)
one finds
) * 2
Cz (t) = Sz2 (0) = Tr ρSz2 (0) = . (8.141)
4
b) The following holds
iωSz t iωSz t
Sx (t) = exp − Sx (0) exp
p z
ρ= , (8.146)
z∗ 1 − p
yield
1 1 + σz σx − i σ y
ρ= , (8.150)
2 σ x + i σy 1 − σz
or
1
ρ= (1 + σx σ x + σ y σy + σz σz ) . (8.151)
2
7. The assumption that ρ represents a pure state implies that it can be
expressed as
ρ = |α α| , (8.152)
0 = β| ρ |β = η β| ρ1 |β + (1 − η) β| ρ2 |β . (8.153)
Since both η and 1 − η are positive, this implies that [recall inequality
(8.13)]
0 = β| ρ1 |β = β| ρ2 |β . (8.154)
Let ρs,n,m be the matrix elements of the operator ρs , where s ∈ {1, 2}, in
a given orthonormal basis, and assume that the first vector of the basis
is taken to be the vector |α . In general
Tr (ρs ) = ρs,n,n , (8.155)
n
2
Tr ρ2s = ρs,n,n . (8.156)
n,m
The requirement Tr (ρs ) = 1 together with Eqs. (8.154) and (8.155) imply
that
ρs,1,1 = α| ρs |α = 1 . (8.157)
N |n = n |n , (8.158)
thus
N = n| N |n = n , (8.159)
and
) 2*
N = n| N 2 |n = n2 , (8.160)
therefore
∆N = 0 . (8.161)
a |α = α |α , (8.162)
thus
2
N = α| N |α = α| a† a |α = |α| , (8.163)
and
) 2* , -
N = α| a† aa† a |α = α| a† a, a† + a† a a |α = |α|2 + |α|4 ,
=1
(8.164)
therefore
∆N = N . (8.165)
O = Tr (ρO) , (8.166)
where O is an operator,
1
ρ = e−Hβ , (8.167)
Z
Z = Tr e−Hβ , (8.168)
and β = 1/kB T and H is the Hamiltonian. For the present case
1
H= ω N+ , (8.169)
2
thus
N = Tr (ρN )
∞
=
n| e−Hβ N |n
n=0
= = ∞
n| e−Hβ |n
n=0
∞
=
ne−n ωβ
n=0
= ∞
=
e−n ωβ
n=0
∞
=
∂ log e−n ωβ
1 n=0
=−
ω ∂β
e−β ω
= ,
1 − e−β ω
(8.170)
and ) *
N 2 = Tr ρN 2
∞
=
n| e−Hβ N 2 |n
n=0
= = ∞
n| e−Hβ |n
n=0
∞
=
n2 e−n ωβ
n=0
= =∞
e−n ωβ
n=0
=∞
1 2 ∂2
ω ∂β 2
e−n ωβ
n=0
= ∞
=
e−n ωβ
n=0
−β ω
e + 1 e−β ω
= 2 ,
(1 − e−β ω )
(8.171)
and therefore
) * e−β ω
(∆N)2 = N 2 − N 2
= = N ( N + 1) . (8.172)
(1 − e−β ω )2
2
x2 = a2 + a† + 2a† a + 1 , (8.175)
2mω
one finds
) 2*
x = Tr x2 ρ
∞
1
= n| x2 e−Hβ |n
Z n=0
∞
1
e− ω(n+ 2 )β n| x2 |n
1
=
Z n=0
∞
1 1
e− ω(n+ 2 )β
1
= n+
mω Z n=0
2
∞
1 1 d
e− ω(n+ 2 )β .
1
= −
mω Z ω dβ n=0
(8.176)
However
∞
e− ω(n+ 2 )β = Z ,
1
(8.177)
n=0
thus
) 2* 1 d
x = log Z −1
mω 2 dβ
d ωβ
1 dβ sinh 2
=
mω 2 sinh ωβ
2
1 ω ωβ
= 2
coth .
mω 2 2
(8.178)
Note that at high temperatures ωβ ≪ 1
) 2* kB T
x ≃ , (8.179)
mω 2
as is required by the equipartition theorem of classical statistical mechan-
ics.
= n=0
=∞
n| e−β ω(N+ 2 ) |n
1
n=0
=∞
e−nβ ω
|n n|
n=0
= ∞
=
e−nβ ω
n=0
∞
= 1 − e−β ω
e−nβ ω
|n n| ,
n=0
(8.180)
where β = 1/kB T . Thus, N is given by
N = Tr (ρN )
∞
= 1 − e−β ω
ne−nβ ω
n=0
∞
∂
= − ω 1 − e−β ω
e−nβ ω
∂β n=0
∂ 1
= − ω 1 − e−β ω
∂β 1 − e−β ω
−β ω
e
= .
1 − e−β ω
(8.181)
Moreover, the following holds
1
N +1 = , (8.182)
1 − e−β ω
N
= e−β ω
, (8.183)
N +1
thus, ρ can be rewritten as
n| ρ |m = d2 αP (α) n |α α |m . (8.185)
2πδnm
∞
2δ nm
dre−(1+ )r2 r2n+1 ,
1
= N
N n!
0
(8.188)
and the transformation of the integration variable
1
t= 1+ r2 , (8.189)
N
1
dt = 1+ 2rdr , (8.190)
N
lead to
ρ= d2 α |α α| P (α) , (8.192)
and where
e−β ω
N = (8.194)
1 − e−β ω
1+2 N
2
;
√x2′
1 1 mω
= √ e−2 1+2 N
π mω (1 + 2 N )
2
1 − x′
= √ e ξ
,
ξ π
where
0
ξ= (1 + 2 N ) ,
mω
and where
e−β ω β ω
1+2 N = 1+2 = coth . (8.198)
1 − e−β ω 2
12. Recall that the LC circuit is a harmonic oscillator.
a) In terms of the annihilation and creation operators
0
Lω ip
a= q+ , (8.199)
2 Lω
0
† Lω ip
a = q− , (8.200)
2 Lω
one has 0
q= a + a† , (8.201)
2Lω
1
H = ω a† a + . (8.202)
2
thus
0 ∞
1
q = Tr (qρ) = n| a + a† e−βH |n = 0 . (8.206)
Z 2Lω n=0
b) Similarly
) 2*
q = Tr q 2 ρ
∞
1 2 −βH
= n| a + a† e |n
2Lω Z n=0
∞
1 1 1
e−β ω(n+ 2 )
1
= ω n+
Lω2 Z n=0
2
1 1 dZ
=− 2
Lω Z dβ
C ω ω
= coth .
2 2kB T
(8.207)
13. In general, ρ0 can be expressed as
('
ρ0 = wi α(i) α(i) , (8.208)
i
= ) *
where 0 ≤ wi ≤ 1, i wi = 1, and where α(i) α(i) = 1. Assume first
*
that the system is initially in the state α(i) . The probability for this to
be the case is wi . In general, the possible results of a measurement of the
observable A are the eigenvalues {an }. The probability pn to *measure the
eigenvalue an given that the system is initially in state α(i) is given by
' (
pn = α(i) Pn α(i) . (8.209)
( *
(i) Pn α(i)
α → )
. *. (8.210)
α(i) Pn α(i)
*
Thus, given that the system is initially in state α(i) the final density
operator is given by
' ( * ) (i)
(i) (i) (i) Pn α(i) α Pn
ρ1 = α Pn α .) * .) *
n α(i) Pn α(i) α(i) Pn α(i)
('
= Pn α(i) α(i) Pn .
n
(8.211)
Averaging over all possible initial states thus yields
('
(i)
ρ1 = wi ρ1 = Pn wi α(i) α(i) Pn = Pn ρ0 Pn . (8.212)
i n i n
14. Since [V (t) , V (t′ )] = 0 the time evolution operator from initial time t0
to time t is given by
t
i
u (t, t0 ) = exp − dt′ V (t′ )
t0
ipi A
= exp x ,
(8.213)
where
t
pi = dt′ f (t′ ) . (8.214)
t0
While the initial state of the entire system at time t0 is given by |Ψ (t0 ) =
|ψi ⊗ |α , the final state at time t is given by
ipi an
Jn = exp x . (8.216)
where
p0 = . (8.221)
x0
In terms of φ (p′ ) the state |Ψ (t) thus can be expressed as
where
∞
pi x0
η= = dt′ f (t′ ) . (8.224)
p0 t0
A = dA′ g (A′ ) A′
−∞
∞
η ′′ 2
= |cn′ | 2
dA′′ e−(ηA ) (A′′ + an′ )
n′
π1/2
−∞
2
= |cn′ | an′ .
n′
(8.225)
B̄ = Tr (Bρf )
thus
an′ −an′′ 2
−η2
ρR = cn′ c∗n′′ e 2
|an′ an′′ | . (8.234)
n′ ,n′′
L = L+ + L− , (8.237)
where
mẋ′2 mω2
L+ = − (1 + λ) x′2 , (8.238)
2 2
and
mẏ ′2 mω2
L− = − (1 − λ) y ′2 . (8.239)
2 2
Thus, the system is composed of√two decoupled harmonic
√ oscillators with
angular resonance frequencies ω 1 + λ (for x′ ) and ω 1 − λ (for y ′ ). In
thermal equilibrium according to Eq. (8.178) one has
√
) ′2 * ω 1 + λβ
x = √ coth , (8.240)
2mω 1 + λ 2
√
) ′2 * ω 1 − λβ
y = √ coth , (8.241)
2mω 1 − λ 2
where β = 1/kB T . Moreover, due to symmetry, the following holds
x′ = y ′ = 0 , (8.242)
′ ′
xy =0. (8.243)
With the help of the inverse transformation, which is given by
x′ + y ′
x= √ , (8.244)
2
x′ − y ′
y= √ , (8.245)
2
one thus finds
x =0, (8.246)
and
) 2 * 1 ) ′2 *
x = x + y′2
2
√ √
coth ω 21+λβ coth ω 1−λβ
2
= √ + √ .
4mω 1+λ 1−λ
(8.247)
16. Using Eq. (5.155), which is given by
p(H) (0)
x(H) (t) = x(H) (0) cos (ωt) + sin (ωt) , (8.248)
mω
one finds that
' ( sin (ωt) ) *
G (t) = cos (ωt) x2(H) (0) + p(H) (0) x(H) (0) . (8.249)
mω
where
e−β ω
N = Tr (ρN ) = , (8.257)
1 − e−β ω
one has
∞ 5 6
1 p′ x′′ x′′ x′′
W (x′ , p′ ) = exp i x′ − |α α x′ + dx′′
2π −∞ 2 2
mω 1/2 ∞
= π
dx′′
2π −∞
! ′′
"2 ! ′′
"2
x′ − x − x α x′ + x − x α p′ − p α
− 2
2∆xα − 2
2∆xα +i x′′
×e
mω 1/2 ∞
= π
dx′′
2π −∞
′′ 2
−
(x′ − (
x α )2 + x
2 ) +i
p′ − p α
x′′
2(∆xα )2
×e ,
(8.266)
thus
2 2
x′ − x α p′ − p α
′1 − 12
′ ∆xα − 12 ∆pα
W (x , p ) = e , (8.267)
π
where [see Eq. (5.49)]
. 0
2 mω
∆pα = α| (∆p) |α = = . (8.268)
2 2∆xα
*
18. At time t > 0 the system is in a coherent state α = α0 e−iωt , where [see
Eqs. (5.53) and (5.243)]
0
mω
α0 = ∆x . (8.269)
2
Thus, with the help of Eq. (8.267) one finds that the Wigner function of
the system at time t is given by
ρ= d2 α |α α| P (α) , (8.275)
and where
e−β ω
N = (8.277)
1 − e−β ω
where
∞ 5 6
1 p′ x′′ x′′ x′′
Wα (x′ , p′ ) = exp i x′ − |α α x′ + dx′′ ,
2π −∞ 2 2
(8.279)
where
0
x0 = . (8.291)
mω
thus
′′ 2 ′′ 2
′′
−
(x′ − x2 ) ′′
−
(x′ + x2 )
∞ x′ − x2 2x2 x′ + x2 2x2
1 p′ x′′ x0 e 0
x0 e 0
W (x′ , p′ ) = ei dx′′ ,
π −∞ π1/2 x0
(8.292)
or
− x′
2 ! "
∞ 2
e x0
1 x′ X2 ip′ X2
′
W (x , p ) =′
− e p0 X− 4
dX , (8.293)
π π1/2 −∞ x0 4
where X = x′′ /x0 and where p0 = /x0 . The integration, which is per-
formed with the help of Eq. (5.139), yields
2 2
2 − x′
2
− p′ 2
x′ p′ 1
W (x′ , p′ ) = e x0 p0
+ − . (8.294)
π x0 p0 2
Note that near the origin the Wigner function W (x′ , p′ ) becomes nega-
tive.
21. The relation (8.67) is proven by
∞
−1
dp′ W (x′ , p′ )
−∞
5 ∞ 6 ∞
′ x′′ ′ x′′ 1 p′ x′′
= x − ρ x + dx′′ dp′ ei
−∞ 2 2 2π −∞
δ(x′′ )
′ ′
= x | ρ |x .
(8.295)
With the help of the identities (3.45) and (3.52) W (x′ , p′ ) can be ex-
pressed as
W (x′ , p′ )
′′ ′′
∞ ∞ ∞ (
p′ x′′ +p′′ x′ − x )−p′′′ (x′ + x2 )
1 ′′ ′′ ′′′ i 2
= dx dp dp e p′′ | ρ |p′′′
(2π)2 −∞ −∞ −∞
p′′ p′′′
∞ ∞ x′ (p′′ −p′′′ ) ∞ x′′ p′ − −
1 ′′ ′′′ i ′′ ′′′ 1 ′′ i
2 2
= dp dp e p | ρ |p dx e .
2π −∞ −∞ 2π −∞
′′ ′′′
δ p′ − p2 − p2
(8.296)
The above result easily leads to (8.68)
δ(p′′ −p′′′ )
′ ′
= p | ρ |p .
(8.297)
22. For a pure state ρ = |α α| one finds that the Wigner function is given
by [see Eq. (8.66)]
∞
1 p′ x′′ x′′ x′′
′
W (x , p ) = ′
dx′′ ei ψα x′ − ψ∗α x′ + , (8.298)
2π 2 2
−∞
p2
H= + V (x) , (8.301)
2m
where p is the canonical conjugate operator to the position operator x.
With the help of Eq. (8.35), which reads
dρ 1
= − [ρ, H] , (8.302)
dt i
one finds that
represents the contribution of the kinetic energy, and where the term
∞ 5 6
1 ip′ x′′ x′′ x′′
Sp = − dx′′ e x′ − [ρ, V (x)] x′ + (8.305)
2πi 2 2
−∞
1= |φn φn | , (8.306)
n
The term V (x′ + x′′ /2)−V (x′ − x′′ /2), which represents an odd function
of x′′ , can be Taylor expanded as
2l+1
x′′ x′′ ∞
= 1 ∂ 2l+1 V x′′
V x′ + −V x′ − =2 2l+1
.
2 2 l=0 (2l + 1)! ∂ (x′ ) 2
(8.320)
∞ 5 6 2l+1
1 2l+1 ip′ x′′ x′′ x′′ ∂ 2l+1 W
dx′′ (x′′ ) e x′ − ρ x′ + = .
2π 2 2 i ∂ (p′ )2l+1
−∞
(8.321)
Employing the above results to evaluate Eq. (8.305) and separating the
first term from all higher order terms yield
2l
∂V ∂W =∞
2i ∂ 2l+1 V ∂ 2l+1 W
Sp = + 2l+1
. (8.322)
∂x′ ∂p′ l=1 (2l + 1)! ∂ (x′ ) ∂ (p′ )2l+1
Note that when ∂ 2l+1 V/∂ (x′ )2l+1 = 0 for l ≥ 1 the above result coin-
cides with the classical equation of motion dW/dt = {H, W }. Thus one
concludes that the quantum time evolution of W of a harmonic oscillator
is identical to the classical one.
24. With the help of Eq. (2.180) one finds that
iξη
exp (−iξX − iηP ) = e− 2 e−iηP e−iξX . (8.325)
δ(X ′′ −X ′ )
∞ 5 6
1 ′′ ′ X ′′ ′ X ′′ ′′ ′
= dX X − ρ X + eiX P .
2π 2 2
−∞
(8.327)
′ ′
25. For the case X = P = 0 the operator Υ is given by
∞ ∞
1
Υ (0, 0) = dξdη e−iξX−iηP , (8.328)
(2π)2
−∞ −∞
δ( η2 +X ′ )
′
= π−1 e2iX P
|X ′ ,
(8.330)
or [see Eq. (3.19)]
Υ (0, 0) |X ′ = π −1 |−X ′ , (8.331)
which implies that Υ (0, 0) is related to the parity operator P (8.78) by
Υ (0, 0) = π−1 P . (8.332)
By employing the relations
a + a† a − a†
X= √ , P = √ , (8.333)
2 2i
together with Eq. (8.328) one finds that
∞ ∞
1 −iξ a+a
√
†
−iη a−a
√
†
P= dξdη e 2 2i
4π
−∞ −∞
∞ ∞
1 − iξ+η
√ a− iξ−η
√ a†
= dξdη e 2 2 .
4π
−∞ −∞
(8.334)
The above result together with Eq. (5.38) yield
X ′ + iP ′ X ′ + iP ′
π−1 D† − √ PD − √
2 2
∞ ∞
′ ′ ′ −iP ′
1 − iξ+η
√ a− X √
+iP
− iξ−η
√ a† − X √
= 2 dξdη e 2 2 2 2
(2π)
−∞ −∞
∞ ∞
′ ′ ′ −iP ′
1 − iξ+η
√ X+iP
√ −X √
+iP
− iξ−η
√ X−iP
√ −X √
= 2 dξdη e 2 2 2 2 2 2
(2π)
−∞ −∞
∞ ∞
1 ′
−X )+iη (P ′ −P )
= dξdη eiξ(X ,
(2π)2
−∞ −∞
(8.335)
X ′ + iP ′ X ′ + iP ′
π−1 D† − √ PD − √ = Υ (X ′ , P ′ ) . (8.336)
2 2
26. The normalized homodyne observable Xφ can be expressed in terms of
the dimensionless position and momentum operators X and P , which are
given by
a + a† a − a†
X= √ , P = √ , (8.337)
2 i 2
and which satisfy [X, P ] = i [see Eq. (5.13)], as
X − iP iφ X + iP −iφ
Xφ = e + e
2 2
= X cos φ + P sin φ .
(8.338)
Xφ cos φ sin φ X
= , (8.339)
Pφ − sin φ cos φ P
X cos φ − sin φ Xφ
= . (8.340)
P sin φ cos φ Pφ
∞
) −iζXφ * ′
e = dXφ′ w Xφ′ e−iζXφ = w̃φ (ζ) . (8.342)
−∞
The comparison between Eq. (8.341) and Eq. (8.343) yields the fol-
lowing relation (recall that Xφ = X cos φ + P sin φ)
w Xφ′ = dXφ′′ dPφ′ W Xφ′′ cos φ − Pφ′ sin φ, Xφ′′ sin φ + Pφ′ cos φ
−∞−∞
∞
1 ′′ ′
× dζ e−iζ (Xφ −Xφ )
2π
−∞
δ (Xφ φ)
′′ −X ′
∞ π
1 ′
cos φ+P ′ sin φ)
′ ′
W (X , P ) = 2 dζ dφ |ζ| w̃φ (ζ) eiζ (X ,
(2π)
−∞ 0
(8.350)
∞ π ∞
1 ′
cos φ+P ′ sin φ−Xφ
′
′
W (X , P ) = ′
dζ dφ dXφ′ |ζ| w Xφ′ eiζ (X ) .
(2π)2
−∞ 0 −∞
(8.351)
ρ= d2 α |α α| P (α) , (8.352)
and where
e−β ω
N = (8.354)
1 − e−β ω
one has
; X ′ +X ′′ 2
2 2 N +1
mω 1/2 1 2 X ′ −X ′′
′′ ′ − −
x | ρ |x = e 2(2 N +1) 2 2
.
π 2 N +1
(8.366)
where
0
x0 = . (8.371)
mω
28. The following holds [see Eqs. (8.8), (8.42)]
a + a†
X0 = √ , (8.376)
2
is related to the position operator x by [see Eq. (5.11)]
x = x0 X0 , (8.377)
where
0
x0 = . (8.378)
mω
The last result implies that
† †
e−iζXφ = eiφa a e−iζX0 e−iφa a
, (8.379)
and thus [see Eq. (8.372) and recall that Tr (AB) = Tr (BA)]
) −iζXφ * ) −iζX0 * ' −iζ √ mω x (
e = e = e . (8.380)
) −iζX *
In other words, e φ
is found to be independent on φ. The expecta-
' √ mω (
−iζ x
tion value e can be calculated by employing the expression
for the matrix elements x′′ | ρ |x′ of the density operator ρ given by Eq.
(8.370)
∞
) −iζXφ * √ mω
e = dx′ x′ | ρe−iζ x
|x′
−∞
2
∞ x′
√ mω − tanh( β 2ω )
′ −iζ x′ e x0
= dx e 0 ,
β ω
−∞ x0 π coth 2
(8.381)
and where β = 1/kB T . Using the identity (5.139) one finds that
) −iζXφ * ζ 2 coth( β ω )
2
−
e =e 4 . (8.382)
1 β ω 1 ) 2*
coth = N + = X0 , (8.383)
2 2 2
e−β ω
N = , (8.384)
1 − e−β ω
e φ
= e− 2 . (8.385)
) −iζX *
The factor e φ
is the characteristic function of the probability dis-
tribution function Pr Xφ′ of Xφ′ , which is denoted below as w Xφ′ ,
and thus it is related to the Fourier transform w̃φ (ζ) of the probability
distribution w Xφ′ by [see Eq. (8.342)]
∞
) −iζX * ′
e φ
= dXφ′ w Xφ′ e−iζXφ = w̃φ (ζ) . (8.386)
−∞
With the help of Eqs. (8.342), (8.344) and (8.385) one finds that the
Fourier transform W̃ (ξ, η) of the Wigner function W (X ′ , P ′ ) satisfy the
following relation for any real φ
ζ 2 X02
and thus
(ξ2 +η2 ) 2
X0
W̃ (ξ, η) = e− 2 . (8.388)
The inverse Fourier transformation [see Eqs. (5.139) and (8.327)] yields
∞ ξ2 X02 ∞ η 2 X02
′ ′ 1 − iξX ′ 1 − ′
W (X , P ) = dξ e 2 e dη e 2 eiηP
2π −∞ 2π −∞
′2 ′2
1 − X +P2
= e 2 X0
2π X02
1 X ′2 +P ′2
= e− 2 N +1 .
π (2 N + 1)
(8.389)
It is easy to see that the above result for the Wigner function W (X ′ , P ′ )
for the dimensionless variables X ′ and P ′ is consistent with Eq. (8.288)
for the Wigner function W (x′ , p′ ) for the displacement x′ and momentum
p′ variables.
29. By using Eqs. (8.70) and employing cylindrical coordinates
thus since w Xφ′ is φ independent [see Eqs. (8.342) and (8.344)] one
has [note that contrary to Eq. (8.351), integration over ζ below is taken
to be over positive values only]
∞ π
′ ′1 ′
W (X , P ) = dζ ζ w̃ (ζ) dφ eiζR cos(φ−θ)
, (8.392)
(2π)2
0 −π
where
∞
′
w̃ (ζ) = dXφ′ w Xφ′ e−iζXφ . (8.393)
−∞
thus
∞
′ 1
′
W (X , P ) = dζ ζ w̃ (ζ) J0 ζ X ′2 + P ′2 , (8.396)
2π
0
or
/∞
dz w̃ √ z zJ0 (z)
X ′2 +P ′2
′ ′ 0
W (X , P ) = . (8.397)
2π (X ′2 + P ′2 )
ζ 2 X02
w̃ (ζ) = e− 2 , (8.398)
where u (t) is the time evolution operator for the system. The Wigner
function W (X ′ , P ′ ; t) can be expressed according to Eq. (8.351) as
∞ π ∞
1 ′
cos φ+P ′ sin φ−Xφ
′
′ ′
W (X , P ; t) = 2 dζ dφ dXφ′ |ζ| w Xφ′ eiζ (X ) ,
(2π)
−∞ 0 −∞
(8.402)
a† eiφ + ae−iφ
Xφ = √ . (8.403)
2
In terms of the density operator ρ (t) one has [recall that Tr (AB) =
Tr (BA)]
*)
w Xφ′ = Tr Xφ′ Xφ′ ρ (t)
*)
= Tr u† (t) Xφ′ Xφ′ u (t) ρ0 ,
(8.404)
( (
where Xφ′ is an eigenvector of Xφ with eigenvalue Xφ′ , i.e. Xφ Xφ′ =
(
Xφ′ Xφ′ . With the help of Eqs. (2.178) and (4.9) one finds that
† iφ
† + ae−iφ −iω(a† a+ 12 )t
a+ 12 )t a e
u† (t) Xφ u (t) = eiω(a √ e
2
† iφ
† a e + ae−iφ −iωta† a
= eiωta a √ e
2
a† ei(φ+ωt) + ae−i(φ+ωt)
= √
2
= Xφ+ωt ,
(8.405)
thus the following holds
* *
Xφ+ωt u† (t) Xφ′ = Xφ′ u† (t) Xφ′ , (8.406)
(
i.e. u† (t) Xφ′ is an eigenvector of Xφ+ωt with eigenvalue Xφ′ . This eigen-
(
′
vector is labeled below as Xφ+ωt . Using these results and the trigono-
metric identities
cos (x + y) = cos x cos y − sin x sin y , (8.407)
sin (x + y) = cos x sin y + sin x cos y , (8.408)
the Wigner function W (X ′ , P ′ ; t) becomes
∞ π ∞
′ ′ 1
W (X , P ; t) = dζ dφ dXφ′ |ζ|
(2π)2
−∞ 0 −∞
*) ′ ′ ′ ′
× Tr ′
Xφ+ωt Xφ+ωt ρ0 eiζ (X cos φ+P sin φ−Xφ )
∞ π ∞
1 ′
= dζ dφ dXφ′ |ζ|
(2π)2
−∞ 0 −∞
*) ′ ′ ′ ′ ′ ′
× Tr Xφ′ ′ Xφ′ ρ0 eiζ (X cos(φ −ωt)+P sin(φ −ωt)−Xφ ) ,
(8.409)
thus
i.e. the time evolution of the Wigner function represents rigid rotation in
phase space at angular velocity ω.
31. The Wigner function W (X ′ , P ′ ) can be expressed as [see Eq. (8.70)]
∞ ∞
′ ′1 ′
+iηP ′
W (X , P ) = dξdη W̃ (ξ, η) eiξX , (8.411)
(2π)2
−∞ −∞
where
and where
ξ + iη
α= √ , (8.413)
2i
thus for the displaced system one has [see Eq. (5.41) and recall that
Tr (AB) = Tr (BA)]
∞ ∞
′ ′ 1 ′
+iηP ′
Wα (X , P ) = 2 dξdη Tr D (α) D (α′ ) ρD† (α′ ) eiξX
(2π)
−∞ −∞
∞ ∞
1 ′∗
−α∗ α′ ′
+iηP ′
= dξdη eαα Tr (D (α) ρ) eiξX ,
(2π)2
−∞ −∞
(8.414)
thus using
′ +α′∗ α′√
−α′∗
′∗
−α∗ α′ iξX ′ +iηP ′ iξ X ′ − α √ +iη P ′ −
eαα e =e 2 2i (8.415)
where
α′ + α′∗
Xα′ ′ = √ , (8.417)
2
α′ − α′∗
Pα′ ′ = √ . (8.418)
i 2
32. In general, the convolution theorem states that
∞ ∞
1 ′′ ′′ ′′ iX ′′ P ′
dX f (X ) g (X ) e = dP ′′ F (P ′′ ) G (P ′ − P ′′ ) ,
2π
−∞ −∞
(8.419)
where F (P ′ ) (G (P ′ )) is the Fourier transform of f (X ′′ ) (g (X ′′ )), i.e.
∞
1′ ′′
P′
F (P ) = dX ′′ f (X ′′ ) eiX , (8.420)
2π
−∞
∞
1 ′′
P′
G (P ′ ) = dX ′′ g (X ′′ ) eiX , (8.421)
2π
−∞
′ ′ ′′ e−( η ′ ′′
)
WR (X , P ) = dP W (X , P ) √ . (8.423)
πη
−∞
33. The normalization constant C is found with the help of Eq. (5.233) [see
Eq. (8.87)]
1 = ψ |ψ
2
= 2 |C|2 1 + e−2|α| cos θ 0 ,
(8.426)
where
With the help of Eqs. (5.35) and (5.41) one finds that
1/2
|α0 + α = ζ 0 D (α0 ) |α , (8.428)
where
and thus
and where
ρ+,+ = |α α| , (8.432)
ρ+,− = ζ 0 |α −α| , (8.433)
ρ−,+ = ζ ∗0 |−α α| , (8.434)
ρ−,− = |−α −α| . (8.435)
The Wigner function W0 of the density operator ρ0 can be expressed in
terms of to the Wigner function W of the density operator ρ [see Eqs.
(8.416) and (8.430)]
where
α0 + α∗0
Xr′ = X ′ − √ , (8.437)
2
α0 − α∗0
Pr′ = P ′ − √ . (8.438)
i 2
The density operator W (Xr′ , Pr′ ) is expressed as [see Eq. (8.431)]
where Wσ1 ,σ2 is the Wigner function of ρσ1 ,σ2 , and where σ 1 , σ2 ∈ {+, −}.
With the help of Eq. (8.74) one finds that [see Eqs. (2.133) and (5.41)]
W+,+ (Xr′ , Pr′ ) = Tr π−1 P |−Zr′ + α −Zr′ + α| , (8.440)
W+,− (Xr′ , Pr′ ) = ζ 0 ζ Tr π −1
P |−Zr′ +α −Zr′ − α| , (8.441)
W−,+ (Xr′ , Pr′ ) = ζ ∗0 ζ ∗ Tr π −1
P |−Zr′ −α −Zr′ + α| , (8.442)
W−,− (Xr′ , Pr′ ) = Tr π −1
P |−Zr′ −α −Zr′ − α| , (8.443)
where P is the parity operator, Zr′ is given by
Xr′ + iPr′
Zr′ = √ , (8.444)
2
and ζ is given by
where
α1 = α′1 + iα′′1 , (8.448)
α2 = α′2 + iα′′2 , (8.449)
and thus [see Eq. (5.139)]
! "
|α1 |2 + |α2 |2 + 2α1 α∗2
Tr π−1 P |α1 α2 | = π −1 exp − . (8.450)
2
With the help of the above results [see Eqs. (8.426), (8.436), (8.439),
(8.444) and (8.450)] one obtains
′ 2
e−2|Z− | + 2 Re ζ 0 ζ 2 e−2|Zr | + e−2|Z+ |
2 2
W = 2 , (8.451)
2π 1 + e−2|α| cos θ0
where
Z± = Zr′ ± α. (8.452)
35. In general, for any smooth function f (ρ) of ρ the following holds
using Eq. (8.37) and the fact that u† (t, t0 ) u (t, t0 ) = 1, i.e. the unitarity
of the time evolution operator. By using this result for the function ρ log ρ
together with the general identity Tr (XY ) = Tr (Y X) [see Eq. (2.133)]
one easily finds that σ is time independent. This somewhat surprising
result can be attributed to the fact that the unitary time evolution that is
governed by the Schrödinger equation is symmetric under time reversal.
In the language of statistical mechanics it corresponds to a reversible
process, for which entropy is preserved.
36. Using the definition of the Pauli matrices (6.136) one finds that
1 1 + kz kx − iky
ρ= , (8.457)
2 kx + iky 1 − kz
and
1 1 + 2kz + k2 2 (kx − iky )
ρ2 = ,
4 2 (kx + iky ) 1 − 2kz + k2
(σ · a) (σ · b) = a · b + iσ · (a × b) , (8.458)
and the fact that all three Pauli matrices have a vanishing trace, one
finds that [compare with Eq. (8.151)]
1 1
Tr (û · σρ) = Tr (û · σ) + Tr ((û · σ) (k · σ))
2 2
1
= Tr ((û · σ) (k · σ))
2
1 i
= Tr (û · k) + Tr (σ · (û × k))
2 2
1
= Tr (û · k)
2
= û · k .
(8.459)
37. Using the definition of the Pauli matrices (6.136) one finds that
1 1 + kz kx − iky
ρ= . (8.460)
2 kx + iky 1 − kz
a) Let λ± be the two eigenvalues of ρ. The following holds
Tr (ρ) = λ+ + λ− = 1 , (8.461)
and
Det (ρ) = λ+ λ− = 1 − k2 /4 , (8.462)
where k2 = kx2 + ky2 + kz2 . Thus
1 ± |k|
λ± = , (8.463)
2
and therefore
σ = f (|k|) , (8.464)
where
1−x 1−x 1+x 1+x
f (x) = − log − log . (8.465)
2 2 2 2
0.7
0.6
0.5
0.4
y
0.3
0.2
0.1
38. First, consider a general functional g (ρ) of the density operator having
the form
df
dg = Tr dρ . (8.472)
dρ
In the above expression the term df /dρ is calculated by simply taking the
derivative of the function f (x) (where x is considered to be a number)
and substituting x = ρ. Alternatively, the change dg can be expressed in
terms of the infinitesimal change dρnm in the matrix elements ρnm of ρ.
To first order in the infinitesimal variables dρnm one has
∂g
dg = dρnm . (8.473)
n,m
∂ρnm
¯ ∂
∇ n,m
= , (8.475)
∂ρnm
and
dρ n,m
= dρnm . (8.476)
and
¯ l · dρ ,
dgl = ∇g (8.478)
where l = 0, 1, 2, ...L.
a) In general, the technique of Lagrange multipliers is very useful for
finding stationary points of a function, when constrains are applied.
A stationary point of σ occurs iff for every small change dρ, which is
¯ 0 , ∇g
orthogonal to all vectors ∇g ¯ 1 , ∇g
¯ 2 , ..., ∇g
¯ L (i.e. a change which
does not violate the constrains) one has
¯ · dρ .
0 = dσ = ∇σ (8.479)
¯ belongs to the
This condition is fulfilled only when the vector ∇σ
¯ ¯ ¯ ¯ L . In other
subspace spanned by the vectors ∇g0 , ∇g1 , ∇g2 , ..., ∇g
words, only when
¯ = ξ 0 ∇g
∇σ ¯ 0 + ξ 1 ∇g
¯ 1 + ξ 2 ∇g
¯ 2 + ... + ξ L ∇g
¯ L, (8.480)
Using Eqs. (8.472), (8.93, (8.95) and (8.96) one finds that
dσ = − Tr ((1 + log ρ) dρ) , (8.482)
dg0 = Tr (dρ) , (8.483)
dgl = Tr (Xl dρ) , (8.484)
thus
! L
"
0 = Tr 1 + log ρ + ξ 0 + ξ l Xl dρ . (8.485)
l=1
The requirement that the last identity holds for any dρ implies that
L
1 + log ρ + ξ 0 + ξ l Xl = 0 , (8.486)
l=1
thus
! L
"
−1−ξ0
ρ=e exp − ξ l Xl . (8.487)
l=1
where
! L
"
Z = Tr − ξ l Xl . (8.489)
l=1
As can be seen from the above expression for Z, the following holds
1 ∂ log Z
Xl = Tr Xl e−βH = − . (8.490)
Z ∂ξ l
Zc = Tr e−βH , (8.492)
and where β labels the Lagrange multiplier associated with the given
expectation value H . By solving Eq. (8.96), which for this case is
given by [see also Eq. (8.490)]
1 ∂ log Zc
H = Tr He−βH = − . (8.493)
Zc ∂β
the Lagrange multiplier β can be determined. Note that the tem-
perature T is defined by the relation β = 1/kB T , where kB is the
Boltzmann’s constant.
and
and therefore
∗ ∗
(ρ1 )n1 ,m1 = ρ(n1 ,n2 ),(m1 ,n2 )
n2
ρ1 = wn(1)
1
|n1 1 1 n1 | , (8.512)
n1
and
ρ2 = wn(2)
1
|n2 2 2 n2 | , (8.513)
n2
(1) (2)
where the eigenvalues satisfy 0 ≤ wn1 ≤ 1 and 0 ≤ wn1 ≤ 1. Similarly, ρ
can be diagonalized as
ρ= wk |k k| , (8.514)
k
and
As can be seen from Eqs. (8.507), (8.508) and (8.512), the following holds
wn(1)
1
= (ρ1 )n1 ,n1
= ρ(n1 ,n2 ),(n1 ,n2 )
n2
= n1 , n2 | ρ |n1 , n2
n2
= n1 , n2 |k wk k |n1 , n2 ,
n2 k
(8.518)
thus
wn(1)
1
= wn1 ,n2 , (8.519)
n2
and similarly
wn(2)
2
= wn1 ,n2 , (8.520)
n1
where
Note that
! "
wn1 ,n2 = wk k| |n1 , n2 n1 , n2 | |k
n1 ,n2 k n1 ,n2
= wk k |k ,
k
(8.522)
thus the normalization condition k |k = 1 together with the require-
ment that
Tr ρ = wk = 1 , (8.523)
k
imply that
i.e.
Tr1 ρ1 = wn(1)
1
=1, (8.525)
n1
and
Tr2 ρ2 = wn(2)
2
=1. (8.526)
n2
(1) (2)
Consider the quantity y wn1 ,n2 /wn1 wn2 , where the function y (x) is
given by
y (x) ≥ 0 (8.530)
(1) (2)
for x ≥ 0. For x = wn1 ,n2 /wn1 wn2 the inequality (8.530) implies that
wn1 ,n2 wn1 ,n2 wn1 ,n2
(1) (2)
log (1) (2)
− (1) (2)
+1≥0. (8.531)
wn1 wn2 wn1 wn2 wn1 wn2
(1) (2)
Multiplying by wn1 wn2 and summing over n1 and n2 yields
wn1 ,n2
wn1 ,n2 log (1) (2)
− wn1 ,n2 + wn(1)
1
wn(2)
2
≥ 0 , (8.532)
n1 ,n2 wn1 wn2 n1 ,n2 n1 n2
thus with the help of Eqs. (8.524), (8.525) and (8.526) one finds that
wn1 ,n2
wn1 ,n2 log (1) (2)
≥0, (8.533)
n1 ,n2 wn1 wn2
2 1
= | n1 , n2 |k | wk log
n1 ,n2
wn1 ,n2
k
! "
2 wk
= | n1 , n2 |k | wk log
n1 ,n2
wn1 ,n2
k
2
− | n1 , n2 |k | wk log wk
n1 ,n2 k
! "
2 wk
= | n1 , n2 |k | wk log
n1 ,n2
wn1 ,n2
k
− wk log wk | n1 , n2 |k |2 ,
k n1 ,n2
=1
(8.535)
thus
− wn1 ,n2 log wn1 ,n2
n1 ,n2
! "
2 wk
= | n1 , n2 |k | wk log
n1 ,n2
wn1 ,n2
k
+σ .
(8.536)
Furthermore, according to inequality (8.530) the following holds
wk
| n1 , n2 |k |2 wk log
wn1 ,n2
k
wk wk
= | n1 , n2 |k |2 wn1 ,n2 log
wn1 ,n2 wn1 ,n2
k
wk
≥ | n1 , n2 |k |2 wn1 ,n2 −1 ,
wn1 ,n2
k
| n1 , n2 |k |2 wk − wn1 ,n2 | n1 , n2 |k |2
k k
=0.
(8.537)
These results together with inequality (8.534) yield
σ1 + σ2 ≥ σ . (8.538)
e−Hβ
ρ= = p+ |+ +| + p− |− −| , (8.539)
Tr (e−Hβ )
where the probabilities p+ and p− are given by
ωβ
e∓ 2
p± = ωβ ωβ , (8.540)
e− 2 +e 2
ω = |e| B/me c is the Larmor frequency [see Eq. (4.22)] and where β =
1/kB T , one finds that
σ = −p+ log p+ − p− log p−
ωβ ωβ ωβ
1 − tanh 2 1 − tanh 2 1 + tanh 2 1 + tanh ωβ2
=− log − log .
2 2 2 2
(8.541)
43. With the help of Eqs. (6.77) and (8.459) one finds that
H| Sz |H = √ , (8.542)
2 2
thus
1
pz+ − (1 − pz+ ) = √ , (8.543)
2
and therefore
π
pz+ = cos2 . (8.544)
8
44. The density operator is given by [see Eq. (8.180)]
∞
ρ = 1 − e−β ω
e−nβ ω
|n n| , (8.545)
n=0
n=0
∞
= − 1 − e−β ω
log 1 − e−β ω
e−nβ ω
n=0
∞
+β ω 1 − e−β ω
ne−nβ ω
.
n=0
(8.546)
and
For a given k the degeneracy index i can take the values 1, 2, · · · , gk . The
set of vectors {|k, 1 , |k, 2 , · · · , |k, gk } forms an orthonormal basis for the
eigensubspace Fk . The closure relation can be written as
gk
1= |k, i k, i| = Pk , (9.3)
k i=1 k
where
gk
Pk = |k, i k, i| (9.4)
i=1
H = H0 + λṼ , (9.6)
H |α = E |α . (9.7)
Chapter 9. Time Independent Perturbation Theory
Qn = 1 − Pn = Pk . (9.8)
k=n
λṼ |α = (E − H0 ) |α . (9.9)
Pn R = RPn = 0 . (9.12)
and similarly
RQn = R . (9.14)
Furthermore, by expressing H0 as
gk
H0 = Ek |k, i k, i| = En Pn + Ek P k , (9.15)
k i=1 k=n
(E − H0 ) R = Qn . (9.17)
The last two results suggest that the operator R can be considered as the
inverse of E − H0 in the subspace of eigenvalue zero of the projector Pn
(which is the subspace of eigenvalue unity of the projector Qn ). Multiplying
Eq. (9.9) from the left by R yields
λRṼ |α = R (E − H0 ) |α . (9.18)
λRṼ |α = Qn |α . (9.19)
Pn = |n n| . (9.23)
Pn |α = |n , (9.24)
namely
n |α = 1 . (9.25)
λṼ |α = (E − H0 ) |α , (9.26)
n| λṼ |α = n| (E − H0 ) |α , (9.27)
or
n| E |α = n| H0 |α + n| λṼ |α , (9.28)
thus
E = En + n| λṼ |α . (9.29)
|k, i k, i| Ṽ |k′ , i k′ , i| Ṽ |n
+λ2
(E − Ek ) (E − Ek′ )
k=n k′ =n
i i
+··· .
(9.30)
Substituting Eq. (9.30) into Eq. (9.29) yields
E = En + λ n| Ṽ |n
n| Ṽ |k, i k, i| Ṽ |n
+λ2
E − Ek
k=n
i
n| Ṽ |k, i k, i| Ṽ |k′ , i k′ , i| Ṽ |n
+λ3
(E − Ek ) (E − Ek′ )
k=n k′ =n
i i
+··· .
(9.31)
Note that the right hand side of Eq. (9.31) contains terms that depend on E.
To second order in λ one finds
2
| k, i| V |n |
E = En + n| V |n + + O λ3 . (9.32)
En − Ek
k=n
i
|k, i k, i| V |n
|α = |n + + O λ2 . (9.33)
En − Ek
k=n
i
The set of vectors {|n, 1 , |n, 2 , · · · , |n, gn } forms an orthonormal basis for
the eigensubspace Fn . Multiplying Eq. (9.9) from the left by Pn yields
Pn λṼ |α = Pn (E − H0 ) |α , (9.34)
Pn λṼ |α = (E − En ) Pn |α . (9.35)
Pn V Pn |α = (E − En ) Pn |α . (9.38)
9.2 Example
H = H0 + V , (9.39)
where
p2 mω2 x2
H0 = + . (9.40)
2m 2
and where
0
mω
V =λ ω x. (9.41)
H0 |n = En |n , (9.42)
Note that, as was shown in chapter 5 [see Eq. (5.164)], the eigenvectors
and eigenvalues of H can be found analytically for this particular case. For
the sake of comparison we first derive this exact solution. Writing H as
0
p2 mω 2 x2 mω
H= + +λ ω x
2m 2
! 0 " 2
p2 mω 2 1
= + x+λ − ωλ2 ,
2m 2 mω 2
(9.44)
one sees that H describes a one dimensional harmonic oscillator (as H0 also
does). The exact eigenenergies are given by
1
En (λ) = En (λ = 0) − ωλ2 , (9.45)
2
and the corresponding exact wavefunctions are
> 0
x′ |n (λ) = x′ + λ |n . (9.46)
mω
where J (∆x) is the translation operator, the exact solution (9.46) can be
rewritten as
! 0 "
x′ |n (λ) = x′ | J −λ |n , (9.48)
mω
or simply as
! 0 "
|n (λ) = J −λ |n . (9.49)
mω
one has
λ ω
m| V |n = √ m| a |n + m| a† |n
2
λ ω √ √
= √ nδ m,n−1 + n + 1δ m,n+1 .
2
(9.54)
Thus En (λ) can be expanded using Eq. (9.32) as
| k, i| V |n |2
En (λ) = En + n| V |n + + O λ3
En − Ek
k=n
=0 i
2
| n − 1| V |n | | n + 1| V |n |2
= En + + + O λ3
En − En−1 En − En+1
1 nλ2 (n + 1) λ2
= ω n+ + ω − ω + O λ3
2 2 2
1 λ2
= ω n+ − ω + O λ3 ,
2 2
(9.55)
in agreement (to second order) with the exact result (9.45), and |n (λ) can
be expanded using Eq. (9.30) as
|k, i k, i| V |n
|n (λ) = |n + + O λ2
En − Ek
k=n
i
|n − 1 n − 1| V |n |n + 1 n + 1| V |n
= |n + + + O λ2
En − En−1 En − En+1
√ √
|n − 1 λ√2ω n |n + 1 λ√2ω n + 1
= |n + − + O λ2
ω ω
λ λ
= |n + √ a |n − √ a† |n + O λ2 .
2 2
(9.56)
Note that with the help of the following identify
0
m ω
p=i −a + a† , (9.57)
2
the last result can be written as
! 0 "
ip
|n (λ) = 1 + λ |n + O λ2 . (9.58)
mω
one has
! 0 "
|n (λ) = J −λ |n + O λ2 , (9.60)
mω
9.3 Problems
1. The volume effect: The energy spectrum of the hydrogen atom was
calculated in chapter 8 by considering the proton to be a point particle.
Consider a model in which the proton is instead assumed to be a sphere
of radius ρ0 where ρ0 ≪ a0 (a0 is Bohr’s radius), and the charge of
the proton +e is assumed to be uniformly distributed in that sphere.
Show that the energy shift due to such perturbation to lowest order in
perturbation theory is given by
e2 2
∆En,l = ρ |Rn,l (0)|2 , (9.61)
10 0
where Rn,l (r) is the radial wave function.
V = Ar , (9.63)
1 1
V (x1 , x2 ) = mω 2 (x1 − a)2 + mω 2 (x2 + a)2 + λmω2 (x1 − x2 )2 ,
2 2
(9.65)
where λ is real. Find the energy of the ground state to lowest non-
vanishing order in λ.
7. A particle having mass m is confined in a potential well of width l, which
is given by
+
0 for 0 ≤ x ≤ l
V (x) = . (9.66)
+∞ elsewhere
l
W (x) = w0 δ x − , (9.67)
2
A perturbation given by
+
w0 if 0 ≤ x ≤ a2 and 0 ≤ y ≤ a
2
W (x, y) = , (9.69)
0 elsewhere
is added.
a) Calculate to lowest non-vanishing order in w0 the energy of the
ground state.
b) The same for the first excite state.
9. Consider a particle having mass m moving in a potential energy given by
mω 2 2
V (x, y) = x + y 2 + βmω 2 xy , (9.70)
2
where the angular frequency ω is a constant and where the dimensionless
real constant β is assumed to be small.
a) Calculate to first order in β the energy of the ground state.
b) Calculate to first order in β the energy of the first excited state.
10. Consider a harmonic oscillator having angular resonance frequency ω 0 .
A perturbation given by
ω1 † †
V = a a + aa (9.71)
2
is added, where a is the annihilation operator and ω1 is a positive con-
stant. Calculate the energies of the system to second order in ω1 .
11. The Hamiltonian of a spin S = 1 is given by
a) ground state.
b) first excited state.
17. Consider a rigid rotator whose Hamiltonian is given by
p21 p2
H= + 2 − αδ (x1 ) − αδ (x2 ) + λδ (x1 − x2 ) , (9.91)
2m 2m
where x1 and x2 are the coordinates of the first and second particle
respectively, p1 and p2 are the corresponding canonically conjugate mo-
mentums, α and λ are both real positive constants and δ () denotes the
delta function. Calculate to first order in λ the energy of the ground state
of the system.
19. In this problem the main results of time independent perturbation theory
are derived using an alternative approach. Consider a general square
matrix
W = D + ΩV , (9.92)
where Ω ∈ R, D is diagonal
n0 |m0 = δ nm , (9.94)
1= |n0 n0 | . (9.95)
n
W |n = λ |n (9.96)
to second order in Ω.
H = H0 + λṼ , (9.97)
H0 |k = Ek |k , (9.98)
k′ |k = δ kk′ . (9.99)
HR = eL He−L , (9.100)
9.4 Solutions
1. The radial force acting on the electron is found using Gauss’ theorem
e2
r2 r > ρ0
Fr (r) = e2 r 3 . (9.103)
r2 ρ r ≤ ρ0
0
3
Enx ,ny ,nz = ω + nx + ny + nz . (9.110)
2
With the help of Eqs. (5.11), (5.13), (5.28), (5.29) and (9.31) together
with the relation
2
r4 = x2 + y 2 + z 2
= x4 + y 4 + z 4 + 2x2 y2 + 2y2 z 2 + 2z 2 x2 ,
(9.111)
one finds that the energy of the ground state Egs is given by
3 ω
Egs = + g 0, 0, 0| r4 |0, 0, 0 + O g2
2
3 ω 2
= + 3g 0, 0, 0| x4 |0, 0, 0 + 6g 0, 0, 0| x2 |0, 0, 0 + O g2
2
2
3 ω
= + 15g + O g2 .
2 2mω
(9.112)
3. The wavefunctions for the unperturbed case are given by
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (9.113)
where for the states relevant to this problem
3/2
1
R10 (r) = 2 e−r/a0 , (9.114a)
a0
3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (9.114b)
2a0
3/2
1 r r
R21 (r) = √ e− 2a0 , (9.114c)
2a0 3a0
0
1
Y00 (θ, φ) = , (9.114d)
4π
0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (9.114e)
2 2π
0
0 1 3
Y1 (θ, φ) = cos θ , (9.114f)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (9.114g)
2 2π
and the corresponding eigenenergies are given by
EI
En(0) = − , (9.115)
n2
where
me e4
EI = . (9.116)
2 2
where
∞
3Aa0
100| V |100 = A dr r3 R10
2
(r) = . (9.119)
2
0
3/2
1
R10 (r) = 2 e−r/a0 , (9.124)
a0
3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (9.125)
2a0
3/2
1 r − r
R21 (r) = √ e 2a0 , (9.126)
2a0 3a0
0
1
Y00 (θ, φ) = , (9.127)
4π
0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (9.128)
2 2π
0
0 1 3
Y1 (θ, φ) = cos θ , (9.129)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (9.130)
2 2π
and the corresponding eigenenergies are given by
EI
En(0) = − , (9.131)
n2
where
me e4
EI = . (9.132)
2 2
The perturbation term V in the Hamiltonian is given by
V = eEz = eEr cos θ . (9.133)
The matrix elements of V are expressed as
∞ 1 2π
′ ∗
′ ′ ′ 3
n l m | V |nlm = eE dr r Rn′ l′ Rnl d (cos θ) dφ cos θ Ylm
′ Ylm .
0 −1 0
(9.134)
a) Disregarding spin this level is non degenerate. To 1st order
(0) (0)
E1 = E1 + 1, 0, 0| V |1, 0, 0 = E1 ,
since
1
d (cos θ) cos θ = 0 ,
−1
dφ e±iφ = 0 , (9.140)
0
one finds
0 0γ 0
0 00 0
M =
γ∗
, (9.141)
00 0
0 00 0
where
γ = 2, 0, 0| V |2, 1, 0
∞ 1 2π
3 ∗
= eE dr r R2,0 R2,1 d (cos θ) dφ cos θ Y00 Y10
0 −1 0
∞
4
eE r r r
= dr 2− e− a0
8 a0 a0
0
1 2π
1 2
× d (cos θ) cos θ dφ .
4π
−1 0
(9.142)
Using
1
2
d (cos θ) cos2 θ = , (9.143)
3
−1
and ∞
x4 e−x dx = 24 (9.144)
0
∞
x5 e−x dx = 120 (9.145)
0
one finds
γ = 2, 0, 0| V |2, 1, 0
∞
4
eE r r r
= dr 2− e− a0
24 a0 a0
0
∞
a0 eE
= dx (2 − x) x4 e−x
24
0
= −3a0 eE .
(9.146)
The eigenvalues of the matrix M are 0, 0, 3a0 eE and −3a0 eE. Thus
to 1st order the degeneracy is partially lifted with subspace of dimen-
(0)
sion 2 having energy E2 , and another 2 nondegenerate subspaces
(0)
having energies E2 ± 3a0 eE.
5. For E = 0 the normalized wavefunctions ψ(0) ′ ′ ′
nx ,ny ,nz (x , y , z ) are given by
ψ(0) ′ ′ ′
nx ,ny ,nz (x , y , z )
= x′ , y ′ , z ′ |nx , ny , , nz
3/2 l l l
2 nx π x′ + 2 ny π y ′ + 2 nz π z ′ + 2
= sin sin sin ,
l l l l
(9.147)
and the corresponding eigenenergies are
2 2
π n2x + n2y + n2z
En(0)
x ,ny ,nz
= , (9.148)
2ml2
where nx , ny , nz ∈ {1, 2, · · · }. The matrix elements of the perturbation
V = qEz are given by
) ′ ′ ′ * 2qE
nx , ny , nz V n′′x , n′′y , n′′z = δ n′x ,n′′x δ n′y ,n′′y In′z ,n′′z , (9.149)
l
where
l/2 l l
n′′z π z ′ + n′z π z ′ +
In′z ,n′′z = dz ′ sin 2
sin 2
z′ . (9.150)
−l/2 l l
Note that In′z ,n′′z = 0 if n′z = n′′z (since for that case the integrand is an
odd function of z ′ ), and thus
) ′ ′ ′ *
nx , ny , nz V n′′x , n′′y , n′′z ∝ δ n′x ,n′′x δ n′y ,n′′y 1 − δ n′z ,n′′z . (9.151)
With the )help of the above result* it is easy to see that all the matrix
elements n′x , n′y , n′z V n′′x , n′′y , n′′z that are needed for first order per-
turbation theory, for both non-degenerate and degenerate energy levels,
vanish, and consequently, to first order in E the energy eigenstates remain
unchanged.
6. To lowest order in perturbation theory the ground state energy is given
by
∞ ∞
Egs = ω+λmω 2
dx1 dx2 ϕ20 (x1 − a) ϕ20 (x2 + a) (x1 − x2 )2 +O λ2 ,
−∞ −∞
(9.152)
= ω + 2λmω2 + a2 + 2λmω 2 a2 + O λ2
2mω
= ω+λ ω + 4mω 2 a2 + O λ2 .
(9.155)
Note that this problem can be also solved exactly by employing the co-
ordinate transformation
x1 + x2
x+ = √ , (9.156)
2
x1 − x2
x− = √ . (9.157)
2
The inverse transformation is given by
x+ + x−
x1 = √ , (9.158)
2
x+ − x−
x2 = √ . (9.159)
2
The following holds
x21 + x22 = x2+ + x2− , (9.160)
and
ẋ21 + ẋ22 = ẋ2+ + ẋ2− . (9.161)
Thus, the Lagrangian of the system can be written as
m ẋ21 + ẋ22
L= − V (x1 , x2 )
2
m ẋ2+ + ẋ2− 1 √
= − mω 2 x2+ + x2− − 2a 2x− + 2a2 + 4λx2−
2 2
= L+ + L− ,
(9.162)
where
mẋ2+ 1
L+ = − mω2 x2+ , (9.163)
2 2
and
!
√ "2
mẋ2− 1 a 2 8λa 2
L− = − mω 2 (1 + 4λ) x− − + . (9.164)
2 2 1 + 4λ 1 + 4λ
2 nx πx′ ny πy ′
ψ(0) ′ ′ ′ ′
nx ,ny (x , y ) = x , y |nx , ny = sin sin , (9.172)
a a a
and the corresponding eigenenergies are
2 2
π n2x + n2y
En(0)
x ,ny
= , (9.173)
2ma2
where nx = 1, 2, · · · and ny = 1, 2, · · · .
a) The ground state (nx , ny ) = (1, 1) is nondegenerate, thus to first
order in w0
2 2
π
E0 = + 1, 1|W |1, 1
ma2
a/2 a/2
2 2
π 4w0 πx πy
= + 2 sin2 dx sin2 dy
ma2 a a a
0 0
2 2
π w0
+ ,
ma2 4
(9.174)
b) The first excite state is doubly degenerate. The matrix of the per-
turbation in the corresponding subspace is given by
1, 2|W |1, 2 1, 2|W |2, 1
2, 1|W |1, 2 2, 1|W |2, 1
a/2
/ a/2
/ a/2
/ a/2
/
2 πx 2 2πy πx 2πx 2πy πy
sin a dx sin a dy sin a sin a dx sin a sin a dy
4w0 0 0 0 0
= 2 a/2 a/2 a/2 a/2
a / 2πx 1πx
/ πy 2πy
/ 2 2πx
/ 2 πy
sin a sin a dx sin a sin a dy sin a dx sin a dy
0 0 0 0
1 16
= w0 4 9π2 ,
16 1
9π2 4
(9.175)
To first order in perturbation theory the eigenenergies are found
by adding the eigenvalues of the above matrix to the unperturbed
(0) (0)
eigenenergy E1,2 = E2,1 . Thus, to first order in w0
5 2 π2 w0 16w0
E1,± = + ± + O w02 . (9.176)
2ma2 4 9π 2
9. For the unperturbed case β = 0 one has
H0 |nx , ny = ω (nx + ny + 1) |nx , ny , (9.177)
where nx , ny = 0, 1, 2, · · · . Using the identities
0
x= ax + a†x , (9.178)
2mω
0
y= ay + a†y , (9.179)
2mω
the perturbation term V1 = βmω 2 xy can be expressed as
ω
V1 = β ax + a†x ay + a†y .
2
a) For the ground state |0, 0 , which is nondegenerate, one has
| nx , ny | V1 |0, 0 |2
E0,0 (β) = ω + 0, 0| V1 |0, 0 +
E0,0 (0) − Enx ,ny
nx ,ny =0,0
=0
2
| 1, 1| V1 |0, 0 |
= ω+
2 ω
2
ωβ
2
= ω−
2 ω
β2
= ω 1− .
8
(9.180)
Thus the degeneracy is lifted and the energies are given by 2 ω (1 ± β/4).
Note that this problem can be also solved exactly by employing the
coordinate transformation
x+y
x′ = √ , (9.182)
2
x−y
y′ = √ . (9.183)
2
The inverse transformation is given by
x′ + y ′
x= √ , (9.184)
2
x′ − y ′
y= √ . (9.185)
2
The following hold
x2 + y 2 = x′2 + y ′2 , (9.186)
2 2 ′2 ′2
ẋ + ẏ = ẋ + ẏ , (9.187)
1 ′2
xy = x − y ′2 . (9.188)
2
Thus, the Lagrangian of the system can be written as
m ẋ2 + ẏ2
L= − V (x1 , x2 )
2
m ẋ′2 + ẏ ′2 mω 2 ′2 βmω 2 ′2
= − x + y ′2 − x − y ′2
2 2 2
= L+ + L− ,
(9.189)
where
mẋ′2 mω 2
L+ = − (1 + β) x′2 , (9.190)
2 2
and
mẏ ′2 mω 2
L− = − (1 − β) y′2 . (9.191)
2 2
Thus, the system is composed of two decoupled harmonic oscillators,
and therefore, the exact eigenenergies are given by
1 1
En+ ,n− = ω 1 + β nx + + 1 − β ny + , (9.192)
2 2
where nx , ny = 0, 1, 2, · · · . To second order in β one thus has
nx − ny nx + ny + 1 2
En+ ,n− = ω nx + ny + 1 + β− β + O β3 .
2 8
(9.193)
ω1 ω1
m| V |n = n (n − 1)δ m,n−2 + (n + 1) (n + 2)δ m,n+2 ,
2 2
(9.194)
thus
1 | m| V |n |2
En (ω 1 ) = ω 0 n + + n| V |n + + O ω31
2 En (0) − Em (0)
m=n
=0
1 ω 21
= ω0 n + + [n (n − 1) − (n + 1) (n + 2)] + O ω 31
2 8ω0
ω2 1
= ω0 1 − 12 n+ + O ω 31 .
2ω0 2
(9.195)
The exact energy eigenvalues can be calculated for this case as follows.
The Hamiltonian H including the perturbation is given by
1 ω1 † †
H = ω 0 a† a + + a a + aa . (9.196)
2 2
b = ua + va† . (9.197)
cosh θ − sinh θ b a
= . (9.202)
− sinh θ cosh θ b† a†
−1 1
H = ω eff b† b + , (9.210)
2
where
0 | 1, m′ | V |1, 0 |2
Em=0 = Em=0 + 1, 0| V |1, 0 + 0 0 = 0 . (9.232)
Em=0 − Em ′
m′ =±1
0
For the degenerate eigenenergy Em=±1 = 2 α the perturbation in
the subspace spanned by {|1, −1 , |1, 1 } is given in matrix form by
01
˙ 2β
Vm=±1 = , (9.233)
10
12. For the unperturbed case V = 0, the eigenvectors and eigenenergies are
related by
0
(Hr + Ha ) |n, σ = En,σ |n, σ , (9.235)
0 1 ωa
En,σ = ωr n + +σ . (9.236)
2 2
0 | n′ , σ′ | V |n, σ |2
En,σ = En,σ + n, σ| V |n, σ + 0 − E0
. (9.237)
En,σ n′ ,σ ′
n′ ,σ ′ =n,σ
Using √
V |n, + = ga† |n, − = g n + 1 |n + 1, − , (9.238)
√
V |n, − = ga |n, + = g n |n − 1, + , (9.239)
one finds for σ = +1
1 ωa g 2 (n + 1)
En,+1 = ωr n + + +
2 2 ω a − ωr
g2
g2 1 ωa + ∆
= ωr + n+ + ,
∆ 2 2
(9.240)
and for σ = −1
1 ωa g2 n
En,−1 = ωr n + − −
2 2 ω a − ωr
g2
g2 1 −ω a + ∆
= ωr − n+ + ,
∆ 2 2
(9.241)
where
∆ = ωa − ωr . (9.242)
g2 1 1 g2
En,σ = ωr + σ n+ + σω a + . (9.243)
∆ 2 2 ∆
Thus, according to the above result (9.243), the energies of the states
(n, +1) and (n + 1, −), which are degenerate for the case where ω r =
ωa and where g = 0, are given to second order in g by
g2 ∆
En,+1 = ωr + (n + 1) + , (9.244)
∆ 2
and
g2 ∆
En+1,−1 = ωr − (n + 1) − . (9.245)
∆ 2
0 1 σ
En,σ = ω n+ + , (9.246)
2 2
g2 1 + Σz
[L, H] = − g a† Σ− + aΣ+ + 2 + a† aΣz , (9.256)
∆ 2
and thus
g2 g2
H′ = ωr + Σz a† a + ωa + Σz
∆ 2 ∆
g2
ωr + ∆ g 3
+ +O .
2 ∆
(9.257)
Note that to second order in g/∆ the states |n, σ are eigenvalues of
H′ , and the following holds
where
g2 1 1 g2
En,σ = ωr + σ n+ + σω a + . (9.259)
∆ 2 2 ∆
∆ = ωa − ωr , (9.262)
or in a matrix form
|n, +
H
|n + 1, −
10 ωn cos θ sin θ |n, +
= ω r (n + 1) + ,
01 2 sin θ − cos θ |n + 1, −
(9.263)
where
ωn = ∆2 + 4g 2 (n + 1) , (9.264)
√
2g n + 1
tan θ = . (9.265)
∆
Thus, the states |n+ and |n− , which are given by [see Eqs. (6.221)
and (6.222)]
θ θ
|n+ = cos |n, + + sin |n + 1, − , (9.266)
2 2
θ θ
|n− = − sin |n, + + cos |n + 1, − , (9.267)
2 2
are eigenstates of H and the following holds
and
N |n + 1, − = (n + 1) |n, − , (9.281)
N |n, + = (n + 1) |n, + . (9.282)
Thus, the operator I, which is defined by
I = N −1/2 S , (9.283)
satisfies
I |n, + = |n + 1, − , (9.284)
I |n + 1, − = − |n, + . (9.285)
Therefore, Eqs. (9.266) and (9.267) can be rewritten as
|n+ = U |n, + , (9.286)
|n− = U |n + 1, − , (9.287)
where
θ θ
U = cos + sin I . (9.288)
2 2
Furthermore, with the help of Eq. (9.275) one finds that [note that
I 2 |n, + = − |n, + and I 2 |n + 1, − = − |n + 1, − ]
I 2gN 1/2
U = exp tan−1 . (9.289)
2 ∆
To first order in g/∆ the following holds [compare with Eq. (9.80)]
g g 3
U = e∆S + O . (9.290)
∆
f) With the help of Eqs. (2.178) and (9.289) one finds that
1 1
H′ = H + [L, H] + [L, [L, H]] + [L, [L, [L, H]]] + · · · , (9.291)
2! 3!
where
L = −Sf (N ) , (9.292)
and
g 3 g 5
f 3 (x) = +O . (9.295)
∆ ∆
Using the commutation relations
, - , -
a† Σ− + aΣ+ , a† a = − a† Σ− + aΣ+ , |+ +| = −a† Σ− + aΣ+ ,
(9.296)
[H, N ] = 0 ,
where
1
[[H, S] , S] = − ∆ 1 + 2 a† a + Σz
2
1
−4 g a† a + a† Σ− + aΣ+ + 2 g a† Σ− + aΣ+ Σz ,
2
(9.303)
where
[[[H, S] , S] , S]
1
= 4 ∆ a† a + a† Σ− + aΣ+ − 2 ∆ a† Σ− + aΣ+ Σz
2
1 1
−4 g a† a + 1 + 2 a† a + Σz
2 2
2
−8 g a† Σ− + aΣ+
1
+2 g 1 + 2 a† a + Σz Σz .
2
(9.305)
By combining the above results one finds that
−1 ′ 4g 4 4g 4 †
H = ωr − 3
+ ξΣz − a aΣz a† a
3∆ 3∆3
1 ωr ξ
+ (ωa + ξ) Σz + +
2 2 2
g 5
+O .
∆
(9.306)
where
g2 4g 2
ξ= 1− . (9.307)
∆ 3∆2
p2x + p2y 1
H0 = + mω 2 x2 + y 2 = ω (Nx + Ny + 1) , (9.308)
2m 2
where Nx = a†x ax , Ny = a†y ay , and
βω
V = L2z
βω
= (xpy − ypx )2
βω , -2
= i ax a†y − a†x ay
3 2 2
4
= −β ω a2x a†y + a†x a2y − ax a†x a†y ay − a†x ax ay a†y
3 2 2
4
= −β ω a2x a†y + a†x a2y − (1 + Nx ) Ny − Nx (1 + Ny ) .
(9.309)
a) For the case β = 0 the ground state |0, 0 is nondegenerate and has
energy E0,0 = ω. Since V |0, 0 = 0 one finds to second order in β
1 | nx , ny | V |0, 0 |2
E0,0 = ω+ 0, 0| V |0, 0 − = ω+O β 3 .
ω nx ,ny =0,0
n x + n y
(9.310)
b) For the case β = 0 the first excited state is doubly degenerate
H0 |1, 0 = 2 ω |1, 0 , (9.311)
H0 |0, 1 = 2 ω |0, 1 . (9.312)
The matrix of V in the basis {|1, 0 , |0, 1 } is given by
1, 0| V |1, 0 1, 0| V |0, 1
0, 1| V |1, 0 0, 1| V |0, 1
1, 0| [(1 + Nx ) Ny + Nx (1 + Ny )] |1, 0 0
=β ω
0 0, 1| [(1 + Nx ) Ny + Nx (1 + Ny )] |0, 1
10
=β ω .
01
(9.313)
Thus to first order in β the first excited state remains doubly de-
generate with energy 2 ω (1 + β). Note - The exact solution can be
found using the transformation
1
ad = √ (ax − iay ) , (9.314)
2
1
ag = √ (ax + iay ) . (9.315)
2
The following holds
3 4 , -
ad , a†d = ag , a†g = 1 ,
1 † 1 †
a†d ad + a†g ag = ax + ia†y (ax − iay ) + a − ia†y (ax + iay )
2 2 x
= a†x ax + a†y ay ,
(9.316)
and
1 † 1 †
a†d ad − a†g ag = a + ia†y (ax − iay ) − a − ia†y (ax + iay )
2 x 2 x
= i ax a†y − a†x ay ,
(9.317)
thus
H0 = ω (Nd + Ng + 1) , (9.318)
2
V = β ω (Nd − Ng ) , (9.319)
and the exact eigen vectors and eigenenergies are given by
3 4
(H0 + V ) |nd , ng = ω nd + ng + 1 + β (nd − ng )2 |nd , ng .
(9.320)
=0, (9.327)
(9.328)
since the integrand is clearly an odd function of y. Thus to first order in
V0 the energies are unchanged
2 2 2
π n
En = + O V02 . (9.329)
2ml2
15. For the case ε = 0 the exact wave functions are given by
0
(0) 2 nπx
ψn (x) = sin , (9.330)
L L
and the corresponding eigenenergies are
2 2 2
π n
En(0) = , (9.331)
2mL2
where n is integer. To first order in ε the energy of the ground state n = 1
is given by
L 2
(0) ε (0)
E1 = E1 + dx ψ1 (x) x + O ε2
L 0
(0) 2ε L πx
= E1 + 2 dx sin2 x + O ε2
L 0 L
(0) ε
= E1 + + O ε2
2
(9.332)
16. For the case λ = 0 the exact wave functions of the eigenstates are given
by
2 nx πx ny πy
ψ(0)
nx ,ny (x, y) = sin sin , (9.333)
l l l
and the corresponding eigenenergies are
2 2
π n2x + n2y
En(0)
x ,ny
= , (9.334)
2ml2
where nx and ny are non-zero integers.
a) The ground state is non degenerate thus to 1st order the energy is
given by
l l 2
(0) (0)
E0 = E1,1 + ψ1,1 W dxdy
0 0
2 2
π
=
ml2
2 2 l l
π πx 2 πy
+ 2 4λ sin2 sin δ (x − lx ) δ (y − ly ) dxdy
ml 0 0 l l
2 2
π πlx πly
= 1 + 4λ sin2 sin2 .
ml2 l l
(9.335)
b) The first excited state is doubly degenerate. The matrix of the per-
turbation W in the eigen subspace is given by
2, 1| W |2, 1 2, 1| W |1, 2
W = ˙
1, 2| W |2, 1 1, 2| W |1, 2
! "
2 2
π sin2 2πll x sin2 πlly sin 2πll x sin πll x sin πll y sin 2πll y
= 4λ
ml2 sin πllx sin 2πll x sin 2πll y sin πll y sin2 πllx sin2 2πll y
! "
2 2
π 4 sin2 πllx cos2 πllx sin2 πll y 4 cos πllx sin2 πllx cos πll y sin2 πll y
= 4λ
ml2 4 cos πll x sin2 πllx cos πll y sin2 πll y 4 sin2 πllx sin2 πll y cos2 πlly
πl
! "
16λ 2 π 2 sin2 πllx sin2 ly cos2 πllx cos πllx cos πlly
= .
ml2 cos πllx cos πll y cos2 πll y
(9.336)
w1 = 0 , (9.337)
and
πly πly
16λ 2 π2 sin2 πlx
l sin2 l cos2 πlx
l + cos2 l
w2 = . (9.338)
ml2
17. The unperturbed Hamiltonian (λ = 0) can be written as
L2 − L2z L2
H= + z
2Ixy 2Iz
2
L 1 1
= + − L2z ,
2Ixy 2Iz 2Ixy
(9.339)
thus the states |l, m (the standard eigenstates of L2 and Lz ) are eigen-
states of H and the following holds
where
2 l (l + 1) 1 1
El,m = + − m2 . (9.341)
2Ixy 2Iz 2Ixy
L2+ + L2−
V =λ . (9.344)
4Ixy
To second order in λ the energy of the ground state is found using Eq.
(9.32)
| l′ , m′ | V |0, 0 |2
E0 = E0,0 + 0, 0| V |0, 0 + +O λ3 . (9.345)
E0,0 − El′ ,m′
l′ ,m′ =0,0
E0 = 0 + O λ3 . (9.348)
H = H1 + H2 + V , (9.349)
where
p21
H1 = − αδ (x1 ) , (9.350)
2m
p22
H2 = − αδ (x2 ) , (9.351)
2m
and
V = λδ (x1 − x2 ) . (9.352)
d2 2m
+ 2 (E + αδ (x1 )) ψ(1) (x1 ) = 0 . (9.353)
dx21
(1) mα2
E0 = − .
2 2
The ground state of H2 can be found in a similar way. Thus, the normal-
ized wavefunction of the only bound state of H1 + H2 , which is obviously
the ground state, is given by
mα mα mα
ψ0 (x1 , x2 ) = 2
exp − 2
|x1 | exp − 2
|x2 | , (9.355)
and
into Eq. (9.96) and collecting terms having the same order in Ω (up to
second order) yield
(D − λn0 ) |n0 = 0 , (9.360)
(D − λn0 ) |n1 + (V − λn1 ) |n0 = 0 , (9.361)
(D − λn0 ) |n2 + (V − λn1 ) |n1 − λn2 |n0 = 0 . (9.362)
We further require normalization
n|n = 1 , (9.363)
n0 |n ∈ R . (9.364)
thus for m = n
m0 | V |n0
m0 |n1 = , (9.372)
λn0 − λm0
thus with the help of Eq. (9.95) one has
m0 | V |n0
|n1 = |m0 . (9.373)
m
λn0 − λm0
n0 | V |m0 m0 | V |n0
λn2 = . (9.375)
m
λn0 − λm0
Thus, using this result together with Eq. (9.371) one finds
λ = λn0 + Ω n0 | V |n0
n0 | V |m0 m0 | V |n0
+ Ω2 + O Ω3 .
m
λn0 − λm0
(9.376)
20. The condition [9.102] together with Eq. (9.98) can be used to evaluate
the matrix elements of L
k| Ṽ |k′
k| L |k′ = λ . (9.377)
Ek − Ek′
With the help of Eq. (2.178) one finds that
22. The energy eigenvalues Ekl of the radial equation of the hydrogen atom,
which is given by [see Eq. (7.61)]
− 2 d2 e2 l (l + 1) 2
− + ukl (r) = Ekl ukl (r) , (9.387)
2µ dr2 r 2µr2
where µ is the reduced mass and e is the electron charge, are [see Eq.
(7.78)]
µe4
Ekl = − , (9.388)
2 2 (k + l)2
where k = 1, 2, 3, · · · . The quantum number l can formally be treated as a
real number, which is not restricted to take integer values only. Consider
the case where the integer l is replaced by l + ǫ, where 0 ≤ ǫ ≪ 1. While
the exact eigenenergies can still be evaluated by Eq. (7.78)
µe4
Ekl = − , (9.389)
2 2 (k + l + ǫ)2
µe4
Ekl = − + klm| VH |klm + · · · , (9.390)
2 2 (k + l)2
where the perturbation VH is given by
2
[(l + ǫ) (l + ǫ + 1) − l (l + 1)]
VH =
2µr2
2
(2l + 1) ǫ
= + O ǫ2 ,
2µr2
(9.391)
By comparing both results for Ekl one finds that
2µ2 e4
klm| r−2 |klm = , (9.392)
4 (2l + 1) (k + l)3
or in terms of the quantum number n = k + l
2
nlm| r−2 |nlm = , (9.393)
a20 (2l + 1) n3
where
2
a0 = . (9.394)
µe2
d |α
i = H |α , (10.1)
dt
where the Hermitian operator H = H† is the Hamiltonian of the system. The
time evolution operator u (t, t0 ) [see Eq. (4.4)] relates the state vector |α (t0 )
at time t0 with its value |α (t) at time t
i (t − t0 )
u (t, t0 ) = exp − H . (10.3)
In this chapter we consider the more general case where H is allowed to vary
in time. We first derive a formal expression for the time evolution operator
u (t, t0 ) applicable for general H. Then we present the perturbation theory
expansion of the time evolution operator, and discuss approximation schemes
to evaluate u (t, t0 ).
Dividing the time interval (t0 , t) into N sections of equal duration allows
expressing the time evolution operator as
N
?
u (t, t0 ) = u (tn , tn−1 ) , (10.4)
n=1
where
tn = t0 + nǫ , (10.5)
Chapter 10. Time-Dependent Perturbation Theory
and where
t − t0
ǫ= . (10.6)
N
Furthermore, according to the Schrödinger equation (4.7), the following holds
iǫ
u (tn−1 + ǫ, tn−1 ) = 1 − H (tn ) + O ǫ2 . (10.7)
N iǫ
=
u1 (t, t0 ) = − lim u0 (t, tn ) H1 (tn ) u0 (tn , t0 )
N→∞ n=1
i /t
=− dt′ u0 (t, t′ ) H1 (t′ ) u0 (t′ , t0 ) ,
t0
(10.13)
and
N−1
= N
= ǫ 2
u2 (t, t0 ) = − lim
N→∞ n=1 m=n+1
iλ /t λ2 /t /t′
O (t) = 1 − dt′ H1I (t′ ) − 2
dt′ dt′′ H1I (t′ ) H1I (t′′ ) , (10.16)
t0 t0 t0
λ2 /t ′
/t′
− 2
dt dt′′ ( H1I (t′ ) H1I (t′′ ) + H1I (t′′ ) H1I (t′ ) )
t0 t0
3
+O λ ,
(10.18)
or
! "2
2 λ2 /t ′ ′
| O (t) | = 1 + 2
dt H1I (t )
t0
λ2 /t /t
− 2
dt′ dt′′ H1I (t′ ) H1I (t′′ ) ,
t0 t0
(10.19)
thus
2 λ2 /t /t
| O (t) | = 1 − 2
dt′ dt′′
t0 t0
or
2 λ2 /t /t
| O (t) | = 1 − 2
dt′ dt′′ ∆H1I (t′ ) ∆H1I (t′′ ) , (10.21)
t0 t0
where
where
i (t − t0 ) iEn (t − t0 )
u0 (t, t0 ) = exp − H0 = exp − |an an | .
n
(10.25)
Assuming that initially at time t0 the system is in state |an , what is the
probability to find it later at time t > t0 in the state |am ? The answer to
this question is the transition probability pnm , which is given by
λ2 /t /t′
− 2
dt′ dt′′ am | H1I (t′ ) H1I (t′′ ) |an
t0 t0
3
+O λ ,
(10.27)
thus
iλ /t
pnm = δ nm − dt′ am | H1I (t′ ) |an
t0
2
λ2 /t ′
/t′ ′′ ′ ′′ 3
− 2
dt dt am | H1I (t ) H1I (t ) |an + O λ .
t0 t0
(10.28)
In what follows, we calculate the transition probability pnm to lowest non-
vanishing order in λ for the case where n = m, for which the dominant
contribution comes from the term of order λ in Eq. (10.28). For simplicity
the initial time t0 , at which the perturbation is turned on, is taken to be zero,
i.e. t0 = 0. We consider below the following cases:
where
Em − En
ω mn = . (10.30)
/t Ωt
′ Ωt sin
dt′ eiΩt = 2ei 2 2
, (10.31)
0 Ω
4 sin2 ωmn
2
t
2
pnm = 2 2
| am | λH1 |an | . (10.32)
ω mn
Note that in the limit t → ∞ one finds with the help of Eq. (10.31) that
4 sin2 Ωt
2
/t iΩt′ ′ 2
lim = lim e dt
t→∞ Ω2 t→∞ 0
/t /t ′ ′′
= lim dt′ dt′′ eiΩ(t −t )
t→∞ 0 0
/t
= 2πδ (Ω) dt′
0
= 2πtδ (Ω) .
(10.33)
In this limit pnm is proportional to the time t, i.e. pnm can be expressed as
pnm = wnm t, where wnm is the transition rate, which is given by
2π
wnm = 2
δ (ω mn ) | am | λH1 |an |2 . (10.34)
The delta function δ (ωmn ) ensures that energy is conserved in the limit
of long time, and transitions between states having different energies are
excluded. However, such transitions have finite probability to occur for any
finite time interval ∆t. On the other hand, as can be see from Eq. (10.32)
(see also the figure below, which plots the function f (x) = sin2 x/x2 ), the
probability is significant only when ω mn ∆t 1, or alternatively when
∆E∆t , (10.35)
where ∆E = ω mn .
0.8
0.6
0.4
0.2
-10 -8 -6 -4 -2 0 2 4 x 6 8 10
2
The function f (x) = sin x/x2 .
In the long time limit, i.e. in the limit t → ∞, the probability pnm is
found using Eq. (10.33) to be proportional to the time t, i.e. pnm = wnm t,
where the transition rate wnm is given by
:
2π
δ (ωmn − ω) | am | λK |an |2 ω mn ≃ ω
wnm ≃ 2π2 2 . (10.39)
2 δ (ω mn + ω) am | λK† |an ω mn ≃ −ω
In many cases of interest the final state |am lie in a band of dense states. Let
wn be the total transition rate from the initial state |an . Assume that the
matrix element am | λK |an does not vary significantly as a function of the
energy Em . For this case the total rate wn can be expressed in terms of the
density of states g (Em ) (i.e. number of states per unit energy) in the vicinity
of the final state |am [see Eq. (10.39)] as
2π
wn = g (Em ) | am | λK |an |2 , (10.40)
10.3.3 H1 is Separable
where f (t′ ) is a real function of time and where H̄1 is time independent
Hermitian operator. To lowest nonvanishing order in λ Eq. (10.28) yields
2
1 /t ′ 2
pnm = 2
dt′ eiωmn t f (t′ ) am | λH̄1 |an . (10.42)
0
10.4 Problems
H = H0 + Hp , (10.43)
where
H0 = ωa† a , (10.44)
a and a† are the annihilation and creation operators (as defined in chapter
5), ω is positive, φ is real and ζ (t) is an arbitrary real function of time t.
2. Consider a particle having mass m moving under the influence of a one
dimensional potential given by
mω20 x2
V (x) = , (10.46)
2
where the angular resonance frequency ω 0 is a constant. A perturbation
given by
and where both α and τ are real. Given that the system was initially at
time t → −∞ in the ground state |0 of the unperturbed Hamiltonian,
find the transition probability pn0 to the number state |n in the limit
t → ∞. Compare your approximated result with the exact result given
by Eq. (5.345).
4. Consider a spin 1/2 particle. The Hamiltonian is given by
H = ωSx , (10.50)
both q (the charge) and E0 (the electric field) are assumed to be con-
stants, and x is the position operator. Calculate to lowest nonvanishing
order in E0 the expectation value x (t) at time t ≥ 0.
10.5 Solutions
1. Expressing the ket vector state as
and
it , -
H0 , a†2 = 2iωta†2 , (10.60)
thus
where
, -
S (ξ, φ) = exp ξ e−2iφ a2 − e2iφ a†2 , (10.65)
and where
t
ξ= dt′ ζ (t′ ) , (10.66)
0
where
∞ 2
1 ′
µ= dt′ eiω0 t f (t′ )
2m ω 0 −∞
α2 1 2 2
= e− 2 ω0 τ .
2m ω 0
(10.70)
The exact result is found from Eq. (5.345)
e−µ µn
pn = . (10.71)
n!
To first order in µ both results agree.
p2,1 = dt e e τ
τa 0
2
2 /a ′ ′ 2πx′ πx′
× dx x sin sin ,
a 0 a a
(10.76)
thus [see Eq. (5.139)
256λ2 9π4 2 τ 2
p2,1 = exp − . (10.77)
81π3 8m2 a4
6. For the present case to second order in λ Eq. (10.27) becomes [see Eq.
(10.25)]
iEm (t−t0 )
e− am | u (t, t0 ) |an
λ2 /t /t′ ′ ′′
=− 2
dt′ dt′′ eiωml t +iωln t am | H1 (t) |al al | H1 (t) |an ,
l 0 0
(10.78)
where
Em − En
ωmn = , (10.79)
or
iEm (t−t0 )
e− am | u (t, t0 ) |an
2 ′
λ Kml Kln /t /t ′ ′′
=− 2
dt′ dt′′ ei(ωml −ω)t +i(ωln −ω)t
l 0 0
λ2 Klm
∗
Kln /t /t′ ′ ′′
− 2
dt′ dt′′ ei(ωml +ω)t +i(ωln −ω)t
l 0 0
2 ′
∗ /t /t
λ Kml Knl ′ ′′
− 2
dt′ dt′′ ei(ωml −ω)t +i(ωln +ω)t
l 0 0
2 ∗
λ Klm ∗ /t
Knl /t′ ′ ′′
− 2
dt′ dt′′ ei(ωml +ω)t +i(ωln +ω)t ,
l 0 0
(10.80)
where
the transition rate wnm can be evaluated in the long time limit. To that
end it is assumed that ω = ±ωml and ω = ±ωln (i.e. it is assumed that the
harmonic perturbation is not at resonance with any first order transition
between the initial |an or final |am states and an intermediate state
|al ), and it is further assumed that ω mn > 0 and that ω ≥ 0. Under
/t ′
these assumptions only the terms proportional to 0 dt′ ei(ωmn −2ω)t in
Eq. (10.84) are taken into account, and consequently the transition rate
wnm becomes
2
2π λ2 Kml Kln
wnm = 4
δ (ωmn − 2ω) . (10.86)
(ωln − ω)
l
and
where x0 is a constant.
What is the range of validity of the zero order approximation? As can be
seen by comparing Eq. (11.6) with Eq. (11.9), the approximation W ≃ W0
is valid when the first term in Eq. (11.6) is negligibly small in absolute value
in comparison with the second one, namely when
2
d2 W dW
2
≪ . (11.11)
dx dx
dλ
≪ 2π . (11.13)
dx
This means that the approximation is valid provided that the change in
wavelength over a distance of one wavelength is small.
where
x
1 i
ϕ± (x) = √ exp ± dx′ p (x′ ) , (11.17)
p
x0
1
J= |C+ |2 − |C− |2 . (11.20)
m
Thus, the current density J associated with the state ϕ+ (x) is positive,
whereas J < 0 for ϕ− (x). Namely, ϕ+ (x) describes a state propagating from
left to right, whereas ϕ− (x) describes a state propagating in the opposite
direction.
V (x ) − E Im( x )
(a) (b)
Γθ
x θ
a a Re( x )
Im( x ) Im( x )
(c) (d)
Θ+ Θ−
a Re( x ) a Re( x )
Fig. 11.1. (a) The turning point at x = a. (b) The integration trajectory Γθ . The
singly connected region Θ+ (c) and Θ− (d).
where
dV
α=− , (11.23)
dx x=a
x = a + ρeiθ , (11.25)
where the integration trajectory Γθ [see Fig. 11.1 (b)] contains two sections,
the first along the real axis from x = a to x = a + ρ and the second along the
′
arc x = a + ρeiθ from θ′ = 0 to θ ′ = θ. With the help of the approximation
(11.26) one finds that
ρ
√ θ
i 2mα ′
I± (θ) = ± dρ′ ρ′ + iρ3/2 dθ′ ei3θ /2
0 0
√ 2i e
3
2 iθ − 1
i 2mα 2 3/2
=± ρ − iρ3/2
3 3
√
2i 2mαρ3/2 32 iθ
=± e
√ 3
2 2mαρ3/2 i(π(1∓ 12 )+ 32 θ)
= e ,
3
(11.27)
thus
√
2 2mαρ3/2
I± (π) = ± , (11.28)
√ 3 3/2
2 2mαρ
I± (−π) = ∓ . (11.29)
3
The last result allows expressing the analytical continuation of the wavefunc-
tion given by Eq. (11.21) for the case x > a and evaluate its value at the
point x = a − ρ. For the case where the singly connected region Θ+ (Θ− ) is
employed, this is done using integration along the trajectory Γπ (Γ−π ), and
the result is labeled as ψ+ (a − ρ) [ψ− (a − ρ)]
Comparing Eqs. (11.30) and (11.31) with Eq. (11.32) shows that for each of
the two choices Θ+ and Θ− the analytical continuation yields one exponential
term having the same form as the one in Eq. (11.32), and another one, which
diverges in the limit x → −∞. Excluding the diverging terms one finds that
continuity of the non diverging term requires that
C+ C−
C= −iπ/4
= , (11.33)
e eiπ/4
and thus the tailored wavefunction is given by
/x
√C|p| exp 1 dx′ |p| x < a
ψ (x) = a . (11.34)
2C 1
/x ′
√p cos dx p − π4 x > a
a
The fact that analytical continuation of the wavefunction in the region x < a
along the trajectory Γπ (Γ−π ) yields only the right to left (left to right) prop-
agating term in the region x > a, and the other term is getting lost along the
way, can be attributed to the limited accuracy of the WKB approximation.
As can be seen from Eq. (11.27), along the integration trajectory Γθ near the
point θ = ±π/3 one term becomes exponentially larger than the other, and
consequently, within the accuracy of this approximation the small term gets
lost.
It is important to keep in mind that the above result (11.34) is obtained
by assuming a particular form for the solution in the region x < a, namely
by assuming that in the classically forbidden region the coefficient of the
exponentially diverging term vanishes. This tailoring role will be employed in
the next section that deals with bound states in a classically accessible region
between two turning points [see Fig. 11.2(a)]. On the other hand, a modified
tailoring role will be needed when dealing with quantum tunneling. For this
case, which will be discussed below, we seek a wave function having the form
V (x ) − E V (x ) − E
(a) (b)
x x
a b a b
C+ 1
/x C− /x
√|p|
exp dx′ |p| + √ exp − 1 dx′ |p| x<a
a |p| a
ψ (x) = /x . (11.35)
C i iπ
√
p exp dx′ p + 4 x>a
a
Thus, in this case only the term describing propagation from left to right
is kept in the region x > a, and the coefficient of the other term in that
region that describes propagation in the opposite direction is assumed to
vanish. Using the same tailoring technique as in the previous case one find
that C+ = 0 and C− = C, and thus
/x
√C|p| exp − 1 dx′ |p| x < a
ψ (x) = a . (11.36)
C i
/x ′ iπ
p√ exp dx p + x > a
4
a
The requirement ψa (x) = ψb (x) can be satisfied for any x in the region
a ≤ x ≤ b only if
b
1 π
dx′ p = + nπ . (11.39)
2
a
Note that the time period T of classical oscillations between the turning
points x = a and x = b is given by
b
dx′
T =2 . (11.43)
v
a
where v (x) = p (x) /m is the local classical velocity. Thus, by choosing the
pre-factor to be real, one finds that the normalized wavefunction is given by
x
0
m 1 π
ψ (x) = 2 cos dx′ p − . (11.44)
pT 4
a
11.4 Tunneling
In this case we consider a classical forbidden region a ≤ x ≤ b bounded by
two turning points at x = a and x = b, namely, it is assumed that E < V (x)
for a ≤ x ≤ b and E > V (x) for x < a and for x > b [see Fig. 11.2(b)].
In classical mechanics a particle cannot penetrate into the potential barrier
in the region a ≤ x ≤ b, however such a process is possible in quantum
mechanics. Consider a solution having the form
/x ′ /x ′
√1 exp i
dx p + √r exp − i dx p x<a
p p
a a
x
/ x
/
C+ C−
ψ (x) = √|p| exp 1 dx′ |p| + √ exp − 1 dx′ |p| a ≤ x ≤ b ,
|p|
b b
t i
/x ′ iπ
√ exp dx p + 4 x>b
p
b
where
b
2
τ = exp − dx′ |p| , (11.49)
a
11.5 Problems
1. Calculate the transmission probability τ of a particle having mass m
and energy E through the potential barrier V (x), which vanishes in the
region x < 0 and which is given by V (x) = U − ax in the region x ≥ 0,
where a > 0 and where U > E.
2. Calculate the transmission probability τ of a particle having mass m and
energy E through the potential barrier V (x) = −mω 2 x2 /2, where ω > 0.
Consider the general case without assuming τ ≪ 1.
3. Consider a particle having mass m moving in a one dimensional double
well potential (see Fig. 11.3), which is assumed to be symmetric, i.e.
V (x) = V (−x). In the limit where the barrier separating the two wells
can be considered as impenetrable, each well is characterized by a set of
eigenstates having eigenenergies {En }. To lowest nonvanishing order in
the penetrability of the barrier calculate the eigenenergies of the system.
11.6 Solutions
1. The classical turning points are x = 0 and x = (U − E) /α. Thus with
the help of Eq. (11.50) one finds that
V (x )
En
x
−a a
Fig. 11.3. Double well potential.
√ (U −E)/α 0
2 2mα U −E
τ = exp − dx − x
α
0
√! "
4 2m 3
= exp − (U − E) 2 .
3 α
(11.51)
2. The factor p/ can be expressed as
.
2 2
p (x) 2m E + mω2 x
= =
;
x 2Ex20
= 2 1+ ,
x0 E0 x2
(11.52)
where x0 = /mω and where E0 = ω. For sufficiently large |x|, namely
for x2 ≫ Ex20 /E0 , one has
p (x) x E
≃ 2 + , (11.53)
x0 E0 x
i EE − 12 iρ2 e2iθ
t ρeiθ 0
exp , (11.57)
2
thus for θ = π this term becomes identical to the reflected term at x/x0 =
−ρ, which is given by
E 1 iρ2
r (ρ)i E0 − 2 exp , (11.58)
2
provided that
2
and therefor the transmission probability τ = |t| is given by
1
τ= 2πE . (11.62)
1 + e− E0
n, L| H |n, L n, L| H |n, R
Hn = . (11.63)
n, R| H |n, L n, R| H |n, R
where
γ = n, L |n, R , (11.67)
where ϕn,L (x) and ϕn,R (x) are the wavefunctions of the states |n, L
and |n, R respectively, i.e.
ϕn,L (x) = x |n, L , (11.69)
ϕn,R (x) = x |n, R . (11.70)
The main contribution to the overlap integral (11.68) comes from the
classically forbidden region |x| ≤ a, where x = ±a are turning points
(i.e., En = V (a) = V (−a)). With the help of Eq. (11.36) one finds that
! "
2 1
/x ′ /a
a |C| exp − dx |p| exp − 1 dx′ |p|
−a x
γ≃ dx (11.71)
|p|
−a
a
a
2 1 ′ dx
= |C| exp − dx |p| , (11.72)
|p|
−a −a
(11.73)
where C is the normalization factor of the WKB wavefunction, which is
approximately given by C = 2 m/T (T is the time period of classical
oscillations of a particle having energy En in a well) in the limit of large
n [see Eq. (11.44)], thus
/a dx
4 |p/m|
a
−a 1
γ≃ exp − dx′ |p| . (11.74)
T
−a
Finally, By diagonalizing the matrix Hn one finds that the two eigenen-
ergies are En (1 ± γ).
4. The radial equation for the case of hydrogen is given by [see Eq. (7.61)]
− 2 d2
+ Veff (r) ukl (r) = Ekl ukl (r) , (11.75)
2µ dr2
where µ ≃ me is the reduced mass (me is the electron’s mass), and where
e2 l (l + 1) 2
Veff (r) = − + . (11.76)
r 2µr2
The eigenenergies Ekl are calculated using the Bohr-Sommerfeld quanti-
zation rule (11.40)
or
√ 1
εI = π k + , (11.86)
2
where the integral I, which is given by
ρ2
(ρ − ρ1 ) (ρ2 − ρ)
I= dρ , (11.87)
ρ
ρ1
Ekl 1
ε=− = 2 . (11.89)
EI 1
l (l + 1) + k + 2
Comparing with the exact result (7.84) shows that the WKB result is a
good approximation provided that the quantum numbers are large.
In this chapter, which is mainly based on Ref. [4], the technique of Feynman’s
path integration is briefly reviewed.
Consider a point particle having mass m and charge q moving under the
influence of electric field E and magnetic field B, which are related to the
scalar potential ϕ and to the vector potential A by
1 ∂A
E = −∇ϕ − , (12.1)
c ∂t
and
B=∇×A. (12.2)
1
mr̈ = q E + ṙ × B . (12.5)
c
In what follows, we consider for simplicity the case where both ϕ and A are
time independent. For this case H becomes time independent, and thus the
quantum dynamics is governed by the time evolution operator, which is given
by Eq. (4.9)
Chapter 12. Path Integration
iHt
u (t) = exp − . (12.6)
The identity operator in the position representation [see Eq. (3.65)] is given
by
1r = d3 r′ |r′ r′ | . (12.9)
Inserting 1r between any two factors in Eq. (12.8) and using the notation
r′a = r′0 , (12.10)
r′b = r′N , (12.11)
t
ǫ= , (12.12)
N
one finds that
thus
N−1
# N−1
#
K (r′b , t; r′a ) = d3 r′n K r′m+1 , ǫ; r′m . (12.14)
n=1 m=0
one has
iHǫ
u (ǫ) = exp −
! 2
"
iǫ p− qc A iǫqϕ
= exp − exp − + O ǫ2 .
2m
(12.16)
Equation (12.122), which is given by
0
iǫV2 1 ir′2 ǫ
exp − = 3/2
d3 r′ exp −i V · r′ , (12.17)
2m (2πi) 2 m
In the next step the identity operator in the momentum representation [see
Eq. (3.71)], which is given by
1p = d3 p′ |p′ p′ | , (12.22)
is inserted to the left of the ket vector |r′m . With the help of Eq. (3.75),
which is given by
1 ip′ · r′
r′ |p′ = 3/2
exp , (12.23)
(2π )
1 ip′ · r′
d3 p′ exp = δ (r′ ) , (12.25)
(2π )3
and thus
m 3/2 iǫ
K r′m+1 , ǫ; r′m = exp Lm + O ǫ3/2 , (12.27)
2πiǫ
where
r′m+1 −r′m 2
m ǫ q A (r′m ) + A r′m+1 r′m+1 − r′m
Lm = −qϕ (r′m )+ · . (12.28)
2 c 2 ǫ
Comparing Eq. (12.28) with the classical Lagrangian of the system, which
is given by Eq. (1.43)
1 q
L = mṙ2 − qϕ + A · ṙ , (12.29)
2 c
shows that Lm is nothing but the Lagrangian at point r′m
Lm = L (r′m ) . (12.30)
As we have discussed above, the terms of order ǫ3/2 in Eq. (12.27) are not
expected to contribute to K (r′b , t; r′a ) in the limit of N → ∞. By ignoring
these terms Eq. (12.14) becomes
! "
N/2 N−1 # N−1
′ ′ Nm 3 ′ i t ′
K (rb , t; ra ) = lim d rn exp L (rm ) .
N→∞ 2πit N m=0
n=1
(12.31)
Recall that the action in classical physics [see Eq. (1.4)] associated with a
given path is given by
S= dt L . (12.32)
where
t
1
0.8
0.6
0.4
0.2
-4 -2 0 2 x 4
-0.2
-0.4
-0.6
-0.8
-1
t
1 2 q
S= dt mṙ − qϕ + A · ṙ . (12.37)
2 c
0
Consider first the case where the vector potential vanishes, i.e. A = 0. For
this case, the system is said to be conservative, and therefor, as we have seen
in chapter 1 [see Eq. (1.29)], the energy of the system
1
E = mṙ2 + qϕ (12.38)
2
is a constant of the motion (see exercise 5 below). In terms of E the action
S (12.37), which is labeled as S0 for this case where A = 0, can be expressed
as
t
1 2
S0 = dt mṙ − qϕ
2
0
t
= dt −E + mṙ2
0
rb
= −Et + m dr · ṙ .
ra
(12.39)
where ra = r (0) and rb = r (t). Employing Eq. (12.38) again allows rewriting
S0 as
rb
Note the similarity between the second factor in the above equation and
between the WKB wavefunction [see Eq. (11.17)]. In the general case, where
A can be nonzero, the phase factor in the path integral becomes [see Eq.
(12.37)]
rb
iS iS0 iq
exp = exp exp dr · A . (12.43)
c
ra
collector
rb , t
impenetrable
long coil
B≠0
ra ,0
source
Fig. 12.1. Two-slit interference experiment with a very long impenetrable cylinder
placed near the gap between the slits.
paths going through the left slit, and another for all paths going through the
right one. Here we disregard paths crossing a slit more than one time, as their
contribution is expected to be small. In general, the difference Θ12 between
the vector potential phase factor in Eq. (12.45) associated with two different
paths r1 (t) and r2 (t) is given by
e
Θ12 = dr · A− dr · A
c
r1 (t) r2 (t)
@
e
= dr · A ,
c
(12.46)
where the closed path integral is evaluated along the path r1 (t) in the forward
direction from ra to rb , and then along the path r2 (t) in the backward
direction from rb back to ra . This integral can be calculated using Stokes’
theorem [see Eq. (12.2)]
@
e e φ
Θ12 = dr · A = ds · B = 2π , (12.47)
c c φ0
where φ is the magnetic flux threaded through the area enclosed by the closed
path, and where
hc
φ0 = (12.48)
e
is the so called flux quantum. While Θ12 vanishes for pairs of paths going
through the same slit, it has the same value Θ12 = 2πφ/φ0 (Θ12 = −2πφ/φ0 )
for all the pairs where r1 (t) goes through the left (right) path and where r2 (t)
goes through the right (left) one. Thus, we come to the somewhat surprising
conclusion that the probability density Pb is expected to be dependent on
the magnetic field. The expected dependence is periodic in the magnetic
flux φ with flux quantum φ0 period. Such dependence cannot be classically
understood, since in this example the electrons can never enter the region in
which the magnetic field B is finite, and thus the Lorentz force vanishes in
the entire region accessible for the electrons outside the impenetrable coil.
iqχ q q iqχ q
exp − p− A− ∇χ exp = p− A , (12.55)
c c c c c
iqχ iqχ
exp − H̃ exp =H, (12.56)
c c
solves the Schrödinger equation with H̃, provided that the state vector |ψ
solves the Schrödinger equation with H, and therefore
iqχ (r′ )
ψ̃ (r′ , t′ ) = exp ψ (r′ , t′ ) . (12.60)
c
Consider a point particle having mass m moving in one dimension along the
x axis under the influence of the potential V (x). The path integral (12.34)
for this case becomes
! "
N/2 N−1
# N−1
Nm i t xm+1 − xm
K (xb , t; xa ) = lim dxn exp L xm , t .
N→∞ 2πit N m=0
n=1 N
(12.61)
The solution of the Euler Lagrange equation, which is given by Eq. (1.8)
d ∂L ∂L
= , (12.63)
dt ∂ ẋ ∂x
yields the classical equation of motion of the system
dV
mẍ = − . (12.64)
dx
N/2 N−1
# N−1
imN imN
K (xb , t; xa ) = lim − dxn exp (xm+1 − xm )2 ,
N→∞ 2π t n=1
2 t m=0
(12.66)
or
α N/2 N−1
# N−1
K (xb , t; xa ) = lim dxn exp −α (xm+1 − xm )2 ,
N→∞ π n=1 m=0
(12.67)
where
imN
α=− . (12.68)
2 t
/
The first integral dx1 can be calculated using the identity
∞
3 4 0π 3 α 4
2 2
dx1 exp −α (x2 − x1 ) − α (x1 − x0 ) = exp − (x2 − x0 )2 ,
2α 2
−∞
(12.69)
/
The second integral dx2 can be calculated using the identity
∞
3 α 4 0 2π 3 α 4
2 2
dx2 exp −α (x3 − x2 ) − (x2 − x0 ) = exp − (x3 − x0 )2 .
2 3α 3
−∞
(12.70)
/
Similarly, the nth integral dxn yields
0
nπ α 2
exp − (xn+1 − x0 ) . (12.71)
(n + 1) α n+1
or
0
m im
K (xb , t; xa ) = exp (xb − xa )2 . (12.73)
2πi t 2 t
As can be seen from the classical equation of motion (12.64), a free particle
moves at a constant velocity. Thus, the classical path satisfying x (0) = xa
and x (t) = xb is given by
(xb − xa ) t′
xc (t′ ) = xa + . (12.74)
t
The corresponding classical action Sc is
m (xb − xa )2
Sc = dt′ L (x, ẋ) = . (12.75)
2t
xc (t′ )
d2 Sc m
=− . (12.76)
dxa dxb t
Thus the propagator can be expressed in terms of the classical action Sc as
;
i d2 Sc i
K (xb , t; xa ) = exp Sc . (12.77)
2π dxa dxb
As we will see below, a similar expression for the propagator is obtained also
for other cases.
x dependent, with the classical path xc (t′ ) and the corresponding classical
action Sc . Consider a general path x (t′ ) satisfying the boundary conditions
x (0) = xa and x (t) = xb . It is convenient to express the path as
where the deviation δ (t′ ) from the classical path xc (t′ ) vanishes at the end
points δ (0) = δ (t) = 0. The action associated with the path x (t′ ) can be
expressed as
S = Sc + S1 + S2 + · · · , (12.80)
where
In the general case, higher orders in such an expansion may play an important
role, however, as will be discussed below, in the classical limit the dominant
contribution to the path integral comes from the lowest order terms.
Claim. S1 = 0.
The first term in Eq. (12.84) vanishes due to the boundary conditions δ (0) =
δ (t) = 0, whereas the second one vanishes because xc (t′ ) satisfies the Euler
Lagrange equation (12.63), thus S1 = 0. The fact that S1 vanishes is a direct
consequence of the principle of least action of classical mechanics that was
discussed in chapter 1.
where
iSc
Pc (xb , t; xa ) = exp , (12.86)
i
K (t) = D [δ (t′ )] exp S2 + O δ 3 , (12.87)
and where
N/2 N−1
#
Nm
D [δ (t′ )] = lim dδ n . (12.88)
N→∞ 2πit n=1
The term K (t) is evaluated by integrating over all paths δ (t′ ) satisfying the
boundary conditions δ (0) = δ (t) = 0.
mẋ2 mω 1 xẋ mω 2 x2
L (x, ẋ) = + − , (12.92)
2 2 2
m 2
2
L (x, ẋ) = ẋc + δ̇ + ω 1 (xc + δ) ẋc + δ̇ − ω 2 (xc + δ) , (12.93)
2
thus the action associated with the path x (t′ ) can be expressed as
where
t
m
Sc = dt′ ẋ2c + ω1 xc ẋc − ω 2 x2c , (12.95)
2
0
t
3 ω1 4
S1 = m dt′ ẋc δ̇ + xc δ̇ + δ ẋc − ω 2 xc δ , (12.96)
2
0
t
m 2
S2 = dt′ δ̇ + ω1 δ δ̇ − ω 2 δ 2 . (12.97)
2
0
As we have seen above, the principle of least action implies that S1 = 0. Note
that in this case the expansion to second order in δ is exact and all higher
order terms vanish. Thus, the exact solution of this problem will also provide
an approximate solution for systems whose Lagrangian can be approximated
by a quadratic one.
d ∂L ∂L
= , (12.98)
dt ∂ ẋ ∂x
for this case yields
ẍ = −ω2 x , (12.99)
thus, indeed the term (mω 1 /2) xẋ doesn’t affect the dynamics. Requiring also
boundary conditions x (0) = xa and x (t) = xb leads to
and
t
dt′ xc ẋc
0
/t
ω dt′ (xb sin (ωt′ ) − xa sin (ω (t′ − t))) (xb cos (ωt′ ) − xa cos (ω (t′ − t)))
0
=
sin2 (ωt)
x2b − x2a
= ,
2
(12.102)
thus, the action is given by
mω ωt mω 1 x2b − x2a
= (xa − xb )2 cot (ωt) − 2xa xb tan + .
2 2 4
(12.103)
To evaluate the propagator according to Eq. (12.85) the factor K (t) has
to be determined. This can be done by employing relation (12.89) for the
case where xa = xb = 0
/ ′
dx Pc (0, t2 ; x′ ) Pc (x′ , t1 ; 0) K (t1 + t2 )
= . (12.104)
Pc (0, t1 + t2 ; 0) K (t1 ) K (t2 )
Exercise 12.4.4. Show that
0
mω
K (t) = . (12.105)
2πi sin (ωt)
Solution 12.4.4. By using Eqs. (12.103) and (12.104) one finds that
3 mω 4 K (t1 + t2 )
dx′ exp i (cot (ωt2 ) + cot (ωt1 )) x′2 = , (12.106)
2 K (t1 ) K (t2 )
thus, using the general integral identity
∞
0
iπ
dx′ exp iαx′2 = , (12.107)
−∞ α
where α is real, one finds that
;
2πi K (t1 + t2 )
= . (12.108)
mω (cot (ωt2 ) + cot (ωt1 )) K (t1 ) K (t2 )
where f (t) is an arbitrary function of time. Substituting this into Eq. (12.110)
yields
thus f (t) = At, where A is a constant. Combining all these results the prop-
agator (12.85) for the present case becomes
ωt 2xa xb
(xa − xb )2 cot (ωt) − 2xa xb tan = x2a + x2b cot (ωt) − .
2 sin (ωt)
(12.114)
d2 Sc mω
=− , (12.116)
dxa dxb sin (ωt)
thus, similar to the case of a free particle [see Eq. (12.77)], also for the present
case of a harmonic oscillator, the propagator can be expressed in terms of
the classical action Sc as
;
i d2 Sc i
K (xb , t; xa ) = exp Sc . (12.117)
2π dxa dxb
The proof of the above result, which requires generalization of the derivation
that led to Eq. (12.117) for the case of a general quadratic Lagrangian, will
not be given here. Another important result, which also is given here without
a proof, generalizes Eq. (12.118) for the case of motion in n spacial dimensions
;
i d2 Sc i
K (rb , t; ra ) = det exp Sc . (12.119)
2π dra drb
12.6 Problems
1. Show that
where x0 is a constant.
8. A particle having mass m is in the ground state of the potential well
V0 (x) = (1/2) mω2 x2 for times t < 0 . At time t = 0 the potential
suddenly changes and becomes V1 (x) = mgx.
a) Calculate the propagator K (xb , t; xa ) from point xa to point xb in
the semiclassical limit for the case where the potential is V1 (x) (i.e.
for the Hamiltonian after the change at t = 0).
b) Use
' the result
( ) of the previous section to calculate the variance
*
(∆x)2 (t) = x2 (t) − x (t) 2 of the position operator x at time t.
12.7 Solutions
Clearly, C (0) = 1. Moreover, with the help of Eq. (2.175) one finds that
dC
= −e−ǫA Aeǫ(A+B) e−ǫB + e−ǫA eǫ(A+B) (A + B) e−ǫB − e−ǫA eǫ(A+B) e−ǫB B ,
dǫ
(12.125)
thus
dC
= −A + (A + B) − B = 0 , (12.126)
dǫ ǫ=0
namely
C (ǫ) = 1 + O ǫ2 ,
and therefor
C (ǫ) = 1 + O ǫ3 , (12.132)
and therefor
eǫ(A+B) = eǫB/2 eǫA eǫB/2 + O ǫ3 . (12.133)
3. The proof is trivial using the identity
∞ 0
′2 ′ π β 2 /4α
e−αx +βx dx′ = e . (12.134)
−∞ α
4. By taking the time derivative of E one has
dE
= mṙ · r̈ + q∇ϕ·ṙ = ṙ· (mr̈ − qE) . (12.135)
dt
However, according to the equation of motion (1.60) the term in the
brackets vanishes, and therefor dE/dt = 0.
5. Assume that the energy eigenstates of the Hamiltonian H are labeled by
|an and the corresponding eigenenergies by En , i.e.
1= |an an | , (12.138)
n
iEn t
= xb |an exp − an |xa .
n
(12.139)
Z= e−βEn , (12.141)
n
Thus, with the help of Eq. (12.115) one finds for the case of a harmonic
oscillator that (recall that sin (ix) = i sinh x and cos (ix) = cosh x)
∞
Z= dx′ K (x′ , −i β; x′ )
−∞
0 ∞
mω mω [cosh (β ω) − 1] x′2
= dx′ exp −
2π sinh (β ω) −∞ sinh (β ω)
;
1
=
2 [cosh (β ω) − 1]
1
= ,
2 sinh βℏω
2
(12.147)
(12.148)
where
0
x0 = . (12.149)
mω
7. Denoting the state ket vector of the system by |ψ (t) and the time evo-
lution operator by u (t) one has
ψ (x′ , t) = x′ |ψ (t)
= x′ | u (t) |ψ (0)
∞
= dx′′ x′ | u (t) |x′′ x′′ |ψ (0)
−∞
∞
= dx′′ K (x′ , t; x′′ ) ψ (x′′ , 0) ,
−∞
(12.150)
′ ′′
where the propagator K (x , t; x ) is given by Eq. (12.73)
;
′ ′′ 1 i (x′ − x′′ )2
K (x , t; x ) = exp , (12.151)
2πiΩtx20 2Ωt x20
and where
Ω= , (12.152)
mx20
thus
;
∞
′ 1 1
ψ (x , t) = dx′′
π 1/4 x0
1/2 2πiΩtx20 −∞
2 2
1 i x′′ i x′ x′′ i x′
× exp − 1− − + .
2 Ωt x0 Ωt x20 2Ωt x0
(12.153)
ẍ = −g . (12.157)
gt2
x = x0 + v0 t − , (12.158)
2
where the constants x0 and v0 are the initial values of the position and
velocity at time t = 0. Given that x = xa at time t = 0 and x = xb at
time t one finds that x0 = xa and
xb − xa gt
v0 = + , (12.159)
t 2
thus the classical trajectory xc (t′ ) is given by
xb − xa g g
xc (t′ ) = xa + + t t′ − t′2 . (12.160)
t 2 2
gt2
xt = − , (12.161)
2
the trajectory xc (t′ ) is expressed as
t′ t′2
xc (t′ ) = xa + (xb − xa − xt ) + xt 2 , (12.162)
t t
and the corresponding velocity ẋc (t′ ) is expressed as
xb − xa − xt 2xt t′
ẋc (t′ ) = + 2 . (12.163)
t t
The Lagrangian along the classical trajectory is given by
1
L (xc , ẋc ) = mẋ2c − mgxc
2
2
xb −xa −xt 2xt t′
m t + t2 t′ t′2
= − mg xa + (xb − xa − xt ) + xt 2 ,
2 t t
(12.164)
and the corresponding action Sc is given by
d2 Sc m
=− , (12.167)
dxa dxb t
thus 0 ! "
2 x2
m im (xb − xa ) + 2xt (xb + xa ) − 3t
K (xb , t; xa ) = exp
2πi t 2t
0 ! 2 x2
"
1 1 i (xb − xa ) + 2xt (xb + xa ) − 3t
= exp
x0 2πiωt ωt 2x20
0 ! 2
"
1 1 i 83 x2t + x2a + 2 (2xt − xa ) (xb − xt ) + (xb − xt )
= exp ,
x0 2πiωt ωt 2x20
(12.168)
where
where
0
x0 = . (12.171)
mω
The wave function at time t is evaluated using the propagator
∞
ψ (x′ , t) = dx′′ K (x′ , t; x′′ ) ψ (x′′ , 0)
−∞
0
1 1 1
=
x0 2πiωt π1/4 x1/2
0
8 2 ′′ 2 ′′ ′ ′ 2
i 3 xt +(x ) +2(2xt −x )(x −xt )+(x −xt )
∞ 2
x′′
− 12
× dx′′ e ωt 2x2
0
x0
−∞
8 2 ′ ′ 2
0 i 3 xt +4xt (x −xt )+(x −xt )
ωt 2x2
1 1 e 0
= 1/2
x0 2πiωt π1/4 x0
x′′ 2
∞ (1− ωt
i
)( x0 ) −ix′′ (x′ −xt )
′′ − 2 +
ωtx2
× dx e 0 .
−∞
(12.172)
With the help of the identity
∞ 0
1 2 1 4a
b2
√ exp −ax + bx dx = e , (12.173)
π a
−∞
d |α
i = H |α . (13.1)
dt
For any given value of the time t the Hamiltonian H (t) is assumed to have
a discrete spectrum
where n = 1, 2, · · · , the momentary eigenenergies En (t) are real, and the set
of momentary eigenvectors is assumed to be orthonormal
The phase factors β n (t) in the expansion (13.4) are chosen to be given by
are the so-called dynamical phases, and the other phase factors
Chapter 13. Adiabatic Approximation
t
γ n (t) = i dt′ n (t′ ) |ṅ (t′ ) (13.7)
are the so-called geometrical phases. As we will see below, choosing the phase
factor β n (t) to be given by Eq. (13.5) ensures that the coefficients an (t)
become constants in the adiabatic limit.
Exercise 13.1.1. Show that the term n (t′ ) |ṅ (t′ ) is pure imaginary.
Solution 13.1.1. Note that by taking the derivative with respect to t (de-
noted by upper-dot) of the normalization condition (13.3) one finds that
ṅ |m + n |ṁ = 0 , (13.8)
thus
∗
n |ṁ = − m |ṅ . (13.9)
The last result for the case n = m implies that n (t′ ) |ṅ (t′ ) is pure imagi-
nary, and consequently γ n (t) are pure real.
Substituting Eq. (13.4) into Eq. (13.1) leads to
i ȧn (t) eiβn (t) |n (t)
n
Em (t)
ȧm (t)+iβ̇ m (t) am (t)+ an (t) eiβn (t) e−iβ m (t) m (t) |ṅ (t) = am (t) .
n
i
(13.11)
m (t)| Ḣ |n (t)
m (t) |ṅ (t) = . (13.14)
En (t) − Em (t)
and the inner product with m (t)|, where m = n, yields the desired identity.
where Λ (t′ ) is arbitrary real function of time. The geometrical phase γ n (t),
which is given by Eq. (13.7)
t
γ n (t) = i dt′ n (t′ ) |ṅ (t′ ) , (13.17)
t0
is transferred to
In the adiabatic limit the terms m (t) |ṅ (t) are considered to be negligibly
small. As can be seen from Eq. (13.14), this limit corresponds to the case
where the rate of change in time of the Hamiltonian approaches zero. In
this limit the coefficients an (t) become constants [see Eq. (13.13)], and the
solution (13.4) thus becomes
In this case the Hilbert space is two dimensional and the Hamiltonian can be
represented by a 2 × 2 matrix, which is conveniently expressed as a combina-
tion of Pauli matrices
H=h
˙ 0I + h · σ , (13.20)
01 0 −i 1 0
σ1 = , σ2 = , σ3 = . (13.21)
10 i 0 0 −1
√
Using the notation h = H ĥ, where H = h · h, and where ĥ is a unit vector,
given in spherical coordinates by
The orthonormal eigenvectors are chosen to be given by [see Eqs. (6.221) and
(6.222)]
cos θ2 exp − iϕ2
|+ =
˙ , (13.24)
sin θ2 exp iϕ
2
− sin θ2 exp − iϕ
2
|− =
˙ , (13.25)
cos θ2 exp iϕ
2
H |± = (h0 ± H) |± . (13.26)
With the help of the last result one thus finds that
1 h 1
γ± = ∓ da · 3 = ∓2Ω , (13.36)
2 |h|
S
where Ω is the solid angle subtended by the close path h (t) as seen from the
origin. Due to the geometrical nature of the last result, these phase factors
were given the name geometrical phases.
Solution 13.5.1. By employing Eqs. (10.21) and (10.27) one finds that (re-
call that Hnn = 0)
/t /t
pnn (t) = 1 − dt′ dt′′ (n| H (t′ ) H (t′′ ) |n) . (13.44)
t0 t0
=
Inserting the identity operator 1 = m |m) (m| between H (t′ ) and H (t′′ )
∗
and recalling that Hmn = Hnm lead to
where
2
/t ′ ′
pmn (t) = dt Hnm (t ) . (13.46)
t0
As can be seen from Eq. (13.39), the matrix elements Hnm (t′ ) are pro-
portional to the oscillatory dynamical phase factors
t
i
Hmn ∝ exp (i (ξ n (t) − ξ m (t))) = exp − dt′ (En (t′ ) − Em (t′ )) .
(13.47)
In the adiabatic limit these terms rapidly oscillate and consequently the prob-
abilities pmn (t) are exponentially small. From the same reason, the dominant
contribution to the integral is expected to come from regions where the en-
ergy gap En (t′ )−Em (t′ ) is relatively small. Moreover, it is also expected that
the main contribution to the total ’survival’ probability pnn will come from
those states whose energy Em (t′ ) is close to En (t′ ). Having this is mind, we
study below the transition probability for the case of a two level system. As
we will see below, the main contribution indeed comes from the region near
the point where the energy gap obtains a minimum.
where Ω is a positive constant, γ is a real constant, and where the initial time
is taken to be t0 = −∞ and the final one is taken to be t = ∞. In spherical
coordinates h (t) is given by
h (t) = H (0, sin θ, cos θ) , (13.49)
where
.
H= Ω 1 + (γt)2 , (13.50)
cot θ = γt , (13.51)
and where ϕ = π/2. Thus, the energy gap 2H obtains a minimum at time
t = 0. As can be seen from Eqs. (13.24) and (13.25), for any curve lying on
a plane with a constant azimuthal angle ϕ, the following holds
* θ̇
+̇ = |− , (13.52)
2
and therefor
* θ̇
− +̇ = , (13.53)
2
and
* *
+ +̇ = − −̇ = 0 . (13.54)
For the present case one finds using Eq. (13.51) that
* 1 γ
− +̇ = − . (13.55)
2 1 + (γt)2
This together with Eqs. (13.39) and (13.46) leads to
2
∞
/ ′ ′ *
p−+ = dt′ ei(ξ+ (t )−ξ− (t )) − (t′ ) +̇ (t′ )
−∞
/ t′ . 2
/ γt′ √ 2
γ ∞
/ exp − 2iΩ
γ 0 dτ 1 + τ2
= dt′
2 −∞ 1 + (γt′ )2
1 , √ √ -2 2
1 ∞/ exp − iΩ 2
γ τ 1 + τ − ln −τ + 1 + τ
2
= dτ .
2 −∞ 1 + τ2
(13.56)
In the limit γ/Ω ≪ the phase oscillates rapidly and consequently p−+ → 0.
The stationary phase points zn in the complex plane are found from the
condition
d 1
0= sinh 2z + z = cosh 2z + 1 , (13.63)
dz 2
thus
1
zn = iπ n + , (13.64)
2
where n is integer. Note, however that the term 1/ cosh z has poles at the same
points. Using the Cauchy’s theorem the path of integration can be deformed
to pass close to the point z−1 = −iπ/2. Since the pole at z−1 is a simple one,
the principle value of the integral exists. To avoid passing through the pole at
z−1 a trajectory forming a half circle "above" the pole with radius ε is chosen
were ε → 0. This section gives the dominant contribution which is iπR, where
R is the residue at the pole. Thus the probability p−+ is approximately given
by
Ω
p−+ ≃ exp −π . (13.65)
γ
The last result can be used to obtain a validity condition for the adia-
batic approximation. In the adiabatic limit p−+ ≪ 1, and thus the condition
πΩ/γ ≫ 1 is required to ensure the validity of the approximation.
H = H0 + H1 . (13.66)
where ml and ωl are the mass and angular frequency of mode l, respectively.
The Hamiltonian of the fast subsystem H1 (x̄), which depends paramet-
rically on x̄, has a set of eigenvectors and corresponding eigenvalues for any
given value of x̄
where 1F is the identity operator on the Hilbert space of the fast subsystem.
The state of the entire system ψ (t) at time t is expanded at any point x̄
using the ’local’ basis {|n (x̄) }
dψ
i = Hψ , (13.72)
dt
leads to
one obtains
With the help of Eqs. (13.67) and (13.74) one finds that
m (x̄)| ∂H
∂xl |n (x̄)
1
Am,n;l = i . (13.82)
εn − εm
As can be seen from the above result (13.83), the adiabatic approximation
greatly simplifies the system’s equations of motion. The effect of the fast
subsystem on the dynamics of the slow one is taken into account by adding a
vector potential Am,m;l (x̄) and a scalar potential εm (x̄) to the Schrödinger
equation of the slow subsystem [compare with Eq. (4.217)]. However, both
potential terms depend on the state m that is being occupied by the fast
subsystem.
Exercise 13.6.1. Show that if m (x̄)| ∂/∂xl |m (x̄) is pure real then
Solution 13.6.1. Note that in general the diagonal elements Am,m;l are real
since pl is Hermitian [see Eq. (13.76)]. On the other hand, if m (x̄)| ∂/∂xl |m (x̄)
is pure real then Amm;l (x̄) is pure imaginary, thus for this case Amm;l (x̄) = 0.
13.7 Problems
p2 mω 20 x2
H (t′ ) = + + xf (t′ ) , (13.89)
2m 2
where p is the variable canonically conjugate to x, the force f (t′ ) is given
by
′2
exp − tτ 2
′
f (t ) = λ √ , (13.90)
πτ
and ω 0 , λ and τ are real constants. When τ is sufficiently large the prob-
lem can be treated using the adiabatic approximation. Expand the state
of the system |ψ (t) in a basis of momentary eigenvectors of the Hamil-
tonian H (t′ ) and derive the equations of motion for the coefficients in
13.8 Solutions
ϕ (t) ϕ (t0 )
γ n (t) → γ̃ n (t) = γ n (t) − + . (13.93)
2 2
2. The momentary eigenenergies of the system are given by
2 2 2
π n
En (t) = 2 , (13.94)
2ma20 1 − α sin2 (ω p t)
3 π3
2ωp α1/2 ≪ . (13.96)
2ma20
b) In general, the term n (t′ ) |ṅ (t′ ) is pure imaginary [see Eq. (13.9)].
On the other hand, the fact that the wavefunctions of one dimensional
bound states can be chosen to be real (see exercise 7 of chapter 4),
implies that n (t′ ) |ṅ (t′ ) is pure real. Thus n (t′ ) |ṅ (t′ ) = 0 and
therefore all geometrical phases vanish.
3. The Hamiltonian H0 ≡ limt→±∞ H (t) is given by
p2 mω 20 x2
H0 = + . (13.97)
2m 2
The eigenvectors |n and eigenenergies En,0 = ω 0 (n + 1/2) of H0 satisfy
the following relation
H0 |n = En,0 |n , (13.98)
f (t′ )
α (t′ ) = − , (13.101)
21/2 mω20 x0
and
f 2 (t′ )
En (t′ ) = En,0 − , (13.104)
2mω20
and
t′
γ n (t′ ) = i dt′′ n (t′′ ) |ṅ (t′′ ) . (13.107)
or
where
λ2 1 2 2
µ= e− 2 ω0 τ . (13.112)
2m ω 0
Note that the above result is identical to (10.69). Note also that the exact
result of this problem is given by [see Eq. (5.345)]
e−µ µn
pn0 = . (13.113)
n!
This chapter discusses the quantization of electromagnetic (EM) field for the
relatively simple case of a free space cavity.
∇·A =0 , (14.5)
and the scalar potential ϕ vanishes in the absence of sources (charge and
current). In this gauge both electric and magnetic fields E and B can be
expressed in terms of A only as [see Eqs. (1.41) and (1.42)]
1 ∂A
E=− , (14.6)
c ∂t
and
B=∇×A. (14.7)
Solution 14.1.1. The gauge condition (14.5) and Eqs. (14.6) and (14.7)
guarantee that Maxwell’s equations (14.2), (14.3), and (14.4) are satisfied
1 ∂ (∇ × A) 1 ∂B
∇×E=− =− , (14.9)
c ∂t c ∂t
1 ∂ (∇ · A)
∇·E=− =0, (14.10)
c ∂t
∇ · B = ∇ · (∇ × A) = 0 , (14.11)
where in the last equation the general vector identity ∇ · (∇ × A) = 0 has
been employed. Substituting Eqs. (14.6) and (14.7) into the only remaining
nontrivial equation, namely into Eq. (14.1), leads to
1 ∂2A
∇ × (∇ × A) = − . (14.12)
c2 ∂t2
Using the vector identity
∇ × (∇ × A) = ∇ (∇ · A) − ∇2 A , (14.13)
1 ∂2A
∇2 A = . (14.14)
c2 ∂t2
Exercise 14.1.2. Consider a solution having the form
∇2 u+κ2 u = 0 , (14.16)
and
d2 q
+ω 2κ q = 0 , (14.17)
dt2
where κ is a constant and where
ω κ = cκ . (14.18)
∇·u =0 . (14.19)
1 d2 q
q∇2 u = u . (14.20)
c2 dt2
Multiplying by an arbitrary unit vector n̂ leads to
∇2 u · n̂ 1 d2 q
= 2 . (14.21)
u · n̂ c q dt2
The left hand side of Eq. (14.21) is a function of r only while the right hand
side is a function of t only. Therefore, both should equal a constant, which is
denoted as −κ2 , thus
∇2 u+κ2 u = 0 , (14.22)
and
d2 q
+ω 2κ q = 0 , (14.23)
dt2
where
ω κ = cκ . (14.24)
where the set {un } forms a complete orthonormal basis spanning the vector
space of all solutions of Eq. (14.16) satisfying the proper boundary conditions
on the conductive walls having infinite conductivity.
Solution 14.1.3. Equation (14.16) should be solved with the boundary con-
ditions of a perfectly conductive surface. Namely, on the surface S enclosing
the cavity we have B · ŝ = 0 and E × ŝ = 0, where ŝ is a unit vector normal
to the surface. To satisfy the boundary condition for E we require that u be
normal to the surface, namely, u = ŝ (u · ŝ) on S. This condition guarantees
also that the boundary condition for B is satisfied. To see this we calculate
the integral of the normal component of B over some arbitrary portion S ′ of
S. Using Eq. (14.7) and Stoke’s’ theorem one finds that
(B · ŝ) dS = 0 . (14.27)
S′
u1 , u2 ≡ (u1 · u2 ) dV , (14.28)
V
where the integral is taken over the volume of the cavity. Using Eq. (14.16)
one finds that
Exercise 14.1.4. Show that the total electric energy in the cavity is given
by
1
UE = q̇n2 , (14.33)
8πc2 n
Solution 14.1.4. Using Eqs. (14.6),(14.7) and (14.25) one finds that the
fields are given by
1
E=− q̇n un , (14.35)
c n
and
B= qn ∇ × un . (14.36)
n
The total energy of the field is given by UE +UB , where UE (UB ) is the energy
associated with the electric (magnetic) field, namely,
1
UE = E2 dV , (14.37)
8π V
and
1
UB = B2 dV . (14.38)
8π V
1 q̇n2 ω2 q2
L = UE −UB = − n n , (14.45)
4πc2 n
2 2
where the symbol overdot is used for derivative with respect to time, and
where ω n = cκn . The Euler-Lagrange equations, given by
d ∂L ∂L
− =0, (14.46)
dt ∂ q̇n ∂qn
∂HF ω2
ṗn = − = − n2 qn , (14.50)
∂qn 4πc
lead also to Eq. (14.23). Note that, as expected, the following holds
HF = UE +UB , (14.51)
and
and
ω2n
ṗn = − qn . (14.56)
4πc2
It is useful to introduce the annihilation and creation operators
0
iωn t ωn 4πic2 pn
an = e qn + , (14.57)
8πc2 ωn
0
ωn 4πic2 pn
a†n = e−iωn t 2
qn − . (14.58)
8πc ωn
The phase factor eiωn t in the definition of an is added in order to make it
time independent. The inverse transformation is given by
;
2πc2
qn = e−iωn t an + eiωn t a†n , (14.59)
ωn
0
ωn
pn = i −e−iωn t an + eiωn t a†n . (14.60)
8πc2
The commutation relations for the these operators are derived directly from
Eqs. (14.52) and (14.53)
, -
an , a†m = δ n,m , (14.61)
, † †-
[an , am ] = an , am = 0 . (14.62)
The Hamiltonian (14.48) can be expressed using Eqs. (14.59) and (14.60) as
1
HF = ω n a†n an + . (14.63)
n
2
The eigenstates are the photon-number states |s1, s2 , ..., sn , ... , which satisfy
[see chapter 5]
1
HF |s1, s2 , · · · , sn , · · · = ωn sn + |s1, s2 , · · · , sn , · · · . (14.64)
n
2
Consider the case where the EM field is confined to a finite volume V , which
for simplicity is taken to have a cube shape with edge L = V −1/3 . The
eigen modes and eigen frequencies of the EM field are found in exercise 1
of this chapter for the case where the walls of the cavity are assumed to
have infinite conductance [see Eq. (14.100)]. It is however more convenient to
assume instead periodic boundary conditions, since the spatial dependence
of the resulting eigen modes [denoted by un (r)], can be expressed in terms of
exponential functions, rather than trigonometric functions [see Eqs. (14.93),
(14.94) and (14.95)]. For this case Eq. (14.63) becomes
1
HF = ω k a†k,λ ak,λ + , (14.68)
2
k,λ
where the eigen frequencies are given by ω k = c |k|. In the limit of large
volume the discrete sum over wave vectors k can be replaced by an integral
∞ ∞ ∞
V
→ dkx dky dkz , (14.70)
k (2π)3 −∞ −∞ −∞
the polarization vectors ǫ̂k,(±;n̂) are defined by [compare with Eqs. (6.221)
and (6.222)]
θ iϕ θ iϕ
ǫ̂k,(+;n̂) = cos e− 2 ǫ̂k,+ + sin e 2 ǫ̂k,− , (14.78)
2 2
θ − iϕ θ iϕ
ǫ̂k,(−;n̂) = − sin e 2 ǫ̂k,+ + cos e 2 ǫ̂k,− , (14.79)
2 2
and the following holds
ǫ̂∗k,(λ;n̂) · ǫ̂k,(λ′ ;n̂) = δ λ,λ′ , (14.80)
k , -
ǫ̂∗k,(λ;n̂) × ǫ̂k,(λ′ ;n̂) = i λ cos θδ λ,λ′ − sin θ 1 − δ λ,λ′ . (14.81)
|k|
The linear momentum PF and angular momentum MF of the field are
taken to be given by
E×B−B×E
PF = dV , (14.82)
8πc
and
A×E−E×A
MF = − dV . (14.83)
V 8πc
With the help of Eqs. (14.6), (14.7) and (14.69), the following general vector
identity
and
Note that for linear polarization ǫ̂∗k′ ,λ′ × ǫ̂k′ ,λ′ = 0, whereas for circular
polarization ǫ̂∗k′ ,λ′ × ǫ̂k′ ,λ′ = iλ′ k/ |k|, where λ′ ∈ {+, −} [see Eq. (14.81)].
14.4 Problems
1. Find the eigen modes and eigen frequencies of a cavity having a pizza
box shape with volume V = L2 d.
2. Casimir force - Consider two perfectly conducting metallic plates placed
in parallel to each other. The gap between the plates is d and the tem-
perature is assumed to be zero. Calculate the force per unit area acting
between the plates.
3. Find the average energy per unit volume of the electromagnetic field in
thermal equilibrium at temperature
' ( T.
2
4. Calculate the variance (∆U) in the energy of the electromagnetic
field in thermal equilibrium at temperature T .
5. Consider an electromagnetic cavity having a set of normal modes. The
waveform of mode n is denoted by un (r), the angular frequency by ω n ,
the annihilation operator by an , and the creation operator by a†n . The
electric filed operator at point r and time t can be expressed as [see Eqs.
(14.6) and (14.67)]
where
Consider the case where all modes in the cavity are in their ground state
except of a single mode, which is in a number state with m photons.
Calculate g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) for such a state.
6. quantum diffraction - Consider the case where sources located in the
left half space z < 0 generate a monochromatic electromagnetic field at
angular frequency ω 0 . The right half space z > 0 is assumed to be a
vacuum free of any matter and sources. Assume the paraxial case, for
which the characteristic angle between the direction of propagation of
the field and the z axis is assumed small. Express the vector potential
operator A (r, t) in the plane z = z ′ > 0 in terms of its value in the plane
z = 0.
7. two-photon states - Space inversion corresponds to the transformation
r → −r. Under space inversion a general quantum state vector |ψ is
transformed to the state P |ψ , where P is the so-called parity opera-
tor P [compare with Eq. (5.103)]. Consider the four two-photon states
|+, + , |+, − , |−, + and |−, − , where
where the operator a†kẑ,+ (a†kẑ,− ) creates a photon having wave vector kẑ
and right (left) handed circular polarization, the operator a†−kẑ,+ (a†−kẑ,− )
creates a photon having wave vector −kẑ and right (left) handed circu-
lar polarization, and |0 is the vacuum state. Construct an orthonormal
basis to the subspace spanned by the vectors |+, + , |+, − , |−, + and
|−, − made of eigenvectors of both the parity operator P and the angular
momentum operator MFz = MF · ẑ.
14.5 Solutions
u1 , u2 = (u1 · u2 ) dV
V
ux1 ux2 dV
V
8
= ax1 ax2
V
L
nx1 π nx2 π
× cos x cos x dx
0 L L
L
ny1 π ny2 π
× sin y sin y dy
0 L L
d
nz1 π nz2 π
× z sin
sin z dz ,
0 d d
8 L2 d
= ax1 ax2 δ nx1 ,nx2 δ ny1 ,ny2 δ nz1 ,nz2 .
V 8
(14.103)
Similar results are obtained for the contribution of the y and z compo-
nents. Thus
L
2
d ∞ ∞ ∞ .
− c dkx dky dkz kx2 + ky2 + kz2 .
π π 0 0 0
(14.105)
.
In polar coordinates u = kx2 + ky2 and θ = tan−1 (ky /kx ) one has
dkx dky = ududθ, thus
2 ′ ∞
0
L π πnz 2
U (d) = c du u u2 +
π 2 n 0 d
z
2 ∞ ∞
L dπ
− c du u dkz u2 + kz2 .
π π2 0 0
(14.106)
Changing the integration variables
2
ud
x= , (14.107)
π
kz d
Nz = , (14.108)
π
leads to
! ′
"
∞
π 2 cL2
U (d) = F (nz ) − dNz F (Nz )
4d3 nz 0
! ∞
"
∞
π 2 cL2 1
= F (0) + F (nz ) − dNz F (Nz ) ,
4d3 2 n 0
z =1
(14.109)
where the function F (ξ) is given by
∞ . ∞
√
F (ξ) = dx x + ξ 2 = dy y . (14.110)
0 ξ2
∞ ∞
1 = 1 ′ 1 ′′′
F (0) + F (n) − dN F (N) = − F (0) + F (0) + · · · ,
2 n=1 0 12 720
(14.112)
π2 cL2
U (d) = − . (14.114)
720d3
The force per unit area (pressure) P (d) is found by taking the derivative
with respect to d and by dividing by the area L2
π2 c
P (d) = − . (14.115)
240d4
The minus sign indicates that the force is attractive.
3. The average energy U in thermal equilibrium is given by Eq. (8.493),
which is given by
∂ log Zc
U =− , (14.116)
∂β
s1, s2 ,...=0
! ∞
"
#
−β ωn (sn + 12 )
= e
n sn =0
! "
# 1
= ,
n 2 sinh β ωn
2
(14.117)
where n labels the cavity modes. Using the last result one finds that
∂ log Zc
U =−
∂β
1
∂ log 2 sinh β ωn
=− 2
n
∂β
ωn β ωn
= coth .
n
2 2
(14.118)
It is easy to see that the above sum diverges since the number of modes
in the cavity is infinite. To obtain a finite result we evaluate below the
difference Ud = U (T ) − U (T = 0) between the energy at temperature
T and the energy at zero temperature, which is given by (recall that
coth (x) → 1 in the limit x → ∞)
ωn β ωn
Ud = coth −1
n
2 2
ωn
= .
n
eβ ωn −1
(14.119)
The angular frequencies ω n of the modes are given by Eq. (14.100). For
simplicity a cubical cavity having volume V = L3 is considered. For this
case Ud is given by (the factor of 2 is due to polarization degeneracy)
∞ ∞ ∞
αn
Ud = 2kB T . (14.120)
nx =0 ny =0 nz =0
eαn−1
.
where n = n2x + n2y + n2z , and where the dimensionless parameter α is
given by
πβ c
α= . (14.121)
L
In the limit where α ≪ 1 (macroscopic limit) the sum can be approxi-
mated by the integral
∞
4π αn
Ud = 2kB T dn n2
8 eαn−1
0
∞
πkB T x3 dx
= ,
α3 ex − 1
0
π4
15
(14.122)
thus the energy per unit volume is given by
Ud π2 (kB T )4
= . (14.123)
V 15 3 c3
' (
4. In general, the energy variance (∆U )2 in thermal equilibrium of a
system having Hamiltonian H can be expressed as [see Eqs. (8.8) and
(8.42)]
' ( ) *
(∆U)2 = H2 − H 2 = Tr ρH2 − (Tr (ρH))2 , (14.124)
5. When only a single mode in the cavity is excited the normalized coher-
ence function g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) becomes [see Eqs. (14.88) and
(14.89) and the definition of g (l) ]
' (
l
a† al
g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) = l
, (14.130)
a† a
dkz ω
= 2 , (14.136)
dω c kz
the wave vector k is given by k = (kx , ky , kz ), and the component kz is
given in terms of the integration variables kx , ky and ω by Eq. (14.134).
/′
The symbol in Eq. (14.135) represents integration over values of kx
and ky for which kz is real, i.e. kx2 + ky2 < ω 2 /c2 [see Eq. (14.134)].
The following commutation relations hold [see Eqs. (14.71), (14.72) and
(14.134)]
3 4
[ak,λ , ak′ ,λ ] = a†k,λ , a†k′ ,λ = 0 , (14.137)
and
3 4 δ (ω−ω′ )
ak,λ , a†k′ ,λ′ = δ λ,λ′ dkz δ (kx −kx′ ) δ ky −ky′ . (14.138)
dω
By applying the Fourier transform to Eq. (14.141) one finds that the
following holds [recall the identity (4.47)]
F −1 A(−) (kx , ky , z, ω)
∞ ∞ ∞
1
= 3/2
dkx dky dω A(−) (kx , ky , z, ω) ei(kx x+ky y+ωt) ,
(2π)
−∞ −∞ −∞
(14.144)
satisfies the following relation [see Eq. (4.47)]
∞ ∞ ∞
(−) ′ ′ ′ ′ ′′ ′′ ∂G (r′ − r′′ , t′ − t′′ )
A (x , y , z , t ) = dx dy dt′′ A(−) (x′′ , y ′′ , 0, t′′ ) ,
∂z ′′
−∞ −∞ −∞
(14.146)
For the case of a monochromatic field the operators A(−) (x′′ , y ′′ , 0, t′′ )
and A(−) (x′ , y ′ , z ′ , t′ ) are expressed as
′′
A(−) (x′′ , y′′ , 0, t′′ ) = A(−) (x′′ , y ′′ , 0) e−ω0 t , (14.148)
(−) ′ ′ ′ ′ (−) ′ ′ ′ −ω0 t′
A (x , y , z , t ) = A (x , y , z ) e . (14.149)
Substituting into Eq. (14.146) yields the following relation between the
time independent operators A(−) (x′′ , y′′ , 0) and A(−) (x′ , y ′ , z ′ ) [see Eq.
(4.47)]
∞ ∞
(−) ′ ′ ′ ∂g (r′ − r′′ )
A (x , y , z ) = 2 dx′′ dy ′′ A(−) (x′′ , y ′′ , 0) ,
∂z ′′
−∞ −∞
(14.150)
With the help of the so-called Weyl’s plane waves expansion the function
g (r) can be expressed as
eikr
g (r) = − , (14.152)
4πr
where r = x2 + y 2 + z 2 . The above result (14.150) is known as the
Rayleigh-Sommerfeld first diffraction integral.
7. The parity operator reverses the direction of propagation (i.e. direction
of the wave vector). On the other hand the vector ǫ̂∗k′ ,λ′ × ǫ̂k′ ,λ′ remains
unchanged under space inversion, and therefore λ′ changes sign under
this transformation, and thus the following holds
P |+, + = |−, − , (14.153)
P |+, − = |+, − , (14.154)
P |−, + = |−, + , (14.155)
P |−, − = |+, + . (14.156)
As can be seen from Eq. (14.86), the following holds
In this chapter the transitions between atomic states that result from inter-
action with an electromagnetic (EM) field are discussed.
15.1 Hamiltonian
Consider an atom in an EM field. The classical Hamiltonian HF of the EM
field is given by Eq. (14.48). For the case of hydrogen, and in the absence
of EM field, the Hamiltonian of the atom is given by Eq. (7.2). In general,
the classical Hamiltonian of a point particle having charge e and mass me
in an EM field having scalar potential ϕ and vector potential A is given by
Eq. (1.62). In the Coulomb gauge the vector potential A is chosen such that
∇ · A = 0, and the scalar potential ϕ vanishes provided that no sources
(charge and current) are present. The EM field is assumed to be sufficiently
small to allows employing the following approximation
e 2 e
p− A ≃ p2 −2 A · p , (15.1)
c c
where p is the momentum vector. Recall that in the Coulomb gauge the vector
operators p and A satisfy the relation p · A = A · p, as can be seen from
Eqs. (6.171) and (6.344). These results and approximation allow expressing
the Hamiltonian of the system as
H = H0 + HF + Hp , (15.2)
where H0 is the Hamiltonian of the atom in the absence of EM field, and
where Hp , which is given by
e
Hp = − A·p , (15.3)
me c
is the coupling Hamiltonian between the atom and the EM field.
In the quantum case The Hamiltonian HF of the EM field is given by Eq.
(14.68)
1
HF = ω k a†k,λ ak,λ + , (15.4)
2
k,λ
Chapter 15. Light Matter Interaction
H0 |{sk,λ } , η = Eη |{sk,λ } , η ,
1
HF |{sk,λ } , η = ω k sk,λ + |{sk,λ } , η . (15.6)
2
k,λ
where ω i,f = Eηi − Eηf / . With the help of Eqs. (14.69) and (15.3) wi,f
can be rewritten as
2
e 4π2 c2 2
wi,f = δ (ω k − ω i,f ) f| ǫ̂∗k,λ · pe−ik·r a†k,λ |i . (15.8)
me c ωk V
4π2 e2 ωk 2
wi,f = δ (ω k − ω i,f ) f| ǫ̂∗k,λ · re−ik·r a†k,λ |i
V
4π2 e2 ωk
= δ (ω k − ω i,f ) |Mi,f |2 ,
V
(15.10)
where the atomic matrix element Mi,f is given by
(a) 4π 2 e2 ω k sk,λ
w(i,sk,λ )→(f,sk,λ −1),λ = δ (ω k + ω i,f ) |Mi,f |2 . (15.13)
V
Note that in this case it is assumed that ω i,f < 0.
The emission (15.12) and absorption (15.13) rates provide the contribu-
(e) (a)
tion of a single mode of the EM field. Let dΓ(i,s)→(f,s+1),λ /dΩ (dΓ(i,s)→(f,s−1),λ /dΩ)
be the total emitted (absorbed) rate in the infinitesimal solid angle dΩ hav-
ing polarization λ. For both cases s denotes the photon occupation number of
the initial state. To calculate these rates the contributions from all modes in
the EM field should be added. In the limit of large volume the discrete sum
over wave vectors k can be replaced by an integral according to Eq. (14.70).
By using the relation ω k = ck, where k = |k|, one finds that
(e)
dΓ(i,s)→(f,s+1),λ V ∞
(e)
= 3 dk k2 w(i,sk,λ )→(f,sk,λ +1),λ
dΩ (2π) 0
∞
e2 (s + 1) 2
= |M i,f | dx x3 δ (x − ω i,f )
2π c3 0
αfs (s + 1) ω3i,f 2
= |Mi,f | ,
2πc2
(15.14)
2
where αfs = e / c ≃ 1/137 is the fine-structure constant. In a similar way,
one finds for the case of absorption that
(a)
dΓ(i,s)→(f,s+1),λ αfs sω3i,f
= |Mi,f |2 . (15.15)
dΩ 2πc2
While the size of an atom aatom is on the order of the Bohr’s radius a0 =
0.53×10−10 m (7.64), the energy difference Eηi −Eηf is expected to be on the
order of the ionization energy of hydrogen atom EI = 13.6 eV (7.66). Using
the relation ωk = Eηi − Eηf / = ck one finds that aatom k ≃ 10−3 . Thus,
to a good approximation the term e−ik·r in the expression for the matrix
element Mi,f can be replaced by unity
Exercise 15.2.1. Show that the selection rule for the magnetic quantum
number m is given by
∆m = mf − mi ∈ {−1, 0, 1} . (15.20)
Therefore Mi,f = 0 [see Eq. (15.16)] unless ∆m ∈ {−1, 0, 1}. The transition
∆m = 0 is associated with linear polarization in the z direction, whereas
the transitions ∆m = ±1 are associated with clockwise and counterclockwise
circular polarizations respectively.
Exercise 15.2.2. Show that the selection rule for the quantum number l is
given by
∆l = lf − li ∈ {−1, 1} . (15.24)
Since both li and lf are non negative integers, and consequently li +lf +2 > 0,
one finds that kf , lf , mf , σf | r |ki , li , mi , σ i can be nonzero only when li =
lf = 0 or |∆l| = 1. However, for the first possibility, for which li = mi = lf =
mf = 0, the wavefunctions of both states |ki , li , mi , σi and |kf , lf , mf , σ f is a
function of the radial coordinate r only [see Eq. (6.130)], and consequently
kf , lf , mf , σ f | r |ki , li , mi , σi = 0. Therefore the selection rule is given by ∆l ∈
{−1, 1}.
∇ × (f V) = f ∇ × V+ (∇f) × V , (15.30)
ǫ̂ · ǫ̂∗ = 1 , (15.33)
and
k × ǫ̂ k × ǫ̂∗
· =1, (15.34)
|k| |k|
and thus
UF = ω |α|2 . (15.35)
Solution 15.3.2. With the help of Eqs. (14.6) and (14.7) one obtains [see
Eq. (14.84)]
1 ∂A
S=− × (∇ × A)
4π ∂t
c ω
= iǫ̂ei(k·r−ωt) α − iǫ̂∗ e−i(k·r−ωt) α∗
2V
k × ǫ̂ i(k·r−ωt) k × ǫ̂∗ −i(k·r−ωt) ∗
× i e α−i e α
|k| |k|
! 2"
c ω k ǫ̂ × (k × ǫ̂) iαei(k·r−ωt)
= |α|2 + Re .
V |k| |k|
(15.37)
The average Poynting vector over time S is given by [see Eq. (15.35)]
c ω |α|2 k cUF k
S = = . (15.38)
V |k| V |k|
ieωa
+| Hp |− = − +| A · r |−
c
= Ωe−iωt + Ω ∗ eiωt ,
2
(15.39)
where (it is assumed that ω ≃ ω a )
0
2πω a
Ω = −2iedp α, (15.40)
V
where
dp = ǫ̂ · +| r |− . (15.41)
ωa ω 1 e−i(ωt+θ1 ) + ei(ωt+θ1 )
H=
˙ i(ωt+θ 1 ) −i(ωt+θ1 ) . (15.44)
2 ω1 e +e −ω a
iωt iωt
|ψ (t) = b+ (t) exp − |+ + b− (t) exp |− . (15.45)
2 2
d b+ 1 ∆ω ω1 e−iθ1 + ei(2ωt+θ1 ) b+
i = iθ 1 −i(2ωt+θ1 ) ,
dt b− 2 ω1 e +e −∆ω b−
(15.46)
where
∆ω = ω a − ω . (15.47)
ω a ω1 e−iωt
H=
˙ , (15.48)
2 ω1 eiωt −ωa
and the equation of motion in the rotating frame can be taken to be given
by
d b+ 1 ∆ω ω1 b+
i = . (15.49)
dt b− 2 ω1 −∆ω b−
The time evolution is found using Eq. (6.138) [see also Eq. (6.268)]
b+ (t)
b− (t)
cos θ − i √∆ω
2
sin θ
2
−i √ ω21 sin θ 2 b+ (0)
ω1 +(∆ω) ω1 +(∆ω)
= ,
−i √ ω21 sin θ cos θ + i √∆ω sin θ b− (0)
ω1 +(∆ω)2 2ω1 +(∆ω)2
(15.50)
where
.
ω 21 + (∆ω)2 t
θ= . (15.51)
2
15.4 Problems
1. Show that
, 2 , 2 --
L , L ,r = 2 2
rL2 +L2 r . (15.52)
fn,n′ = 1 . (15.54)
n′
τ2
E (t) = E0 ẑ , (15.55)
τ 2 + t2
where τ is a constant having the dimension of time, is externally applied.
Calculate the probability p2p to find the atom in the sub-shell 2p at time
t → ∞.
5. Consider a particle having mass m and charge q moving in a one dimen-
sional harmonic oscillator having angular resonance frequency ω. Cal-
culate using the dipole approximation the rate of spontaneous emission
from the number state |n to the ground state |0 .
6. A hydrogen atom is initially in its ground state. An electric field given
by E0 cos (ωt), where both E0 and ω are constants, is externally applied.
Assume that ω > EI , where EI is the ionization energy of the atom.
Calculate the rate of ionization.
15.5 Solutions
me e4 1 1
ωi,f = − + , (15.60)
2 3 22 12
thus in terms of the Bohr’s radius a0 [see Eq. (7.64)] one has
2
(se) 33 α5fs me c2 Mi,f
Γi→f,λ = .
28 a0
3/2
1
R10 (r) = 2 e−r/a0 , (15.63)
a0
3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (15.64)
2a0
3/2
1 r − r
R21 (r) = √ e 2a0 , (15.65)
2a0 3a0
0
1
Y00 (θ, φ) = , (15.66)
4π
0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (15.67)
2 2π
0
0 1 3
Y1 (θ, φ) = cos θ , (15.68)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (15.69)
2 2π
where a0 is Bohr’s radius [see Eq. (7.64)], one finds for the radial part
that
∞ √
3 27 6
dr r R10 R21 = a0 , (15.70)
35
0
and
1 2π
∗
d (cos θ) dφ sin θe−iφ Y00 Y1−1 = 0 , (15.74)
−1 0
0 1 2π 0
1 −iφ ∗ 1
d (cos θ) dφ sin θe Y00 Y11 =− , (15.75)
2 3
−1 0
0 1 2π 0
1 ∗ 1
d (cos θ) dφ sin θe iφ
Y00 Y1−1 = , (15.76)
2 3
−1 0
1 2π
∗
d (cos θ) dφ sin θeiφ Y00 Y11 = 0 . (15.77)
−1 0
Thus, by combining all these results one finds that the inverse lifetime of
the states (2, 1, −1), (2, 1, 0) and (2, 1, 1) is given by
2
33 α5fs me c2 27 √ 1
Γ (se) = 2 = , (15.78)
28 35 1.06 × 10−9 s
whereas the lifetime of the state (2, 0, 0) is infinite (in the dipole approx-
imation).
4. The probability p2pm to find the atom in the state |n = 2, l = 1, m is
calculated using Eq. (10.42) together with Eq. (7.84)
2
e2 E02 τ 2 ∞
/ 3EI ′ τ
p2pm = 2
′ i
dt e 4 t | 2, 1, m| z |1, 0, 0 |2 . (15.79)
−∞ τ 2 + t′2
where
µe4
EI = (15.80)
2 2
is the ionization energy. The following holds
∞
/ 3EI ′ τ 1 ∞/ dxeix
dt′ ei 4 t =
−∞ τ2 +t′2 Ω −∞ 1 + x 2
Ω
∞
/ dxeix
=Ω ,
−∞ (x − iΩ) (x + iΩ)
(15.81)
where
3EI τ
Ω= , (15.82)
4
thus with the help of the residue theorem one finds that
(se) q 2 (nω)3 2
dΓ|n →|0 ,λ = | 0| z |n |2 ǫ̂∗k,λ · ẑ dΩ , (15.87)
2π c3
where ǫ̂∗k,λ is the polarization unit vector. With the help of Eqs. (5.11),
(5.28) and (5.29) one finds that
(se) q 2 (nω)3 n 2
dΓ|n →|0 ,λ = δ n,1 ǫ̂∗k,λ · ẑ dΩ . (15.88)
2π c3 2mω
Integrating over dΩ in spherical coordinates θ and φ with the help of the
relation
π 1
4π
dΩ cos2 θ = dφ d (cos θ) cos2 θ = , (15.89)
−π −1 3
and summing over the two orthogonal polarization yields the total rate
of spontaneous emission
(se) 2q 2 ω2
Γ|n →|0 = δ n,1 . (15.90)
3mc3
−3/2 16a40 k′ a0
= π1/2 eE0 a0 V −1/2 cos θ0 3 .
i (k′ a0 )2 + 1
(15.94)
′
The rate of ionization w is obtained by summing over k [see Eq. (10.39)]
2π
w= δ (∆Ek′ − ω) |Mk′ |2 , (15.95)
k′
where
2 ′2
k
∆Ek′ = + EI (15.96)
2me
is the change in the energy of the electron and where EI = me e4 /2 2 is
the ionization energy of the atom [see Eq. (7.66)]. Replacing the sum by
an integral according to (14.70) yields
thus
3
256e2 me E02 a40 (k0 a0 )
w= 6 , (15.98)
3 3
(k0 a0 )2 + 1
where
2me ( ω − EI )
k0 = . (15.99)
Note that for a given amplitude E0 the rate w obtains its maximum
value, which is given by [see Eq. (7.64)]
√
27 3 E02 a30
wmax = , (15.100)
16
when the angular frequency ω is chosen such that k0 a0 = 3−1/2 .
other words, when the particles are considered as indistinguishable the sub-
space corresponding to any given vector of occupation numbers n̄ is rather
than being gn̄ - fold degenerate (as in the approach where the particles are
considered to be distinguishable) is taken to be nondegenerate. The identical
particle postulate of quantum mechanics states that identical particles should
be considered as indistinguishable. Consequently, a basis for the Hilbert space
of the many-particle system can be constructed from the set of ket vectors
{|n̄ }n̄ . The ket vector |n̄ represents a state that is characterized by a vector
of occupation numbers n̄ = (n1 , n2 , · · · ), where the integer ni is the number
of particles that are in the single particle state |ai . Such a basis is considered
to be both orthonormal, i.e.
n̄1 |n̄2 = δ n̄1 ,n̄2 , (16.2)
where δ n̄1 ,n̄2 = 1 if n̄1 = n̄2 and δ n̄1 ,n̄2 = 0 otherwise, and complete
(a)
2 …
1 3
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0
1 …
3 2
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0
3 …
2 1
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0
(b)
…
=
Fig. 16.1. In this example the number of particles is N = i ni = 3, where the
occupation numbers are given by n̄ = (n1 , n2 , n3 , n4 , · · · ) = (0, 2, 1, 0, · · · ). When
the particles are considered as distinguishable
? [see panel (a)] the corresponding
subspace is gn̄ degenerate, where gn̄ = N!/ ni ! = 3. On the other hand, when
i
the particles are considered as indistinguishable [see panel (b)], the corresponding
subspace is nondegenerate.
3 4
where ai , a†j denotes anti-commutation, i.e.
+
Solution 16.1.1. For Bosons this result is trivial [see Eqs. (16.6) and (16.7)].
It is also trivial for Fermions when i = j. Finally, for Fermions when i = j
one has
Ni Nj = a†i ai a†j aj = −a†i a†j ai aj = a†i a†j aj ai = −a†j a†i aj ai = a†j aj a†i ai = Nj Ni .
(16.12)
16.2 Bosons
Based on Eqs. (16.2), (16.4), (16.6) and (16.7) a variety of results can be
obtained:
Solution 16.2.1. Trivial by Eq. (2.179), which states that for any operators
A and B
ai |0 = 0 . (16.15)
Solution 16.2.2. The norm of the vector ai |0 can be expressed with the
help of Eqs. (16.4) and (16.7)
3 4
0| a†i ai |0 = 0| a†i , ai + ai a†i |0
= − 0 |0 + 0, 0, · · · , ni = 1, 0, · · · |0, 0, · · · , ni = 1, 0, · · · ,
(16.16)
thus with the help of the normalization condition (16.2) one finds that
0| a†i ai |0 = 0 and therefore ai |0 = 0.
Solution 16.2.3. With the help of Eqs. (16.4), (16.13) and (16.15) one finds
that
16.3 Fermions
2
The anti-commutation relations (16.8) for the case i = j yields a†i = 0. As
can be seen from Eq. (16.4), this implies that the only possible occupation
numbers ni are 0 and 1. This result is known as the Pauli’s exclusion principle,
according to which no more than one Fermion can occupy a given single
particle state. For Fermions Eq. (16.4) can be written as (recall that 0! =
1! = 1)
n1 n2
|n̄ = a†1 a†2 · · · |0 , (16.21)
Solution 16.3.1. The norm of the vector ai |0 can be expressed with the
help of Eqs. (16.21) and (16.9)
3 4
0| a†i ai |0 = 0| a†i , ai − ai a†i |0
+
= 0 |0 − 0, 0, · · · , ni = 1, 0, · · · |0, 0, · · · , ni = 1, 0, · · · ,
(16.23)
thus with the help of the normalization condition (16.2) one finds that
0| a†i ai |0 = 0 and therefore ai |0 = 0.
where Ni = a†i ai .
Solution 16.3.2. Using Eqs. (16.8), (16.9) and (16.21) one finds that
n1 n2
Ni |n̄ = a†i ai a†1 a†2 · · · |0
=
2 nj n1 n2 ni
= (−1) j<i
a†1 a†2 · · · a†i ai a†i · · · |0
n1 n2 ni
= a†1 a†2 · · · a†i ai a†i · · · |0 .
(16.25)
For the case ni = 0 this yields [see
ni 3 Eq. 4 (16.22)] Ni |n̄ = 0, whereas for the
†
case ni = 1 one has ai ai = ai , ai −a†i ai = 1−a†i ai , thus Ni |n̄ = |n̄ .
†
+
Both cases are in agreement with Eq. (16.24).
Solution 16.3.3. According to Eq. (16.8) a†i a†j = −a†j a†i . For i = j this
2
yields a†i = 0. These relations together with Eq. (16.21) leads to Eq.
(16.27) (note that 1 − ni = 1 if ni = 0 and 1 − ni = 0 if ni = 1). Similarly,
ni
Eq. (16.26) is obtained by using the identity ai a†i = 1 − a†i ai and by
considering both possibilities ni = 0 and ni = 1.
|bj bj | = 1 . (16.30)
j
|ai ai | = 1 , (16.31)
i
The single particle state |ai can be expressed in the notation of many
particle states as a†i |0 , whereas the single particle state |bj can be expressed
as b†j |0 , where the operator b†j , which is the creation operator of the single
particle state |bj , is given by [see Eq. (16.32)]
The creation operator b†j is the Hermitian conjugate of the annihilation op-
erator
bj = bj |ai ai . (16.34)
i
where |r′ is a single particle position eigenvector, and where ψi (r′ ) = r′ |ai
is the wavefunction of the single particle state |ai .
Expressing the single particle state |ai in the notation of many particle
states as a†i |0 allows expressing the single particle state |r′ in the notation
of many particle states as Ψ † (r′ ) |0 [see Eq. (16.35)], where the operator
Ψ † (r′ ), which is given by
Note that while ψi (r′ ) is a wave function, Ψ (r′ ) is an operator on the Hilbert
space of the many particle system.
, -
Exercise 16.4.1. Calculate Ψ (r′ ) , Ψ † (r′′ ) ∓ , where [A, B]∓ = AB ∓ BA
for general operators A and B, and where the minus sign is used for Bosons
and the plus sign for Fermions.
Solution 16.4.1. With he help of Eqs. (16.7) and (16.9) one finds that
, - 3 4
Ψ (r′ ) , Ψ † (r′′ ) ∓ = ψi (r′ ) ψ∗i′ (r′′ ) ai , a†i′
∓
i,i′
= r′ |ai ai |r′′
i
= r′ |r′′ ,
(16.38)
thus [see Eq. (3.66)]
, -
Ψ (r′ ) , Ψ † (r′′ ) ∓ = δ (r′ − r′′ ) . (16.39)
d3 r′ ρ (r′ ) = N , (16.42)
where
and where
N= Ni (16.44)
i
The operator ρ (r′ ) is called the number density operator, and the operator
N is called the total number of particles operator.
= a†i ai
i
=N .
(16.45)
Consider a single particle observable such as the observable BSP , which was
introduced in the previous section [see Eqs. (16.28), (16.29), (16.30)]. It is
convenient to employ the single particle basis {|bj }j , which is made of single-
particle eigenvectors of BSP that satisfy BSP |bj = β j |bj [see Eq. (16.28)],
in order to construct creation b†j and annihilation bj operators. In the many-
particle case, the same physical variable that BSP represents for the single
particle case is represented by the operator B, which is given by
B= β j b†j bj . (16.46)
j
This can be seen by recalling that the operator b†j bj represents the number
of particles in the single particle state |bj and that β j is the corresponding
eigenvalue. With the help of Eqs. (16.28), (16.29), (16.30) and (16.33) (16.34)
the operator B can be expressed in terms of the operators a†i and ai
To see that the above expression indeed represents the two particle interaction
consider the expectation value n̄| V |n̄ with respect to the many body state
|n̄ = |n1, n2 , · · · . The following holds [see Eqs. (16.6) , (16.7), (16.8) and
(16.9)]
b†j b†j ′ bj ′ bj = ±b†j b†j′ bj bj ′
3 4
= ±b†j b†j′ , bj ± bj b†j′ bj′
∓
3 4
= ±b†j ∓ bj , b†j′ ± bj b†j ′ bj ′
∓
1
V = vj,j ′ Nj (Nj ′ − δ j,j′ ) . (16.52)
2
j,j ′
Separating the terms for which j = j ′ from the terms for which j = j ′ yields
1
V = vj,j′ Nj Nj ′ + vj,j Nj (Nj − 1) , (16.53)
2
j<j ′ j
(16.55)
thus
1
V = ai′ , ai′′ | VTP |ai′′′′ , ai′′′ a†i′ a†i′′ ai′′′ ai′′′′ . (16.56)
2 i′ ,i′′ ,i′′′ ,i′′′′
16.6 Hamiltonian
Consider the case where the single-particle Hamiltonian is given by
The matrix element ai′ | p2SP |ai can be written using the wavefunctions
ψi (r′ ) = r′ |ai [recall Eq. (3.29), according to which r′ | p |α = −i ∇ψα
for a general state |α ]
2
ai′ | p2SP |ai = d3 r′ (∇ψ∗i′ (r′ )) · (∇ψi (r′ )) . (16.60)
2m
Thus, in terms of the quantized field operator Ψ (r′ ) [see Eqs. (16.36) and
(16.37)] the operator T can be expressed as
2
T = d3 r′ ∇Ψ † (r′ ) · ∇Ψ (r′ ) . (16.61)
2m
Integration by parts yields an alternative expression
2
T =− d3 r′ Ψ † (r′ ) ∇2 Ψ (r′ ) . (16.62)
2m
Similarly, the many-particle potential energy operator is found using Eq.
(16.47) [recall Eq. (3.23), according to which r′ | f (r) |α = f (r′ ) ψα (r′ ) for
a general state |α and for a general function f (r)]
U = ai′ | USP (r′ ) |ai a†i′ ai
i,i′
1
V = d3 r′ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′ ) Ψ † (r′′ ) Ψ (r′′ ) Ψ (r′ ) . (16.65)
2
Combining all these results yields the total many-particle Hamiltonian
2
H= d3 r′ ∇Ψ † (r′ ) · ∇Ψ (r′ )
2m
+ d3 r′ USP (r′ ) Ψ † (r′ ) Ψ (r′ )
1
+ d3 r′ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′ ) Ψ † (r′′ ) Ψ (r′′ ) Ψ (r′ ) .
2
(16.66)
Exercise 16.6.1. Show that the Heisenberg equation of motion for the field
operator Ψ (r′ ) is given by
d
i Ψ (r′ , t)
dt
2
= − ∇2 + USP (r′ ) Ψ (r′ , t)
2m
+ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′′ , t) Ψ (r′′ , t) Ψ (r′ , t) .
(16.67)
Note that in the absence of two-particle interaction the above equation for
the field operator Ψ (r′ , t) is identical to the single-particle Schrödinger equa-
tion for the single particle wavefunction ψ (r′ ). Due to this similarity the
many-particle formalism of quantum mechanics is sometimes called second
quantization.
Solution 16.6.1. The Heisenberg equation of motion [see Eq. (4.37)] is given
by
dΨ
i = − [H, Ψ ]− . (16.68)
dt
For general operators A, B and C the following holds
[AB, C]− = A [B, C]± ∓ [A, C]± B
= A [B, C]± − [C, A]± B .
(16.69)
Below we employ this relation for evaluating commutation relations. For
Fermions the upper sign (anti-commutation) is chosen, whereas for Bosons
the lower one is chosen (commutation). With the help of Eqs. (16.39), (16.40)
and (16.41) one finds (for both Bosons and for Fermions) that
where
Uk′ −k′′ = k′′ | USP (r′ ) |k′
1 ′ ′′ ′
= d3 r′ USP (r′ ) ei(k −k )·r ,
V V
(16.80)
and the many-particle interaction operator is given by [see Eq. (16.56)]
1
V = k′ , k′′ | VTP |k′′′′ , k′′′ a†k′ a†k′′ ak′′′ ak′′′′ , (16.81)
2
k′ ,k′′ ,k′′′ ,k′′′′
where
1 ′′′′
−k′ )·r′ i(k′′′ −k′′ )·r′′
k′ , k′′ | VTP |k′′′′ , k′′′ = d3 r′ d3 r′′ VTP (r′ , r′′ ) ei(k e .
V2 V V
(16.82)
The assumption that VTP (r′ , r′′ ) is a function of the relative coordinate r =
r′ − r′′ only, together with the coordinates transformation
r′ + r′′
r0 = , (16.83)
2
r = r′ − r′′ , (16.84)
yields (note that r′ = r0 + r/2 and r′′ = r0 − r/2)
k′ , k′′ | VTP |k′′′′ , k′′′
1 ′′′′ ′ ′′′ ′′ i(k′′′′ −k′ −k′′′ +k′′ )·r
= 2 d3 r0 ei(k −k +k −k )·r0 d3 r VTP (r0 + r/2, r0 − r/2) e 2
V V V
1 i(k′′′′ −k′ −k′′′ +k′′ )·r
= δ k′ +k′′ ,k′′′ +k′′′′ d3 r vTP (r) e 2 ,
V V
(16.85)
where
Thus the only allowed processes for this case are those for which the total
momentum is conserved, i.e. k′ + k′′ = k′′′ + k′′′′ . Using the notation
where
1 iq·r
vq = d3 r vTP (r) e 2 . (16.89)
V V
16.8 Spin
where the quantum number σ indicates the spin state. The single-particle
orthonormality condition reads
In the momentum representation the single particle state |k′ , σ has a wave-
function given by [see Eq. (16.75)]
1 ′ ′
r′ |k′ , σ = √ eik ·r , (16.95)
V
and thus the quantized field operator Ψσ (r′ ) is given by [see Eq. (16.37)]
1 ′ ′
Ψσ (r′ ) = √ eik ·r ak′ ,σ . (16.96)
V k′
The Fermi wave vector is chosen such that the number of single particle states
for which |k′ | ≤ kF is N. Since the density of states per spin in k′ space is
V/8π3 one finds that
V 4 3
2 πk = N , (16.99)
8π3 3 F
thus
3π 2 N
kF3 = . (16.100)
V
The Fermi energy ǫF is the corresponding energy
2 2
kF
ǫF = . (16.101)
2m
The density of states D (ǫ) per spin and per unit volume is given by
1
D (ǫ) = δ (ǫ − ǫk′ ) . (16.102)
V
k′
where ǫk′ is given by Eq. (16.97). By replacing the sum by an integral one
finds that
2 ′2
1 k
D (ǫ) = δ ǫ−
V ′ 2m
k
∞
2 ′2
1 V k
= 4π dk′ k′2 δ ǫ −
V 8π3 2m
0
3/2 ∞ √
1 2m
= 2 2
dǫ′ ǫ′ δ (ǫ − ǫ′ )
4π
0
m √
= 2 3 2mǫ .
2π
(16.103)
3N 2 kF2
E0 = . (16.105)
5 2m
16.10 Problems
1. Find the many-particle interaction operator V for the case where the
two-particle potential is a constant VTP (r1 , r2 ) = V0 .
2. The same for the Coulomb interaction
e2
VTP (r1 , r2 ) = . (16.106)
|r1 − r2 |
3. Show that
dρ
+ ∇J = 0 , (16.107)
dt
where ρ (r′ ) = Ψ † (r′ ) Ψ (r′ ) is the number density operator [see Eq.
(16.43)] and where the current density operator J is given by
, -
J (r′ ) = Ψ † (r′ ) ∇Ψ (r′ ) − ∇Ψ † (r′ ) Ψ (r′ ) . (16.108)
2im
4. Consider two identical Bosons having mass m in a one dimensional po-
tential U (x) well given by
+
0 if 0 ≤ x ≤ L
U (x) = . (16.109)
∞ else
The particles interact with each other via a two-particle interaction given
by VTP = −V0 Lδ (x1 − x2 ), where V0 is a constant. Calculate the ground
state energy to lowest nonvanishing order in V0 .
5. By definition, an ideal gas is an ensemble of non-interacting identical
particles. The set of single particle eigenenergies is denoted by {εi }. Cal-
culate the average energy H and the average number of particles N in
thermal equilibrium as a function of the temperature T and the chemical
potential µ for the case of
a) Fermions.
b) Bosons.
where ak′ and a†k′ are Boson annihilation and creation operators corre-
sponding to the single particle state |k′ , and where ǫk′ and λ are real
coefficients.
8. Consider a system of identical spinless Bosons, whose Hamiltonian is
given by
ξ k′ † †
H= ǫk′ a†k′ ak′ + ak′ a−k′ + ak′ a−k′ , (16.113)
2
k′ k′
a) Find a condition that the function F (r′ , r′′ ) must satisfy in order to
ensure that the state |γ is normalized.
b) Consider the case where F (r′ , r′′ ) can be expressed as F (r′ , r′′ ) =
Af1 (r′ ) f2 (r′′ ), where A is a normalization constant (which is chosen
such that γ |γ = 1) and where both functions f1 () and f2 () are
normalized according to
2 2
1= dr′ |f1 (r′ )| = dr′ |f2 (r′ )| .
16.11 Solutions
1. In general V is given by Eq. (16.88) where for this case
vq = V0 δ q,0 , (16.118)
thus
V0
V = a†k′ a†k′′ ak′′ ak′
2
k′ ,k′′
V0 3 4
= a†k′ a†k′′ ak′′ , ak′ + ak′ a†k′′ ak′′ .
2 −
k′ ,k′′
(16.119)
With the help of Eq. (16.69) one finds that [see also Eqs. (16.6), (16.7),
(16.8) and (16.9)]
3 4 3 4
a†k′′ ak′′ , ak′ = a†k′′ [ak′′ , ak′ ]± − ak′ , a†k′′ ak′′
− ±
= −δ k′ ,k′′ ak′ ,
(16.120)
[for Fermions the upper sign (anti-commutation) is taken, whereas for
Bosons the lower one is taken (commutation)], thus
N (N − 1)
V = V0 , (16.121)
2
where N is the total number of particles operator. Note that N (N − 1) /2
is the number of interacting pairs in the system.
2. For this case the Fourier transform f (q) of the function 1/ |r| is needed
1
= d3 q f (q) eiq·r . (16.122)
|r|
Applying the Laplace operator ∇2 and using the identity
1
∇2 = −4πδ (r) (16.123)
|r|
yield
3. With the help of Eq. (16.67) and its Hermitian conjugate one finds that
dρ dΨ † (r′ ) dΨ (r′ )
= Ψ (r′ ) + Ψ † (r′ )
dt dt dt
1 2 , † ′ -
=− Ψ (r ) ∇ Ψ (r′ , t) − ∇2 Ψ † (r′ , t) Ψ (r′ ) ,
2
i 2m
(16.129)
∗
where the assumptions USP (r′ ) = USP (r′ ) and VTP
∗
(r′ , r′′ ) = VTP (r′ , r′′ )
have been made, thus
dρ
+ ∇J = 0 . (16.130)
dt
Note the similarity between this result and the continuity equation that
is satisfied by a single-particle wavefunction [see Eq. (4.73)].
4. For the unperturbed case, i.e. when V0 = 0, the single-particle wavefunc-
tions of the normalized eigenstates are given by
0
2 jπx
ψj (x) = sin , (16.131)
L L
where j = 1, 2, · · · , and the corresponding single-particle eigenenergies
are
2 2 2
π j
εj = . (16.132)
2mL2
For this case the ground state is the many-particle state |GS = |n1 = 2, n2 = 0, n3 = 0, · · · ,
i.e. the state for which both particles are in the j = 1 single-particle state.
In perturbation theory to first order in V0 the energy of this state is given
by [see Eq. (9.32)]
L L
1, 1| VTP |1, 1 = dx1 dx2 ψ1 (x1 ) ψ1 (x2 ) VTP (x1 , x2 ) ψ1 (x1 ) ψ1 (x2 )
0 0
L
= −V0 L dx1 ψ41 (x1 )
0
3
= − V0 ,
2
(16.135)
thus
2 2
π 3
E= − V0 + O V02 . (16.136)
mL2 2
5. The grandcanonical partition function [see Eq. (8.495)] is evaluated by
summing over all many-particle states
Zgc = Tr e−βH+βµN
= n1, n2 , · · · , ni , · · · | e−βH+βµN |n1, n2 , · · · , ni , · · · ,
n1, n2 ,···
(16.137)
where
H= εi a†i ai , (16.138)
i
N= a†i ai , (16.139)
i
and β = 1/kB T , thus one finds that
#
Zgc = e−βni (εi −µ) . (16.140)
i ni
and
! "
−βni (εi −µ)
log Zgc = log e . (16.141)
i ni
εi e−β(εi −µ)
= ,
i
1 + e−β(εi −µ)
(16.143)
and
H = εi fBE (εi ) ,
i
where
1
fBE (ε) = (16.150)
exp [β (ε − µ)] − 1
is the Bose-Einstein function .
6. The operators ak and a†k satisfy [see Eqs. (16.6), (16.7), (16.8) and (16.9)]
3 4
[ak′ , ak′′ ]± = a†k′ , a†k′′ =0, (16.151)
±
3 4
ak′ , a†k′′ = δ k′ ,k′′ . (16.152)
±
Using the definition (16.110) together with Eqs. (16.6) and (16.8) these
conditions become
3 4 3 4
vk′ uk′′ a†−k′ , ak′′ + uk′ vk′′ ak′ , a†−k′′ =0, (16.155)
± ±
3 4 3 4
vk′ uk′′ a−k′ , a†k′′ + uk′ vk′′ a†k′ , a−k′′ =0, (16.156)
± ±
3 4 3 4
uk′ uk′′ ak′ , a†k′′ + vk′ vk′′ a†−k′ , a−k′′ = δ k′ ,k′′ . (16.157)
± ±
Note that by inverting the transformation between the operators ak , a−k ,
a†k and a†−k and the operators bk , b−k , b†k and b†−k , which can be expressed
in matrix form as [see Eq. (16.110)]
bk uk 0 0 v k ak
b−k 0 u−k v−k 0 a−k
† =
bk 0 vk uk 0 a†k , (16.158)
b†−k v−k 0 0 u−k a†−k
one finds that
ak u−k 0 0 −vk bk
a−k
b−k
1 0 uk −v−k 0
† = † .
ak uk u−k − vk v−k 0 −vk u−k 0 bk
a†−k −v−k 0 0 uk b†−k
(16.159)
This result together with Eq. (16.111) imply that the expectation value
Vb | a†k ak |Vb is given by
2
vk
Vb | a†k ak |Vb = Vb | b−k b†−k |Vb , (16.160)
uk u−k − vk v−k
thus for both Bosons and Fermions [see Eq. (16.154)]
2
vk
Vb | a†k ak |Vb = . (16.161)
uk u−k − vk v−k
a) For the case of Fermions one finds using Eq. (16.9) that the condi-
tions (16.155), (16.156) and (16.157) become (recall that [A, B]+ =
[B, A]+ )
(vk′ uk′′ + uk′ vk′′ ) δ k′ ,−k′′ = 0 , (16.162)
(vk′ uk′′ + uk′ vk′′ ) δ k′ ,−k′′ = 0 , (16.163)
(uk′ uk′′ + vk′ vk′′ ) δ k′ ,k′′ = δ k′ ,k′′ , (16.164)
thus
vk u−k + uk v−k = 0 , (16.165)
2 2
uk + vk = 1 . (16.166)
These conditions are guarantied to be satisfied provided uk and vk
are expressed using a single real parameter θ k as
uk = cos θk , vk = sin θk , (16.167)
u−k = cos θk , v−k = − sin θ k . (16.168)
For this case Eq. (16.159) becomes
ak cos θ k 0 0 − sin θk bk
a−k 0 cos θk sin θk 0 b−k
† = † , (16.169)
ak 0 − sin θ k cos θ k 0 bk
†
a−k sin θk 0 0 cos θk b†−k
b) For the case of Bosons one finds using Eq. (16.7) that the condi-
tions (16.155), (16.156) and (16.157) become (recall that [A, B]− =
− [B, A]− )
−vk u−k + uk v−k = 0 , (16.171)
2 2
uk − vk = 1 . (16.172)
These conditions are guarantied to be satisfied provided that uk and
vk are expressed using a single real parameter θk as
uk = cosh θk , vk = sinh θk , (16.173)
u−k = cosh θk , v−k = sinh θk . (16.174)
For this case Eq. (16.161) thus becomes
where
3 4
Hk′ = ǫk′ a†k′ ak′ + λ ak′ + a†k′ , (16.177)
and where
H̄ = U † HU , (16.184)
where
! "
U = exp − Lk′ , (16.185)
k′
which yields
where
uk′ = u−k′ = cosh θ k′ , (16.191)
vk′ = v−k′ = sinh θk′ , (16.192)
the identities
sinh (2θ k′ ) = 2 sinh θk′ cosh θ k′ , (16.193)
cosh (2θ k′ ) = sinh2 θk′ + cosh2 θk′ , (16.194)
cosh (2θk′ ) + 1
cosh2 θk′ = , (16.195)
2
cosh (2θk′ ) − 1
sinh2 θk′ = , (16.196)
2
3 4
and the commutation relation bk′ , b†k′ = 1, one finds that
(16.201)
and thus, the eigenenergies are given by
;
2
ξ k′
En̄ = ǫk′ 1 − nk′
ǫk′
k′
;
2
ǫk′ ξ k′
+ 1− − 1 ,
′
2 ǫk′
k
(16.202)
where the nonnegative integer nk′ is the number of so-called quasi parti-
cles in state k′ .
9. Consider the state |α (r′ ) , which is defined by
where α (r′ ) ∈ C and where the operator Dα(r′ ) is given by [see for com-
parison Eq. (5.36)]
/
dr′ (α(r′ )Ψ † (r′ )−α∗ (r′ )Ψ (r′ ))
Dα(r′ ) = e . (16.204)
For general operators A and B the following holds [see Eq. (2.180)]
1 1
eA+B = eA eB e− 2 [A,B] = eB eA e 2 [A,B] , (16.205)
provided that
2
= dr′ |α (r′ )| ,
(16.207)
thus [see for comparison Eq. (5.39)]
and thus |α (r′ ) is normalized. With the help of Eq. (16.208) together
with the relation Ψ (r) |0 = 0 one finds that
/ 2 /
1
dr′ |α(r′ )| dr′ α(r′ )Ψ † (r′ )
|α (r′ ) = e− 2 e |0 . (16.210)
/ ′ 2
/ 3 / 4
= e− 2 dr |α(r )| e dr α(r )Ψ (r ) Ψ (r) + Ψ (r) , e
′ ′ ′ † ′
1
dr′ α(r′ )Ψ † (r′ )
|0
/ 2 /
− 12 dr |α(r
′ ′
)| e ′
dr α(r )Ψ ′ † ′
(r ) |0 ,
= α (r) e
(16.213)
that is
The expectation value with respect to the number operator N [see Eqs.
(16.42) and (16.43)] is given by
= d3 r′ |α (r)|2 ,
Z 2 EI
En = − , (16.216)
n2
where [see Eq. (7.66)]
me e4
EI = , (16.217)
2 2
and where me is the electron’s mass. The position wavefunction ψ n,l,m (r)
of a single-electron energy eigenstates having orbital quantum numbers
n, l and m is given by [see Eq. (7.92)]
(Z)
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (16.218)
(Z)
where the radial wavefunction Rnl (r) is obtained by substituting e2
by Ze2 in the radial wave function Rnl (r) of hydrogen [see Eqs. (7.127),
(7.128) and (7.129)]. The ground state |Υ (when the Coulomb interaction
between the electrons is disregarded) is given by [see Eq. (16.21)]
where a†n,l,m,σ are creation operators and where |0 represents the state
where all occupation numbers are zero. The energy of the unperturbed
ground state is −2 × 22 EI = −8EI [see Eq. (16.216)]. The Coulomb in-
teraction between the electrons is described by the two-particle operator
[see Eq. (16.106)]
e2
VTP (r1 , r2 ) = . (16.220)
|r1 − r2 |
and where
2
a0 = (16.223)
me e2
is the Bohr’s radius [see Eq. (7.64)]. The integration over r2 is performed
in spherical coordinated, where the z axis is chosen in the direction of
the vector r1
∞ 1 2π
27 1 −
4r1 4r2 d (cos θ2 )
α= 2 5 dr1 e a0
dr2 r22 e− a0
dφ2
π a0 r1 + r22 − 2r1 r2 cos θ2
2
0 −1 0
∞
r1 ∞
7
2 1 4r
− a1 1
4r
− a2
4r
− a2
= 4π 4π dr1 r12 e 0 dr2 r22 e 0 + dr2 r2 e 0
π 2 a50 r1
0 0 r1
5
= ,
2
(16.224)
thus the ground state energy is −8EI + Υ | V |Υ = − (11/2) EI . Note
that the fact the energy correction Υ | V |Υ is comparable with the un-
perturbed value of −8EI suggests that the accuracy of the first order
perturbation approximation is relatively poor.
11. With the help of the commutation relations (16.39), (16.40) and (16.41)
one finds that
γ |γ = dr′ dr′′ dr′′′ dr′′′′ F (r′ , r′′ ) F ∗ (r′′′ , r′′′′ )
+ dr′ dr′′ dr′′′ dr′′′′ F (r′ , r′′ ) F ∗ (r′′′ , r′′′′ ) 0| Ψ (r′′′′ ) Ψ † (r′ ) Ψ (r′′′ ) Ψ † (r′′ ) |0
2
= dr′ dr′′ |F (r′ , r′′ )| + F (r′ , r′′ ) F ∗ (r′′ , r′ ) .
(16.225)
a) The condition is
2
1= dr′ dr′′ |F (r′ , r′′ )| + F (r′ , r′′ ) F ∗ (r′′ , r′ ) . (16.226)
= |A|2 1 + |γ 12 |2 ,
(16.227)
where
1
g (r′′′′′ ) = dr′ dr′′ dr′′′ dr′′′′ f1 (r′ ) f2 (r′′ ) f1∗ (r′′′ ) f2∗ (r′′′′ )
1 + |γ 12 |2
× 0| Ψ (r′′′′ ) Ψ (r′′′ ) Ψ † (r′′′′′ ) Ψ (r′′′′′ ) Ψ † (r′ ) Ψ † (r′′ ) |0
|f1 (r′′′′′ )|2 + |f2 (r′′′′′ )|2 + γ 12 f1∗ (r′′′′′ ) f2 (r′′′′′ ) + γ ∗12 f1 (r′′′′′ ) f2∗ (r′′′′′ )
= .
1 + |γ 12 |2
(16.229)
c) The number of particles is given by
where |ϕ0 is the ground state of the free electron gas [see Eq. (16.98)],
thus
1 ′ ′′
−r′ )
Cσ (r′ − r′′ ) = eik ·(r . (16.232)
V
|k′ |≤k F
thus
13. First consider the unperturbed case, where the electron-electron Coulomb
interaction is disregarded. The ground state
is given by Eq. (16.98), and its energy E0 = (3N/5) 2 kF2 /2m by Eq.
(16.105), where kF is the Fermi wave vector. To first order in perturbation
(1)
theory the energy of the ground state becomes EGS = E0 + ∆E, where
the energy shift ∆E due to electron-electron Coulomb interaction is given
by [see Eqs. (9.32), (16.65), (16.93) and (16.106)]
1
∆E = d3 r′ d3 r′′ VTP (r′ , r′′ ) ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 ,
2
σ′ ,σ ′′
(16.237)
where
e2
VTP (r′ , r′′ ) = . (16.238)
|r′ − r′′ |
With the help of the expansion (16.96) and the commutation relations
(16.91) and (16.92) one finds that
ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0
1 ′′ ′′′ ′′ ′ ′′′′ ′
= 2 ei(k −k )·r ei(k −k )·r ϕ0 | a†k′′′′ ,σ′ a†k′′′ ,σ′′ ak′′ ,σ′′ ak′ ,σ′ |ϕ0
V ′ ′′ ′′′ ′′′′
k ,k ,k ,k
1 ′′
−k′ )·r′′ i(k′ −k′′′′ )·r′
= − 2 δ σ′ ,σ′′ ei(k e ϕ0 | a†k′′′′ ,σ′ ak′′ ,σ′ |ϕ0
V
k′ ,k′′ ,k′′′′
1 ′′
−k′′′ )·r′′ i(k′ −k′′′′ )·r′
+ ei(k e ϕ0 | a†k′′′′ ,σ′ ak′ ,σ′ a†k′′′ ,σ′′ ak′′ ,σ′′ |ϕ0 .
V2
k′ ,k′′ ,k′′′ ,k′′′′
(16.239)
The only nonvanishing terms in the second line are those for which either
k′ = k′′′′ and k′′ = k′′′ or k′ = k′′′ and k′′ = k′′′′ . For the second case
the two possibilities σ′ = σ ′′ and σ ′ = σ′′ are separately considered
(16.241)
For N ≫ 1 the single summation terms are negligibly small
ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0
2
1 1 ′ ′ ′′
= 2 1 − 2 δ σ′ ,σ′′ eik ·(r −r ) ,
V V
|k′′ |,|k′′ |≤kF |k′ |≤kF
(16.242)
or [see Eqs. (16.232) and (16.235)]
(16.246)
or [see Eq. (16.100)]
) 2 * kB T
x = . (17.7)
mω20
However, as can be seen from Eq. (17.5), when F0 = 0 the steady state
solution is given by x (t) = 0, contradicting thus the equipartition theorem.
This can be fixes by introducing yet another term f (t) in the equation of
motion representing fluctuating force
The fluctuating
) * force has vanishing mean f (t) = 0, however its variance
is finite f 2 (t) > 0. In exercise 1 below the autocorrelation function of the
fluctuating force f (t) is found to be given by (17.181)
Similarly to the classical case, also in the quantum case unphysical be-
havior is obtained when damping is disregarded. This happens not only for
the above discussed example of a driven resonator. For example, recall that
for a general quantum system driven by a periodic perturbation the time
dependent perturbation theory predicts in the long time limit constant rates
of transition between states [e.g., see Eq. (10.38)]. Such a prediction can
yield correct steady state population of quantum states only when damping
is taken into account.
Damping and fluctuation in a quantum system can be taken into account
by introducing a thermal bath, which is assumed to be weakly coupled to the
system under study. Below this technique is demonstrated for two cases. In
the first one, the system under study (also referred to as the closed system)
is a mechanical resonator, and in the second one it is taken to be a two level
system. In both cases the open system is modeled by assuming that the closed
system is coupled to a thermal bath in thermal equilibrium.
p2 1
H0 = + mω20 x2
2m 2
1
= ω0 a† a + ,
2
(17.10)
where
0
mω0 ip
a= x+ , (17.11)
2 mω 0
0
mω0 ip
a† = x− , (17.12)
2 mω 0
and where
, †-
a, a = 1 . (17.13)
Ht = H0 + Hr + V , (17.16)
where λk are coupling constants. The bath operators satisfy regular harmonic
oscillator commutation relations
3 4 , - 3 4
[a, bk ] = a, b†k = a† , bk = a† , b†k = 0 , (17.19)
The states of the thermal bath are assumed to be very dense, thus one can
replace the sum over k with an integral
2
|λk | exp [−iω k (t − t′ )]
k
∞
2
≃ dΩ |λ (Ω)| exp [−iΩ (t − t′ )] ,
−∞
(17.31)
2πδ(t−t′ )
2
= π |λ (ω 0 )| a (t) .
(17.32)
γ = π |λ (ω 0 )|2 , (17.33)
one has
The fluctuation terms F (t) and F † (t) represent noisy force acting on the
resonator.
From Eqs. (17.34), (17.35), (17.14), and (17.15) one finds that
ṗ + γp + mω 20 x = f (t) , (17.38)
where
where
1
n̂0 = , (17.44)
eβ ω0 −1
and where β = 1/kB T .
Solution 17.2.2. The modes of the thermal bath are assumed to be in ther-
mal equilibrium. In general, thermal averaging of an operator Ok , associated
with mode #k in the thermal bath, is given by [see Eqs. (8.8) and (8.42)]
Ok = Tr (ρk Ok ) , (17.45)
Z = Tr e−βHr,k , (17.47)
1
Hr,k = ω k b†k bk + , (17.48)
2
and β = 1/kB T . Using these expressions one finds that [see Eq. (8.170)]
' ( 1
b†k (t) bk (t) = β ωk ≡ n̂k . (17.49)
e −1
Moreover, using the full bath Hamiltonian Hr one can easily show that
' (
bk bl = b†k b†l = 0 , (17.52)
' (
b†k (t) bl (t) = δ kl n̂k , (17.53)
and
' (
bk (t) b†l (t) = δ kl (n̂k + 1) . (17.54)
The fluctuating forces are given by Eqs. (17.36) and (17.37). We calculate
below some correlation functions of these forces. Using Eq. (17.51) one finds
) *
F (t) = F † (t) = 0 . (17.55)
Replacing the sum over k with an integral, as in Eq. (17.31), and taking into
account only modes that are nearly resonant with the cavity mode one finds
) † *
F (t) F (t + t′ ) = 2γ n̂0 δ (t′ ) , (17.57)
where
1
n̂0 = . (17.58)
eβ ω0 −1
Similarly
) *
F (t) F † (t + t′ ) = 2γ (n̂0 + 1) δ (t′ ) , (17.59)
and
) *
F (t) F (t + t′ ) = F † (t) F † (t + t′ ) = 0 . (17.60)
) † *
Exercise 17.2.3. Show that the expectation value a a in steady state is
given by
) † *
a a = n̂0 . (17.61)
Solution 17.2.3. Multiplying Eq. (17.34) by the integration factor e(iω0 +γ)t
yields
d
ae(iω0 +γ)t = F (t) e(iω0 +γ)t . (17.62)
dt
The solution is given by
t
′
a (t) = a (t0 ) e(iω0 +γ)(t0 −t) + dt′ F (t′ ) e(iω0 +γ)(t −t) . (17.63)
t0
= n̂0 1 − e−2γ(t−t0 ) .
(17.66)
The assumption γ (t − t0 ) ≫ 1 allows writing this result as
) † *
a a = n̂0 . (17.67)
) † *
The last result a a = n̂0 verifies that the resonator reached thermal
equilibrium in steady state. Similarly, the next exercise shows that in the
classical limit the equipartition theorem of classical statistical mechanics is
satisfied.
) *
Exercise 17.2.4. Calculate x2 in steady state.
, -
Solution 17.2.4. According to Eq. (17.14) and a, a† = 1 the following
holds
) 2* ) *
x = a† + a a† + a
2mω0
) *
= a†2 + a2 + a† a + aa†
2mω0
) *
= a†2 + a2 + 2a† a + 1 .
2mω0
(17.68)
) †2 * ) 2 *
As can be seen from Eq. (17.60), a = a = 0. Thus, with the help of
Eq. (17.67) one has
) 2*
x = (2n̂0 + 1)
2mω0
β ω0
= coth ,
2mω0 2
(17.69)
in agreement with Eq. (8.178). In the classical limit where kB T ≫ ω0 one
has
) 2 * kB T
x = , (17.70)
mω20
in agreement with the classical equipartition theorem.
Hq =
˙ Ω (t) · σ , (17.71)
2
where Ω (t) is a 3D real vector, and where the components of the Pauli matrix
vector σ are given by
01 0 −i 1 0
σx = , σy = , σz = . (17.72)
10 i 0 0 −1
Let P = σ be the vector of expectation values P = ( σ x , σy , σz ). We
refer to this vector as the polarization vector. With the help of Eq. (4.38),
which is given by
dP
= Ω (t) × P . (17.75)
dt
The time varying ’effective magnetic field’ Ω (t) is taken to be given by
While ω 0 , which is related to the energy gap ∆ separating the TLS states by
ω 0 = ∆/ , is assumed to be stationary, the vector ω1 (t) is allowed to vary
in time, however, it is assumed that |ω 1 (t)| ≪ ω0 .
As we did in the previous section, damping is taken into account using a model
containing reservoirs having dense spectrum of oscillator modes interacting
with the TLS. Furthermore, since the ensembles are assumed to be dense,
summation over modes is done with continuos integrals. The Hamiltonian H
of the entire system is taken to be given by
H = Hq
01 00
σ+ = , σ− = , (17.78)
00 10
where i = 1, 2. While the coupling to the first bath (with coupling constant
Γ1 ) gives rise to TLS decay through spin flips, the coupling to the second
bath (with coupling constant Γϕ ) gives rise to pure dephasing.
and
dσ+ 1 Γ1
= [σ+ , Hq ] − + Γϕ σ +
dt i 2
i3 † 4
+ −V1 σz + 2σ + Vϕ + Vϕ† ,
(17.82)
where
0
Γ1 iφ1
V1 = e dωe−iω(t−t0 ) a1 (t0 , ω) , (17.83)
2π
and
0
Γϕ iφ2
Vϕ = e dωe−iω(t−t0 ) a2 (t0 , ω) . (17.84)
4π
Solution 17.3.1. With the help of the identities
[σz , σ + ] = 2σ + , (17.85)
[σz , σ − ] = −2σ− , (17.86)
[σ+ , σ − ] = σz , (17.87)
one finds that the Heisenberg equation of motion (4.38) for σ z is given by
dσz 1
= [σz , Hq ]
dt i0
Γ1
− 2i dω eiφ1 σ + a1 (ω)
2π
0
Γ1
+ 2i dω e−iφ1 a†1 (ω) σ − ,
2π
(17.88)
for σ + by
dσ+ 1
= [σ+ , Hq ]
dt i0
Γ1
−i dω e−iφ1 a†1 (ω) σ z
2π
0
Γϕ
+ 2i dω eiφ2 σ + a2 (ω)
4π
0
Γϕ
+ 2i dω e−iφ2 a†2 (ω) σ+ ,
4π
(17.89)
for a1 (ω) by
0
da1 (ω) Γ1 −iφ1
= −iωa1 (ω) − i e σ− , (17.90)
dt 2π
and for a2 (ω) by
0
da2 (ω) Γϕ −iφ2
= −iωa2 (ω) − i e σz . (17.91)
dt 4π
Integrating the equations of motion for the bath operators a1 (ω) and a2 (ω)
yields
a1 (ω) = e−iω(t−t0 ) a1 (t0 , ω)
0
Γ1 −iφ1 t ′ −iω(t−t′ )
−i e dt e σ− (t′ ) ,
2π t0
(17.92)
and
a2 (ω) = e−iω(t−t0 ) a2 (t0 , ω)
0
Γϕ −iφ2 t ′ −iω(t−t′ )
−i e dt e σ z (t′ ) .
4π t0
(17.93)
We now substitute these results into the Eqs. (17.88) and (17.89) and make
use of the following relations
′
dω e−iω(t−t ) = 2πδ (t − t′ ) , (17.94)
t
1
δ (t − t′ ) f (t′ ) dt′ = sgn (t − t0 ) f (t) . (17.95)
t0 2
where sgn(x) is the sign function
+
+1 if x > 0
sgn (x) = , (17.96)
−1 if x < 0.
to obtain
dσz 1
= [σz , Hq ]
dt i0
Γ1
− 2i dωeiφ1 σ + e−iω(t−t0 ) a1 (t0 , ω)
2π
− Γ1 σ + σ−
0
Γ1
+ 2i dωe−iφ1 eiω(t−t0 ) a†1 (t0 , ω) σ−
2π
− Γ1 σ + σ− ,
(17.97)
and
dσ+ 1
= [σ+ , Hq ]
dt i0
Γ1
−i dωe−iφ1 eiω(t−t0 ) a†1 (t0 , ω) σz
2π
Γ1
+ σ + σz
20
Γϕ
+ 2i dωeiφ2 σ+ e−iω(t−t0 ) a2 (t0 , ω)
4π
Γϕ
+ σ+ σ z
20
Γϕ
+ 2i dωe−iφ2 eiω(t−t0 ) a†2 (t0 , ω) σ+
4π
Γϕ
− σz σ + .
2
(17.98)
1
σ+ σ − = (1 + σ z ) , (17.99)
2
1
σ− σ + = (1 − σ z ) , (17.100)
2
σz σ + = −σ + σz = σ + , (17.101)
and assuming the case where the dominant contribution to the TLS dynamics
comes from the bath modes near frequency ω0 (recall that ω 0 = ∆/ , where
∆ is the energy gap separating the TLS states), one finds that
' (
V1† (t′ ) V1 (t)
Γ1 ′ ′
' (
= 2 dω dω′ e−iω (t−t ) a†1 (t0 , ω) a1 (t0 , ω ′ )
2π
Γ1 ′
= 2 dωe−iω(t−t ) n (ω)
2π
≃ 2 Γ1 n̂0 δ (t − t′ ) ,
(17.106)
1
n̂0 = . (17.107)
eβ ω0 −1
Similarly
' (
V1 (t) V1† (t′ ) = 2
Γ1 (n̂0 + 1) δ (t − t′ ) , (17.108)
) † ′ * 2 Γϕ
Vϕ (t ) Vϕ (t) = n̂0 δ (t − t′ ) , (17.109)
2
) * Γϕ
Vϕ (t) Vϕ† (t′ ) = 2 (n̂0 + 1) δ (t − t′ ) , (17.110)
2
and
' (
V1 (t′ ) V1 (t) = V1† (t′ ) V1† (t)
) *
Vϕ (t′ ) Vϕ (t) = Vϕ† (t′ ) Vϕ† (t) = 0 .
(17.111)
t
1 i Γ1 ′
+ dt′ [σ + , Hq ] + −V1† σz + 2 σ + Vϕ + Vϕ† σ+ e( 2 +Γϕ )(t −t) .
0 i
(17.115)
In the second step these expressions for the TLS operators are substituted
into Eqs. (17.112) and (17.113). In this final step, correlations are disregarded
(e.g. the expectation value' of a term
( having ' the (form σ+ V1† V1 is evaluated
using the approximation σ+ V1† V1 ≃ σ + V1† V1 ). The expectation values
of bath operators are calculated with the help of the results of the previous
section. This approach yields the following results
1 t ′ ( Γ21 +Γϕ )(t′ −t) ' † ′ (
σ + V1 = dt e V1 (t ) V1 (t) σz (t′ )
i 0
i Γ1 n̂0
=− Pz ,
2
(17.116)
' ( i Γ n̂
1 0
V1† σ− = Pz , (17.117)
2
' (
V1† σz = −i Γ1 n̂0 P+ , (17.118)
and
) *
σ + Vϕ + Vϕ† σ + = i Γϕ n̂0 P+ , (17.119)
thus
Ṗz = (Ω (t) × P)z − Γ1 [1 + (2n̂0 + 1) Pz ] , (17.120)
and
Γ1
Ṗ+ = (Ω (t) × P)+ − + Γϕ (2n̂0 + 1) P+ . (17.121)
2
A similar equation can be obtained for Ṗ− , which together with Eq. (17.121)
can be written as
Γ1
Ṗx = (Ω (t) × P)x − + Γϕ (2n̂0 + 1) Px , (17.122)
2
Γ1
Ṗy = (Ω (t) × P)y − + Γϕ (2n̂0 + 1) Py . (17.123)
2
Consider the case where ω 1 (t) = 0, i.e. Ω (t) = ω 0 ẑ [see Eq. (17.76)]. For
this case Eqs. (17.120) and (17.121) become
Ṗz = −Γ1 [1 + (2n̂0 + 1) Pz ] , (17.124)
Γ1
Ṗ± = ±iω0 − + Γϕ (2n̂0 + 1) P± . (17.125)
2
In the long time limit the solution is given by P± (t → ∞) = 0 and
Pz (t → ∞) = Pz0 , where [see Eq. (17.58)]
1 β ω0
Pz0 = − = − tanh . (17.126)
2n̂0 + 1 2
Note that Eq. (17.126) is in agreement with the Boltzmann distribution law
of statistical mechanics, according to which in thermal equilibrium the prob-
ability to occupy a state having energy ǫ is proportional to exp (−βǫ) (recall
that Pz is the probability to occupy the upper state of the TLS minus the
probability to occupy the lower one). In terms of the decay times T1 and T2 ,
which are defined by
the equations of motion for the general case, which are known as optical
Bloch equations, are given by
Px
Ṗx = (Ω (t) × P)x − , (17.129)
T2
Py
Ṗy = (Ω (t) × P)y − , (17.130)
T2
Pz − Pz0
Ṗz = (Ω (t) × P)z − . (17.131)
T1
17.4 Problems
where χ′ (ω) and χ′′ (ω) are respectively the real and imaginary parts
of the magnetic susceptibility χ (ω) (i.e. χ (ω) = χ′ (ω) + iχ′′ (ω)).
Note that the term proportional to χ′ (ω) is ’in phase’ with respect
to the driving magnetic filed in the x direction [recall that Bx =
B1 cos (ωt)], whereas the second term, which is proportional to χ′′ (ω)
is ’out of phase’ [i.e. proportional to sin (ωt)] with respect to Bx .
Calculate χ (ω).
6. A dilute gas of hydrogen atoms at temperature T is illuminated by a laser
having intensity IL (in units of power per unit area), circular polarization
and an angular frequency ω that is tuned close to the transition angular
frequency ω 0 from the ground state |n = 1, 0 = 0, m = 0 to the excited
state |n = 2, l = 1, m = 1 . The atoms are characterized by longitudinal
T1 and transverse T2 relaxation times. Calculate the probability pe in
steady state to find an atom in the excited state.
7. The Unruh-Davies Effect - The correlation function C (r′ , t′ ) is de-
fined by
c2 aτ
x (τ) = cosh −1 . (17.136)
a c
The proper time τ is the time as being measured by a clock com-
moving with the observer, and it is related to the time t in the fixed
inertial frame by
c aτ
t= sinh . (17.137)
a c
Consider the case where the observer is moving in an electromagnetic
field at temperature T = 0. Show that the effective temperature
of the electromagnetic field as being measured by the accelerated
observer is given by
a
TUD = . (17.138)
2πkB c
8. two-mode squeezing - Consider a system whose Hamiltonian is given
by
H = H0 + Hp , (17.139)
where
3 4
H0 = ω 0 (1 + η) B1† B1 + (1 − η) B2† B2 , (17.140)
the annihilation Bn and creation Bn† operators satisfy the following com-
mutation relations
3 4
Bn′ , Bn† ′′ = δ n′ ,n′′ , (17.142)
3 4
[Bn′ , Bn′′ ] = Bn† ′ , Bn† ′′ = 0 , (17.143)
where n′ , n′′ ∈ {1, 2}, the real parameters ω 0 , η and φ are real, and ζ (t)
is a real function of time t.
a) Show that the time evolution of the state vector of the system |ψ is
given by
where
t
ξ = ω0 dt′ ζ (t′ ) . (17.146)
0
b) Show that
and
where |0, 0 is the ground state of H0 . Calculate ∆Xθ ∆Pθ with re-
spect to the state |ξ, φ , where the operators Xθ and Pθ are defined
by
P2 = i A2 − A†2 , (17.155)
and
B1 + B2
A1 = √ , (17.156)
2
B1 − B2
A2 = √ . (17.157)
2
g) Show that
∞ ∞
*
S (ξ, 0) = dX1′ dX2′ e−ξ X1′ , eξ X2′ X1′ , X2′ | , (17.161)
−∞ −∞
X1 ± X2
X± = √ . (17.162)
2
′ ′
Calculate the joint probability distribution Px X+ , X− to obtain
′ ′
the values X+ and X− in a measurement of X+ and X− , respectively,
when the system is in the state |ξ, 0 .
17.5 Solutions
1. In the absence of any externally applied driving force, i.e. when Fex = 0,
the classical equation of motion is given by (17.8)
where f (t) represents a random force acting on the resonator due to the
coupling with the thermal bath at temperature T . Bellow we consider
statistical properties of the fluctuating function x(t). However, since some
of the quantities we define may diverge, we consider a sampling of the
function x(t) in the finite time interval (−τ /2, τ /2), namely
+
x(t) −τ /2 < t < τ /2
xτ (t) = . (17.164)
0 else
2πδ(ω+ω′ )
∞
1
= lim dω xτ (ω)xτ (−ω) .
τ →∞ τ −∞
(17.168)
Since x(t) is real xτ (−ω) = x∗τ (ω). In terms of the power spectrum Sx (ω),
which is defined as
1
Sx (ω) = lim |xτ (ω)|2 , (17.169)
τ →∞ τ
one has
) 2* ∞
x = dω Sx (ω) . (17.170)
−∞
where
∞
1
f(t) = √ dωf (ω)e−iωt . (17.172)
2π −∞
Assuming that in the vicinity of ω0 , i.e. near the peak of the integrand,
the spectral density Sf (ω) is a smooth function on the scale of the width
of the peak γ, and also assuming that ω 0 ≫ γ, one approximately finds
that
∞ ∞
1 dω
dω Sx (ω) ≃ Sf (ω 0 ) 2 2
−∞ m2 (ωγ) + (ω 20 − ω 2 )
−∞
Sf (ω 0 ) ∞ dα
=
m2 ω 30 −∞ (αγ/ω 0 )2 + (1 − α2 )2
Sf (ω 0 ) ∞ dα
≃
m ω 0 −∞ (αγ/ω 0 )2 + 1
2 3
πω0 /γ
π
= 2 2 Sf (ω0 ) .
m γω 0
(17.175)
γkB T 1
Sx (ω) = . (17.177)
mπ (ωγ) + (ω 20 − ω2 )2
2
2πδ(ω+ω′ )
∞
1 ′
= lim dωe−iωt f(ω)f(−ω)
τ →∞ τ −∞
∞
′
= dωe−iωt Sf (ω) .
−∞
(17.180)
Using Eq. (17.176) and assuming as before that Sf (ω) is smooth function
near ω = ω 0 allow determining the coefficient C (t′ )
∞
mγkB T ′
C (t′ ) = dωe−iωt = 2mγkB T δ (t′ ) . (17.181)
π −∞
2πδ(t′ )
2. Using the definition (17.39) and Eqs. (17.57), (17.59) and (17.60) one has
m ω0
f (t) f (t + t′ ) = −
), 2 -, -*
× F † (t) − F (t) F † (t + t′ ) − F (t + t′ )
= m γω0 (2n̂0 + 1) δ (t′ )
+1 ′ eβ ω0
= m γω0 δ (t )
−1 eβ ω0
β ω0
= m γω0 coth δ (t′ ) .
2
(17.182)
(2n̂0 + 1) 1 − e−2γ(t−t0 )
g (τ ) = cos (ω 0 τ ) e−γτ
mω 2
coth β 2ω 1−e −2γ(t−t0 )
= cos (ω 0 τ ) e−γτ .
mω 2
(17.186)
In steady state, i.e. for γ (t − t0 ) ≫ 1, the autocorrelation function g (τ )
becomes
β ω
g (τ ) = coth cos (ω 0 τ ) e−γτ . (17.187)
2mω 2
4. The Bloch equation (17.131) for this case becomes
Pz − Pz0
Ṗz = −ΓT Pz − , (17.188)
T1
thus in steady state
Pz0
Pz = . (17.189)
1 + ΓT T1
Clearly, by symmetry, Px = Py = 0 in steady state.
5. The Hamiltonian of the closed system is given by
Hq =
˙ Ω (t) · σ , (17.190)
2
where
1 + T22 (ω − ω 0 )2
Pz = Pz0 . (17.208)
1 + T22 (ω − ω 0 )2 + ω21 T1 T2
iT2 ω 1 [1 + iT2 (ω − ω0 )]
PR+ = Pz0 , (17.209)
1 + T22 (ω − ω 0 )2 + ω21 T1 T2
thus
iT2 ω 1 [1 + iT2 (ω − ω 0 )]
P+ = 2 Pz0 e−iωt . (17.210)
1 + T22 (ω − ω0 ) + ω21 T1 T2
Px = P+ + P− , (17.212)
6. In the rotating frame the Bloch equations are given by Eqs. (17.204) and
(17.205), where the Rabi frequency ω1 is given by Eq. (15.42). In steady
state, i.e. when ṖR+ = 0, Eq. (17.205) yields
−iω1 Pz
PR+ = 1 . (17.214)
i (ω − ω0 ) − T2
∗
The following holds (note that PR− = PR+ )
iω 1 (PR+ − PR− ) Pz
=− , (17.215)
2 T1L
−1
where T1L , which is given by
−1 ω21 T2
T1L = , (17.216)
1 + (ω0 − ω)2 T22
is the laser-induced transition rate, and thus Eq. (17.204) can be rewrit-
ten as
Pz Pz − Pz0 Pz − Pz0T
Ṗz = − − =− , (17.217)
T1L T1 T1T
−1
where T1T , which is given by
1 1 1
= + , (17.218)
T1T T1L T1
is the effective longitudinal decay rate, and Pz0T which is given by
T1T Pz0
Pz0T = , (17.219)
T1
1 + Pz0T 1 + T1TTP
1
z0
pe = = , (17.220)
2 2
−1
and thus pe ≃ (1 + Pz0 ) /2 when T1L ≪ T1−1 , and pe ≃ 1/2 in the
−1 −1
opposite limit when T1L ≫ T1 . The Rabi frequency ω 1 given by Eq.
(15.42) can be expressed as [note that the the laser intensity IL is the
magnitude of the time averaged Poynting vector S given by Eq. (15.38)]
0
2e |dp | 2π
ω1 = IL , (17.221)
c
where the matrix element dp is given by [see Eq. (15.62)]
x − iy
dp = n′ = 2, l′ = 1, m′ = 1| √ |n = 1, l = 0, m = 0
2
0 ∞ 1 2π
1 ∗
= dr r3 R21 R10 d (cos θ) dφ sin θe−iφ Y11 Y00
2
0 −1 0
15/2
2
=− a0 ,
35
(17.222)
where a0 is Bohr’s radius [see Eq. (7.64)], and thus [see Eq. (17.216)]
2 2
218 π e a0 IL
−1 310 2 c T2
T1L =
1 + (ω 0 − ω)2 T22
IL σλ
hc
= ,
1 + (ω 0 − ω)2 T22
(17.223)
where
218 παfs
σ= ω 0 T2 a20 = 0.101 × ω 0 T2 a20 , (17.224)
310
αfs = e2 / c ≃ 1/137 is the fine-structure constant, and λ = 2πc/ω 0 is
the laser wavelength.
7. With the help of Eq. (14.69), the commutation relations (14.71) and
(14.72), the relations
ωk = c |k| , (17.225)
ǫ̂∗k,λ · ǫ̂k,λ′ = δ λ,λ′ , (17.226)
ǫ̂k,λ · k = ǫ̂∗k,λ · k = 0 , (17.227)
and the thermal expectation values (17.51), (17.52), (17.53) and (17.54)
one finds that the correlation function (17.135) can be expressed as
2πc2 ′ ′
' ( ′ ′
' (
C (r′ , t′ ) = e−i(k·r −ωk t ) ak,λ a†k,λ + ei(k·r −ωk t ) a†k,λ ak,λ
ωk V
k,λ
and thus
! " 2
′ 2 π 2 1 − coth2 π
K 2 πkB T
C (0, t ) = =− ′ .
πct′2 K2 πc sinh2 πkB T t
(17.233)
(17.234)
thus
2 c 1
C (r′ , t′ ) = . (17.235)
π (r ) − (ct′ )2
′ 2
c) With the help of Eqs. (17.136), (17.137) and (17.235) one finds that
the value of the correlation function C (r′ , t′ ) as being measured by
the accelerated observer is given by
C (x (τ 2 ) − x (τ 1 ) , t (τ 2 ) − t (τ 1 ))
2 a2 1
=
πc3 cosh aτ 2 − cosh aτ 1 2 − sinh aτ 2 − sinh aτ 2
c c c c
a2 1
=− 3
πc cosh a(τ 2 −τ 1 ) − 1
c
a2 1
=− 3
.
πc 2 sinh a(τ 2 −τ 1 )
2
2c
(17.236)
The above can be rewritten as [see Eqs. (17.137) and (17.138)]
πkB TUD 2
2
C (x (τ 2 ) − x (τ 1 ) , t (τ 2 ) − t (τ 1 )) = − ,
πc sinh2 πkB TUD (τ 2 −τ 1 )
(17.237)
which implies that the effective temperature is TUD [see Eq. (17.233)].
and
it 3 4
H0 , B1† B2† = 2iω 0 tB1† B2† , (17.244)
thus
where
3 4
S (ξ, φ) = exp ξ e−2iφ B1 B2 − e2iφ B1† B2† , (17.249)
and where
t
ξ = ω0 dt′ ζ (t′ ) . (17.250)
0
3 4
−ξ e−2iφ B1 B2 − e2iφ B1† B2† , ξ 2 B1 = −ξ 3 e2iφ B2† ,
(17.253)
3 4
−ξ e−2iφ B1 B2 − e2iφ B1† B2† , −ξ 3 e2iφ B2† = ξ 4 B1 , (17.254)
..
.
thus
and similarly,
B1 + B2 + B1† + B2†
Xθ = cos θ √
2
B1 − B2 + B1† − B2†
+ sin θ √
2
B1 + B †
= (cos θ + sin θ) √ 1
2
B2 + B †
+ (cos θ − sin θ) √ 2
2
π
= cos θ − B1 + B1†
4
π
+ cos θ + B2 + B2† .
4
(17.260)
S † (ξ, φ) Xθ S (ξ, φ)
π
= cos θ − cosh ξ B1 + B1†
4
π
+ cos θ + cosh ξ B2 + B2†
4
π
− cos θ − sinh ξ B2† e2iφ + B2 e−2iφ
4
π
− cos θ + sinh ξ B1† e2iφ + B1 e−2iφ .
4
(17.261)
ξ, φ| Xθ |ξ, φ = 0 , (17.262)
ξ, φ| (∆Xθ )2 |ξ, φ
π π
= cos θ − cosh ξ − cos θ+ sinh ξe−2iφ
4 4
π π
× cos θ − cosh ξ − cos θ+ sinh ξe2iφ
4 4
π π
+ cos θ + cosh ξ − cos θ− sinh ξe−2iφ
4 4
π π
× cos θ + cosh ξ − cos θ− sinh ξe2iφ .
4 4
(17.263)
ξ, φ| (∆Xθ )2 |ξ, φ = cosh (2ξ)−sinh (2ξ) cos (2θ) cos (2φ) . (17.264)
B1 − B1† + B2 − B2†
Pθ = i cos θ √
2
B1 − B1† − B2 + B2†
+ i sin θ √
2
π
= i cos θ − B1 − B1†
4
π
+ i cos θ + B2 − B2† .
4
(17.265)
S † (ξ, φ) Pθ S (ξ, φ)
π
= i cos θ − cosh ξ B1 − B1†
4
π
+ i cos θ + cosh ξ B2 − B2†
4
π
+ i cos θ + sinh ξ B1 e−2iφ − B1† e2iφ
4
π
+ i cos θ − sinh ξ B2 e−2iφ − B2† e2iφ
4
ξ, φ| Pθ |ξ, φ = 0 , (17.266)
ξ, φ| (∆Pθ )2 |ξ, φ = cosh (2ξ)+sinh (2ξ) cos (2θ) cos (2φ) . (17.267)
Σ− = −B1 B2 , (17.269)
Σ+ = B1† B2† , (17.270)
Σx = Σ+ + Σ− , (17.272)
Σy = −i (Σ+ − Σ− ) , (17.273)
Σz = [Σ+ , Σ− ] . (17.274)
and
or
B1 B2 |0, 0 = 0 , (17.283)
B1 B1† + B2† B2 |0, 0 = |0, 0 , (17.284)
the state |ξ, φ = S (ξ, φ) |0, 0 can be easily expanded in the basis of
number states |n1 , n2 B
where
n1 n2
B1† B2†
|n1 , n2 B = √ √ |0, 0 , (17.286)
n1 ! n2 !
and where |0, 0 is the ground state of H0 . With the help of Eq.
(17.238) one finds for the case |ψI (0) = |0, 0 that
thus
∞
=
e2ni(φ−ω0 t) tanhn ξ
|ψ (t) = − n=0 |n, n B ,
cosh ξ
(17.288)
or
or
O1 = Tr (ρeff O1 ) ,
A21 − A†2 2 †2
1 − A2 + A2
= B1 B2 − B1† B2† ,
2
and thus the operator S (ξ, φ) [see Eq. (17.145)] for the case φ = 0 is
given by [see Eqs. (17.258) and (17.259)]
3 4
S (ξ, 0) = exp ξ B1 B2 − B1† B2†
ξ
= exp A21 − A†2 2
1 − A2 + A2
†2
2
ξ A21 − A†2
1 ξ A22 − A†2
2
= exp exp − .
2 2
(17.294)
With the help of the above result (17.295) one finds that the probabil-
′ ′
ity distribution function Px X+ , X− is a joint normal distribution
given by
′ ′ ′ ′ 2
Px X+ , X− = ψS X+ , X−
! "
′ 2 ′ 2
1 e2ξ X+ ′
+ X− + e−2ξ X+′
− X−
= exp −
π 2
′
X+ 2 ′
X− 2 ′ X′
2ρx X+
1 −
− + −
2(1−ρ2
x)
σx σx σ2
e x
= ,
2πσ2x 1 − ρ2x
(17.296)
where
and
' ) ′ * 2
( 1
′ ′ ′
X+ − X+ |X− |X− = 1 − ρ2x σ2x = . (17.301)
2 cosh (2ξ)
In this chapter two models are discussed, the London’s model, in which a
macroscopic wavefunction is introduced to describe the state of a supercon-
ductor, and the model by Bardeen, Cooper and Schrieffer (BCS), which pro-
vides an insight on the underlying microscopic mechanisms that are respon-
sible for superconductivity.
Taking the curl of Eq. (18.20) and employing Eqs. (18.19), (18.21) and (18.23)
together with the general vector identity
∇ × ∇ × B = ∇ (∇ · B) − ∇2 B (18.24)
lead to
1 1 ∂2B
∇2 B = 2 B + c2 ∂t2 , (18.25)
λL
where
;
m⋆s c2
λL = (18.26)
4πn⋆s qs⋆2
is the London penetration depth in Gaussian units (λL = m⋆s /µ0 n⋆s qs⋆2 in
SI units). In terms of the superconducting plasma frequency ωp,s , which is
given by
4πn⋆s qs⋆2
ω 2p,s = , (18.27)
m⋆s
dθ m⋆
− = ⋆2 s ⋆2 J2s + qs⋆ ϕ . (18.30)
dt 2qs ns
Taking the time derivative of Eq. (18.18) and employing Eq. (18.1) together
with the last result yield the first London equation
m⋆s ∂Js 1
+ ⋆ ⋆ ∇J2s =E. (18.32)
qs n⋆s
⋆2 ∂t 2qs ns
|Js | q ⋆2 n⋆
≃ s ⋆ s |B| . (18.33)
l0 ms c
∂Js 1
+ ∇J2s = 0 . (18.36)
∂t 2qs⋆ n⋆s
or
∂vs
m⋆s = qs⋆ E . (18.38)
∂t
The above relation (18.38) is analogous to the classical equation of motion
given by Eq. (18.4) for the case of vanishing magnetic field. The absence of
any damping term in Eq. (18.38) represents the nullification of resistance in
superconductors.
Flux Quantization. Consider a close curve C inside a superconductor. In-
tegrating Eq. (18.18), which is given by
where
hc
φs = (18.43)
qs⋆
hc
φs = . (18.44)
2e
As was shown above, the second London equation implies that the super-
current density Js vanishes deep inside a superconductor. Consider a close
curve C inside a superconductor and assume that the distance between any
point on C and the nearest surface is much larger than the London penetra-
tion depth λL . For such a curve the left hand side of Eq. (18.42) vanishes,
and consequently
φC = nφs , (18.45)
i.e. the magnetic flux is quantized in units of the superconducting flux quan-
tum.
The state vector of the junction |φ is expressed in terms of basis states |φL
and |φR as
1/2 1/2
|φ = nL eiθL |φL + nR eiθR |φR , (18.46)
where nL,R and θ L,R are all real, and where the normalized states |φL and
|φR , which represent, respectively, the left and right ports of the junction, are
orthogonal to each other, i.e. φL |φR = 0. The Hamiltonian of the system
is taken to be given by
H = EL |φL φL | + ER |φR φR |
+ geiφ |φL φR | + ge−iφ |φR φL | ,
(18.47)
where EL,R , g and φ are all real (to ensure that H is Hermitian). The energy
expectation value is given by
φs
φ| H |φ = nL EL + nR ER + Ic cos Θ ,
2π
where φs = hc/2e is the flux quantum [see Eq. (18.44)], the so-called critical
current Ic is given by
√
4e nL nR g
Ic = , (18.48)
c
and the relative phase Θ is given by
Θ = θL − θR − φ . (18.49)
The Schrödinger equation, which reads
d |φ
i = H |φ , (18.50)
dt
yields
! " ! "
1/2 1/2
d nL eiθL EL geiφ nL eiθL
i 1/2 = −iφ 1/2 , (18.51)
dt nR eiθR ge ER nR eiθR
or
1/2
dnL 1/2 dθL EL 1/2 1/2
+ inL = −i nL + gnR e−iΘ , (18.52)
dt dt
1/2
dnR 1/2 dθR ER 1/2 1/2
+ inR = −i nR + gnL eiΘ , (18.53)
dt dt
or
dnL dθ L EL cIc −iΘ
+ 2inL + = −i e , (18.54)
dt dt 2e
dnR dθR ER cIc
+ 2inR + = −i eiΘ . (18.55)
dt dt 2e
Let I (t) and V (t) be the current through and voltage across a Josephson
junction, respectively, at time t. The energy UJ of the junction can be eval-
uated by calculating the work done by the source
t
UJ = dt′ I (t′ ) V (t′ ) . (18.60)
With the help of the first (18.56) and second (18.58) Josephson relations this
becomes
Θ
Ic
UJ = dΘ′ sin Θ′ , (18.61)
2e
thus up to a constant UJ is given by
where
Ic φ Ic
EJ = = s . (18.63)
2e 2πc
The energy UJ (18.62) can be expressed as [compare with Eq. (18.384)
below]
;
2
I
UJ = −EJ 1 − . (18.64)
Ic
LJ I 2
UJ = −EJ + + O I4 , (18.65)
2
where
φs
LJ = (18.66)
2πcIc
is the so-called Josephson inductance. Note, however, that an inductor-like
behavior of a Josephson junction is expected only when I ≪ Ic .
In terms of the superconducting flux quantum φs [see Eq. (18.44)] Eq. (18.39)
can be rewritten as
qs⋆ n⋆s
Js = ∇θGI , (18.67)
m⋆s
where
2π
∇θGI = ∇θ − A. (18.68)
φs
The phase factor θGI is commonly called the gauge invariant phase.
Consider an integral over Js (18.67) along a path going through a Joseph-
son junction from point r1 on the interface between the first superconductor
and the barrier to point r2 on the interface between the second supercon-
ductor and the barrier. The phase difference Θ is obtained by integrating
∇θGI
r2 r2
2π
Θ= dr · ∇θGI = θ (r2 ) − θ (r1 ) − dr · A . (18.69)
r1 φs r1
18.3 RF SQUID
A radio frequency (RF) superconducting quantum interference device (SQUID)
is made of a superconducting loop interrupted by a Josephson junction (see
Fig. 18.1). Consider a close curve C going around the loop. The requirement
that the phase θ of the macroscopic wavefunction is continues reads
@
2nπ = dr · ∇θ , (18.70)
C
where n is integer. The section of the close curve C inside the superconductor
is denoted by C − and the integral through the junction is denoted as an
integral from point r1 to point r2 . With the help of Eq. (18.39) the above
condition becomes
r2
m⋆s m⋆s 2π
2nπ = dr · Js + dr · Js + dr · A . (18.71)
qs⋆ n⋆s r1 qs⋆ n⋆s C− φs C
Consider the case where the curve is chosen such that the supercurrent density
Js vanishes everywhere on the curve C − (i.e. inside the superconductor the
distance between any point on C − and the nearest surface is much larger than
the London penetration depth λL ). For this case Eq. (18.71) becomes
2πφ
2nπ = Θ + , (18.72)
φs
where
r2 r2
m⋆s
Θ= dr · Js = dr · ∇θGI (18.73)
qs⋆ n⋆s r1 r1
is the gauge invariant phase difference across the junction [see Eqs. (18.67)
and (18.69)] and where
@
φ= dr · A (18.74)
C
is the magnetic flux threaded through the area enclosed by the closed path
C [see Eq. (18.41)].
The junction’s critical current is labeled by Ic . It is assumed that the
junction has capacitance, which is denoted by CJ . Consider the case where a
magnetic flux that is denoted by φe is externally applied. The total magnetic
flux φ threading the loop is given by
φ = φe + ΛIs , (18.75)
where Is is the circulating current flowing in the loop and Λ is the self induc-
tance of the loop.
18.3.1 Lagrangian
The Lagrangian of the system [see Eq. (1.16)] can be expressed as a function
of the dimensionless flux coordinate Φ, which is defined by
2πφ
Φ= , (18.76)
φs
and its time derivative Φ̇. According to Faraday’s law of induction the voltage
across the capacitor (in Gaussian units) is
φ̇
V =− , (18.77)
c
and therefore the kinetic energy of the system T is the capacitance energy
2
CJ φ̇ CJ φ2s Φ̇2
T = = . (18.78)
2c2 8π2 c2
The potential energy U has two contributions, the inductive energy (in
Gaussian units)
ΛIs2 (φ − φe )2 φ2 (Φ − Φe )2
= = s , (18.79)
2c 2Λc 8π 2 Λc
where
2πφe
Φe = (18.80)
φs
is the normalized external flux, and the Josephson energy UJ [see Eqs. (18.62)
and (18.72)]
φs Ic
UJ = − cos Φ . (18.81)
2πc
Thus the Lagrangian L = T − U is given by
φ2s
E0 = , (18.84)
8π2 Λc
the junction’s plasma frequency ω p is given by
0 0
c 2ecIc
ωp = = , (18.85)
LJ CJ CJ
where LJ = φs /2πcIc is the Josephson inductance [see Eq. (18.66)], the di-
mensionless potential u (Φ; Φe ) is given by
d ∂L ∂L
= , (18.88)
dt ∂ Φ̇ ∂Φ
thus
Λ Φ̈
+ Φ − Φe + β L sin Φ = 0 . (18.89)
LJ ω 2p
With the help of Eqs. (18.72), (18.75) and (18.77) the equation of motion can
be rewritten as
Is = Ic sin Θ + CJ V̇ . (18.90)
The above equation states that the circulating current Is equals the sum of
the current Ic sin Θ through the Josephson junction and the current CJ V̇
through the capacitor.
φ = φe + φi , (18.91)
where the term φi represents the flux generated by both, the circulating
current in the RF SQUID Is and by the current in the inductor of the LC
resonator IL
φi = ΛIs + M IL , (18.92)
Fig. 18.2. The LC resonator that is coupled to the RF SQUID allows readout.
where Λ is the self inductance of the loop. Similarly, the magnetic flux ϕ in
the inductor of the LC resonator is given by
φi Λ M Is
= . (18.94)
ϕ M L IL
Inverting the above relation allows expressing the currents Is and IL in terms
of φi = φ − φe and ϕ
φi Mϕ
Is = − , (18.95)
Λ (1 − K 2 ) ΛL (1 − K 2 )
ϕ Mφi
IL = − , (18.96)
L (1 − K 2 ) ΛL (1 − K 2 )
where the dimensionless constant K is given by
M
K=√ . (18.97)
ΛL
Exercise 18.3.1. Show that the equations of motion governing the dynam-
ics of the system are given by
2πMϕ
Λ Φ̈ Φ − Φe − φ L
2
=− s
− β L sin Φ , (18.98)
LJ ω p 1 − K2
and
C ϕ̈ ϕ − φ2πΛ
sM
(Φ − Φe )
=− + Iin . (18.99)
c L (1 − K 2 )
Λ − 2φΛL
i ϕM
+ ϕL
=
2c (1 − K 2 )
2
ϕ2 φi − Mϕ
L
= +
2cL 2cΛ (1 − K 2 )
2
Cω 2e ϕ2 φ2s Φ − Φe − 2πMϕ
φs L
= + ,
2c2 8π2 cΛ (1 − K 2 )
(18.101)
where
0
c
ωe = (18.102)
LC
is the LC angular resonance frequency. The total potential energy U is given
by
Iin ϕ φs Ic
U = UI − − cos Φ , (18.103)
c 2πc
where the term −Iin ϕ/c is the potential energy of the current source and
− (φs Ic /2πc) cos Φ is the Josephson energy [see Eq. (18.81)]. With the help of
the above relations one finds that the Lagrangian of the system L = T − U
can be expressed as
L = L0 + L1 , (18.104)
(Φ − Φe,eff )2
uK (Φ; Φe,eff ) = − 2β L cos Φ , (18.107)
1 − K2
and where the effective external flux Φe,eff is given by
2πM ϕ
Φe,eff = Φe + . (18.108)
φs L
Note that L1 depends on the effective external flux Φe,eff , which, in turn,
depends on the coordinate ϕ of the LC resonator [see Eq. (18.108)]. This
dependence gives rise to the coupling between the LC resonator and the RF
SQUID. The Euler - Lagrange equations (1.8), which are given by
d ∂L ∂L
= , (18.109)
dt ∂ Φ̇ ∂Φ
d ∂L ∂L
= , (18.110)
dt ∂ ϕ̇ ∂ϕ
leads to Eqs. (18.98) and (18.99).
With the help of Eqs. (18.95) and (18.96) one finds that the equations of
motion (18.98) and (18.99) can be rewritten as
CJ φ̈
Is = Ic sin Θ − , (18.111)
c
and
C ϕ̈
Iin = + IL . (18.112)
c
While Eq. (18.111) expresses the law of current conservation in the SQUID
loop, Eq. (18.112) expresses the same law in the LC resonator.
For a given value of the coordinate ϕ, local minima points of the potential
uK (Φ; Φe,eff ) are found by solving [see Eq. (18.107)]
Φ − Φe,eff
0= + sin Φ . (18.113)
β L (1 − K 2 )
When β L 1 − K 2 < 1 the above equation has a single solution, which to
first order in β L 1 − K 2 is given by
Φ = Φe,eff + β L 1 − K 2 sin Φe,eff . (18.114)
As will be shown below, when the dynamics of the LC resonator can be
considered as slow in comparison with the dynamics of the RF SQUID, i.e.
when ω e ≪ ωp , the effective resonance frequency of the LC resonator, which
is denoted by ω e,eff , becomes periodically dependent on the magnetic flux Φe
that is externally applied to the RF SQUID. This dependency can be utilized
for magnetic fields sensing using the system under study.
Exercise 18.3.2. Consider the case where β L ≪ 1, K 2 ≪ 1 and ω e ≪ ωp .
Show that for this case the effective value of the angular resonance frequency
of the LC resonator is approximately given by
β L K 2 cos Φe
ω e,eff = ω e 1 + . (18.115)
2
Solution 18.3.2. In the limit ωe ≪ ω p the coordinate Φ is expected to be
given by Eq. (18.114), i.e. it is assumed to adiabatically follow the assumed
slow dynamics of the LC resonator. When the coupling between the LC res-
onator and the RF SQUID is weak, i.e. when K 2 ≪ 1, the effective resonance
frequency of the LC resonator is expected to be given by
c2 ∂ 2 L
ω 2e,eff = − , (18.116)
C ∂ϕ2
where the Lagrangian is given by Eq. (18.104), and where the second order
derivative is calculated at a local minima point of the potential energy of the
system. With the help of the relation [see Eq. (18.108)]
∂ ∂Φe,eff ∂ 2πM ∂
= = , (18.117)
∂ϕ ∂ϕ ∂Φe,eff φs L ∂Φe,eff
and the expansion [see Eqs. (18.107) and (18.114)]
uK (Φ; Φe,eff ) = −2β L cos (Φe,eff ) + O β 2L , (18.118)
one finds to first order in β L that (recall that it is assumed that K 2 =
M 2 /ΛL ≪ 1)
ω 2e,eff = ω 2e 1 + β L K 2 cos Φe , (18.119)
in agreement with Eq. (18.115) when β L ≪ 1.
18.3.3 Hamiltonian
The variables canonically conjugate to Φ and ϕ are given by [see Eqs. (1.20)
and (18.104)]
∂L 2E0 ΛΦ̇
Q= = , (18.120)
∂ Φ̇ LJ ω2p
∂L C ϕ̇
q= = 2 . (18.121)
∂ ϕ̇ c
The Hamiltonian is given by [see Eq. (1.22)]
H = QΦ̇ + q ϕ̇ − L = H0 + H1 , (18.122)
where
c2 q 2 Cω 2e ϕ2 Iin ϕ
H0 = + − , (18.123)
2C 2c2 c
and where
LJ ω2p Q2
H1 = + E0 uK (Φ; Φe,eff ) . (18.124)
4E0 Λ
Quantization is achieved by regarding the variables {Φ, Q, ϕ, q} as Her-
mitian operators satisfying the following commutation relations [see Eqs.
(3.6), (3.7) and (3.8)]
[Φ, Q] = [ϕ, q] = i , (18.125)
and
[ϕ, Φ] = [ϕ, Q] = [q, Φ] = [q, Q] = 0 . (18.126)
In terms of the annihilation operators A, which is given by [see Eq. (5.9)]
0
1 Cω e i
A= √ ϕ+ . q , (18.127)
2 c2 Cωe
c2
and the corresponding number operator N , which is given by [see Eq. (5.14)]
1 c2 q 2 Cω 2e ϕ2 1
N = A† A = + − , (18.128)
ωe 2C 2c2 2
the Hamiltonian H0 becomes
0
1
H0 = ω e N+ − Iin A + A† , (18.129)
2 2Cω e
and the term Φe,eff becomes [see Eq. (18.108)]
0
K ωe
Φe,eff = Φe + A + A† . (18.130)
2 E0
Solution 18.3.3. With the help of Eqs. (18.91), (18.107) and (18.124) one
finds that
2
∂H 2E0 2π Mϕ
−c =c φ − φe −
∂φe 1 − K2 φs L
φi − Mϕ
L
= ,
Λ (1 − K 2 )
(18.132)
in agreement with Eq. (18.131) [see Eq. (18.95)].
Consider the case where the externally applied magnetic flux φe is chosen to
be close to a half integer value in units of the superconducting flux quantum
φs . The potential uK (18.107) can be expressed as
(Φr − Φe,eff,r )2
uK = + 2β L cos Φr , (18.133)
1 − K2
where Φe,eff,r and Φr are defined by [see Eq. (18.108)]
2πMϕ
Φe,eff = Φe + = π + Φe,eff,r , (18.134)
φs L
Φ = π + Φr . (18.135)
Consider the case where Φe,eff,r = 0 (i.e. Φe,eff = π). For this case to
second order in Φr the potential uK is given by
1 − βL 1 − K 2 2
uK = 2β L + Φr + O Φ4r . (18.136)
1 − K2
Thus if β L 1 − K 2 > 1 the point Φr = 0 becomes a local maxima point of
u. The corresponding potential barrier centered at Φr = 0 (i.e. at Φ = π)
separates two symmetric potential wells on the right and on the left (see Fig.
18.3). At sufficiently low temperatures only the two lowest energy levels are
expected to be occupied. In this limit the Hamiltonian of the system can be
expressed in the basis of the states | and | , that represent localized
states in the left and right well, respectively, having opposite circulating cur-
rents. In this range the device can be used as an artificial two-level system
(TLS), i.e. as a quantum bit (qubit in short).
Fig. 18.3. Eigenstates of H1 . (a)-(c) The first 3 lowest energy states for the case
Φe,eff,r = 0. (d) The energy of the two lowest states vs. Φe,eff,r .
−1 1
H = ωe A† A +
2
ωf
+ (| |−| |)
2
ω∆
+ (| |+| |)
2
− g A + A† (| |−| |) .
(18.137)
Exercise 18.3.4. Let Icc (−Icc ) be the circulating current associated with
the state | (| ). Express the coefficient ωf in terms of Icc and the exter-
nally applied magnetic flux φe .
Solution 18.3.4. To ensure consistency with Eq. (18.131), i.e. to satisfy the
requirement
∂H ∂H
Icc = −c | | =c | | , (18.138)
∂φe ∂φe
2Icc φs Icc
ωf = φe − = (Φe − π) . (18.139)
c 2 e
As will be shown below, the energy ω ∆ is the smallest value of the qubit
energy gap, which is obtained when ω f = 0 [see Eq. (18.144) below]. Note
that it can be estimated using the WKB result (11.74) for the energy gap of a
double well potential. The coefficient g, which is called the coupling constant,
is given by [see Eq. (18.130)]
0
Icc K ωe
g=− . (18.140)
4e E0
Exercise 18.3.5. Consider the decoupled case, i.e. the case where g = 0.
Find the eigenstates and eigenenergies of the qubit.
|+ cos θ2 sin θ2 |
= , (18.141)
|− − sin θ2 cos θ2 |
where
ω∆
tan θ = , (18.142)
ωf
and the corresponding eigenenergies are
ωa
ε± = ± , (18.143)
2
where
.
ωa = ω2f + ω 2∆ . (18.144)
and
hold, where
Σz = |+ +| − |− −| , (18.147)
Σ+ = |+ −| , (18.148)
Σ− = |− +| , (18.149)
−1 1 ωa
H = ωe A† A + + Σz
2 2
− g A + A† [cos θ Σz − sin θ (Σ+ + Σ− )] ,
(18.150)
or
−1 1 ωa
HJC = ω e A† A + + Σz
2 2
+ g1 A† Σ− + AΣ+ ,
(18.152)
and g1 is given by
g1 = g sin θ . (18.154)
and
HJC |n + 1, − = ωe (n + 1) |n + 1, −
∆ √
+ |n + 1, − + g1 n + 1 |n, + ,
2
(18.156)
where
∆ = ωe − ωa , (18.157)
or in a matrix form
|n, +
HJC
|n + 1, −
10 ωn cos θ n sin θn
= ω e (n + 1) +
01 2 sin θn − cos θn
|n, +
× ,
|n + 1, −
(18.158)
where
.
ωn = ∆2 + 4g12 (n + 1) , (18.159)
√
2g1 n + 1
tan θn = − . (18.160)
∆
Thus, the states |n+ and |n− , which are given by [see Eqs. (6.221) and
(6.222)]
θn θn
|n+ = cos |n, + + sin |n + 1, − , (18.161)
2 2
θn θn
|n− = − sin |n, + + cos |n + 1, − , (18.162)
2 2
are eigenstates of HJC and the following holds
where
3 ωn 4
En± = ω e (n + 1) ±
2
0
∆2
= ω e (n + 1) ± + (n + 1) g12 .
4
(18.164)
where
∆
Eg = (18.166)
2
is the ground state energy.
While in the RWA the term VBS is disregarded, its effect, gives rise to the
so-called Bloch-Siegert shift.
Solution 18.3.7. As can be seen from Eq. (18.153), the perturbation VBS is
proportional to g1 . The exact eigenstates of HJC are given by Eqs. (18.161),
(18.162) and (18.165). All diagonal matrix elements of VBS vanish, and con-
sequently the lowest nonvanishing order of the perturbation expansion is the
second one [see Eq. (9.32)]. The nonvanishing matrix elements of VBS are
evaluated below to first order in g1
) ′ −1
n+ VBS |0, − = g1 δ n′ ,1 , (18.167)
) ′ −1
n− VBS |0, − = g1 cot θδ n′ ,0 , (18.168)
) ′ −1 √
n− VBS |n+ = g1 nδ n′ ,n−2 , (18.169)
and
) ′ −1
n− VBS |n−
√ √
= g1 cot θ n + 1δ n′ ,n−1 + n + 2δ n′ ,n+1 .
(18.172)
−1 ∆
Eg = + ωBS,0 , (18.173)
2
and the energies of the excited states by
−1
En± = (n + 1) (ωe ± ωBS )
0
∆2
± + (n + 1) g12 + ω BS,0 ,
4
(18.174)
where
g12
ω BS = , (18.175)
ωe + ωa
and where
1 cot2 θ
ω BS,0 = −g12 + . (18.176)
ωe + ω a ωe
The following holds
−1 g12
(En− − Eg ) = (n + 1) ωe − ω BS + + O g14 , (18.177)
∆
and
−1 g12
(En+ − E0+ ) = n ω e + ω BS − + O g14 , (18.178)
∆
thus in the linear regime and when g12 / |∆| ≪ ωe the system has two resonance
frequencies given by ω e ± ω BS ∓ g12 /∆.
18.3.6 Damping
The effect of damping on both a resonator and on a TLS has been discussed
in the previous chapter. In this section the effect of damping on the coupled
resonator-qubit system is being studied.
Exercise 18.3.8. Employ the RWA to derive equations of motion for the
operators A, Σz and Σ− .
Solution 18.3.8. With the help of Eqs. (4.37) and (18.152) together with
the commutation relations
, -
A, A† = 1 , (18.179)
[Σz , Σ+ ] = 2Σ+ , (18.180)
[Σz , Σ− ] = −2Σ− , (18.181)
[Σ+ , Σ− ] = Σz , (18.182)
one obtains (recall that in the RWA the term VBS is disregarded)
dA
= −iω e A − ig1 Σ− , (18.183)
dt
dΣz
= 2ig1 Σ− A† − AΣ+ , (18.184)
dt
dΣ−
= −iω a Σ− + ig1 AΣz , (18.185)
dt
where g1 = g sin θ [see Eq. (18.154)].
Damping can be taken into account by introducing the cavity decay rate
γ e [see Eq. (17.33)] and the qubit decay times T1 and T2 [see Eqs. (17.127)
and (17.128)]. The equation of motion for the cavity operator A (18.183)
leads to an equation of motion for the expectation value A = A [see Eq.
(17.34)], and the qubit equations of motion (18.184) and (18.185) lead to
equations of motion for the expectation values Pz = Σz and P− = Σ−
[see Eqs. (17.120) and (17.121)]
dA
+ (iωe + γ e ) A + ig1 P− = 0 , (18.186)
dt
dPz Pz − Pz0
+ 2ig1 (AP+ − P− A∗ ) = − , (18.187)
dt T1
dP− P−
+ iω a P− − ig1 APz = − , (18.188)
dt T2
where Pz0 is the value of Pz in thermal equilibrium [see Eq. (17.126)].
Consider the low temperature limit, for which kB T ≪ ω a and conse-
quently Pz0 ≃ −1 [see Eq. (17.126)]. In this limit Eq. (18.188) can be sim-
plified by employing the approximation −ig1 APz ≃ ig1 A, which allows ex-
pressing Eqs. (18.186) and (18.188) in a matrix form as
d A A
+ iM =0, (18.189)
dt P− P−
where
ωe − iγ e g1
M= , (18.190)
g1 ω a − iγ a
The macroscopic Maxwell’s equations (in Gaussian units) for the electric field
E, electric displacement D, magnetic induction B and magnetic field H in
the presence of external charge density ρext and external current density Jext
are given by
4π 1 ∂D
∇×H = Jext + , (18.196)
c c ∂t
1 ∂B
∇×E = − , (18.197)
c ∂t
∇ · D = 4πρext , (18.198)
∇·B = 0 . (18.199)
For an isotropic and linear medium the following relations hold
D = E+4πP , (18.200)
D = ǫE , (18.201)
P = χe E , (18.202)
B = H + 4πM , (18.203)
B = µH , (18.204)
M = χm H , (18.205)
where P is the electric polarization, ǫ = 1 + 4πχe is the permittivity (di-
electric constant of the medium), χe is the electric susceptibility, M is the
magnetization, µ = 1 + 4πχm is the permeability and χm is the magnetic
susceptibility.
In general, a scalar function f (r, t) can be Fourier expanded as
F = FL + FT , (18.208)
where the longitudinal part is given by FL = (q̂ · F) q̂, the transverse one
is given by FT = (q̂ × F) × q̂, and where q̂ = q/ |q| is a unit vector in the
direction of q. The following holds ∇ · F = ∇ · FL and ∇ × F = ∇ × FT .
Recall that for a general scalar φ and a vector A the following holds
∇ · (φA) = φ∇ · A + A · ∇φ ,
∇ × (φA) = φ∇ × A − A × ∇φ ,
thus
and
With the help of the above relations the Maxwell’s equations (18.196),
(18.197), (18.198) and (18.199) can be Fourier transformed into
4π iω
iq × HT (q, ω) = Jext (q, ω) − D (q, ω) , (18.211)
c c
ω
q × ET (q, ω) = B (q, ω) , (18.212)
c
iq · DL (q, ω) = 4πρext (q, ω) , (18.213)
q · BL (q, ω) = 0 . (18.214)
While the external charge density ρext is related to D by the relation [see
Eq. (18.198)]
∇ · D = 4πρext , (18.215)
the induced charge density ρind , which is defined as the change in charge
density with respect to the unperturbed case, is related to the electric polar-
ization by the relation ∇ · P = − ρind , and the total charge density ρind + ρext
is related to the electric field E by the relation
ρext (q, ω)
ǫ (q, ω) = . (18.219)
ρext (q, ω) + ρind (q, ω)
ϕext (q, ω)
ǫ (q, ω) = , (18.224)
ϕ (q, ω)
or
4π ρind (q, ω)
ǫ (q, ω) = 1 − . (18.225)
|q|2 ϕ (q, ω)
1 1
m r̈+ ṙ = q E + ṙ × B , (18.226)
τ tr c
where τ tr is the so-called scattering time. For simplicity the applied magnetic
field is assumed to vanish, i.e. B = 0. In terms of the current density vector
J, which is related to the velocity vector v = ṙ by the relation
J
v= , (18.227)
qn
Eq. (18.226) yields
m ∂J 1
+ J =E. (18.228)
q2 n ∂t τ tr
The current density J is related to the induced charge density ρind by the
continuity equation (18.7)
dρind
+∇·J =0 . (18.229)
dt
Applying ∇· to Eq. (18.228) and using Eqs. (18.216) and (18.229) lead to
d2 ρind 1 dρind
+ = −ω 2p (ρind + ρext ) , (18.230)
dt2 τ tr dt
where ω p , which is given by
4πq 2 n
ω 2p = , (18.231)
m
is the so-called plasma frequency.
By employing Fourier expansion
q 2 nτ tr
σ0 = , (18.238)
m
and Eq. (18.230) becomes
ω 2p
ρind (ω) = ρ (ω) . (18.239)
ω 2 − ω 2p + τiωtr ext
Thus the dielectric function in the long wavelength limit ǫ (0, ω) is given
by [see Eq. (18.219)]
4πiσ (ω)
ǫ (0, ω) = 1 + . (18.241)
ω
Alternatively, in terms of the so-called skin depth δ sd , which is given by
0
c 2
δ sd = , (18.242)
ωp ωτ tr
2ic2 1
ǫ (0, ω) = 1 + . (18.243)
δ 2sd ω 2 1 − iωτ tr
4π ρind (q, 0)
ǫ (q, 0) = 1 − . (18.244)
|q|2 ϕ (q, 0)
∂n
ρind (r) = −q 2 ϕ (r) . (18.246)
∂µ
When the thermal energy kB T is much smaller than the Fermi energy ǫF
the factor ∂n/∂µ is approximately the density of states at the Fermi energy
ǫF , which is given by [see Eq. (16.103)]
∂n m2 vF
≃ 2 3 , (18.247)
∂µ π
2 4πq 2 m2 vF
kTF = . (18.249)
π2 3
The above result (18.248) together with Eq. (18.244) yield
2
kTF
ǫ (q, 0) = 1 + . (18.250)
|q|2
where both the normal contribution σn (ω) and the superconductivity one
σ s (ω) are evaluated using the Drude model expression (18.237).
As was discussed above, the first London equation for the homogeneous
case (18.37) suggests that the resistance of superconductors vanishes. Ac-
counting for this by taking the scattering time τ tr to be effectively infinite
yields [see (18.237)]
qs⋆2 n⋆s
σs (ω) = i . (18.252)
ωm⋆s
For the case where normal conductance is carried by electrons having mass
me , charge qe , density ne and scattering time τ tr,e the normal conductivity
σ n (ω) is given by [see (18.237)]
qe2 nn τ tr,e 1
σn (ω) = . (18.253)
me 1 − iωτ tr,e
The dielectric constant ǫ (ω) in the two fluid model is thus given by [see
Eq. (18.241)]
or in terms of the skin depth δ sd [see Eq. (18.242)] and the London penetration
depth λL [see Eq. (18.26)]
2i 1 1
ǫ (ω) = 1 + 2 − , (18.255)
(δ sd k) 1 − iωτ tr,e (λL k)2
where k = ω/c. Note that 1/δ 2sd ∝ nn [see Eq. (18.242)], whereas 1/λ2L ∝ ns
[see Eq. (18.26)]. Note also that the ratio between these characteristic length
scales is given by
;
λL m⋆s ne qe2 ωτ tr,e
= . (18.256)
δ sd me n⋆s qs⋆2 2
The dielectric function in the zero frequency limit given by Eq. (18.250)
represents the effect of screening by free charge carriers (i.e. by conducting
electrons) of externally applied electric field. However, in the derivation of
Eq. (18.250) the screening by localized charges (i.e. ions in the lattice) has
not been taken into account.
The contribution of free charge carriers, which is denoted by ǫe , to the
total dielectric function ǫ is given according to the Thomas-Fermi approxi-
mation by [see Eq. (18.250)]
2
kTF
ǫe = 1 + . (18.257)
|q|2
The contribution of localized charges (i.e. lattice vibrations), which is denoted
by ǫi , is taken to be given by Eq. (18.240). When ωτ tr,i ≫ 1, where τ tr,i is
the effective scattering time of the localized charges, Eq. (18.240) yields
ω 2p,i
ǫi = 1 − , (18.258)
ω2
where ω p,i , which is given by
4πqi2 ni
ω 2p,i = , (18.259)
mi
is the ion plasma frequency and where mi , qi and ni are the ionic mass, charge
and density, respectively.
The total potential ϕ can be expressed as
ϕ = ϕext + ϕe + ϕi , (18.260)
ǫ = ǫe + ǫi − 1 , (18.264)
The above result (18.265) indicates that the effect of lattice vibrations be-
comes important only when ω ωp,i .
Let ρ (r′ ) be the electron density in a medium having volume V. Classi-
cally, the two-particle Coulomb interaction VTP (r1 , r2 ) = e2 / |r1 − r2 | [see
Eq. (16.106)] gives rise to energy V given by [see Eqs. (16.65) and (16.43) for
comparison with the analogous second-quantization expression]
1
V = d3 r′ d3 r′′ VTP (r′ , r′′ ) ρ (r′ ) ρ (r′′ ) . (18.266)
2
With the help of the Fourier expansion
1 ′ ′
ρ (r′ ) = 3/2 √
d3 q′ ρ (q′ ) eiq ·r (18.267)
(2π) V
and Eqs. (4.47) and (16.127) one finds that [see Eq. (16.128) for comparison
with the analogous second-quantization expression]
1 4πe2
V = d3 q ρ (q) ρ (−q) . (18.268)
2V |q|2
The effect of induced charges in the medium (i.e. screening) can be taken
into account by dividing by the dielectric constant of the medium ǫ [see Eq.
(18.224)]
1 4πe2
V = d3 q ρ (q) ρ (−q) . (18.269)
2V |q|2 ǫ
The expression for the Coulomb energy (18.269) together with the dielec-
tric constant (18.265) lead to the effective interaction coefficient for a pair of
electrons having wave vectors k′ and k′′ and energies ǫk′ and ǫk′′ respectively
4πe2 4πe2 1
vk′ ,k′′ = 2 = 2 2 Ω2
, (18.270)
|q| ǫ |q| + kTF 1 − p,i
2 ω
where
q = k ′′ − k′ , (18.271)
ǫk′′ − ǫk′
ω = , (18.272)
and where
2
2 |q|
Ωp,i = 2 2
ω 2p,i . (18.273)
|q| + kTF
The fact that ǫ−1 becomes negative when ω < Ωp,i indicates that the
effective (i.e. phonon mediated) electron-electron interaction becomes attrac-
tive in the limit of low frequencies. The characteristic energy interval Ωp,i
in which the interaction becomes attractive is of the order of the so-called
Debye energy ǫD , which represents the largest energy of an acoustic phonon
in the lattice.
↑ labels spin up state, ↓ labels spin down state, ǫk′ is the energy of both single
particle states |k′ , ↑ and |k′ , ↓ , ǫF is the Fermi energy [see Eq. (16.101)] and
+
1 |ǫk′ − ǫF | < ǫD
ζ k′ = . (18.276)
0 otherwise
The coupling constant g > 0 gives rise for an effective electron-electron at-
tracting interaction [see Eq. (18.270)]. The interaction is assume to couple
pairs of electrons whose energies are inside an energy interval of width 2ǫD
around the Fermi energy ǫF .
As can be seen from the comparison with the more general many-particle
interaction operator V given by Eq. (16.93), the BCS Hamiltonian con-
tains only interaction terms that represents annihilation (the factor Bk′ =
a−k′ ,↓ ak′ ,↑ ) and creation (the factor Bk† ′′ = a†k′′ ,↑ a†−k′′ ,↓ ) of electrons pairs
having zero total angular momentum. Moreover, the summation is restricted
only to the energy interval of width 2ǫD in which attractive interaction is
expected, and the effective interaction coefficients are all assumed to be iden-
tical [see for comparison Eq. (18.270)].
−∆ ζ k′ Bk† ′ − ∆∗ ζ k′ Bk′ ,
k′ k′
(18.278)
where
g
∆= ζ k′ Bk′ . (18.279)
V
k′
Note that the identity a†k′ ,↓ ak′ ,↓ = 1 − ak′ ,↓ a†k′ ,↓ has been employed
= to derive
the first term of HMF (and the resultant constant term k′ (ǫk′ − ǫF ) has
been removed).
where
ǫk′ − ǫF −∆ζ k′
Mk′ = , (18.282)
−∆∗ ζ k′ − (ǫk′ − ǫF )
where
bk′ ,↑ ak′ ,↑
= Uk−1 . (18.290)
b†−k′ ,↓ a†−k′ ,↓
′
where σ ∈ {↑, ↓} and the number operator Nk,σ (with respect to the bk,σ and
b†k,σ operators ) is given by
|∆|
sin (2θk′ ) = − . , (18.296)
(ǫk′ − ǫF )2 + |∆|2
ǫk′ − ǫF
cos (2θk′ ) = . , (18.297)
(ǫk′ − ǫF )2 + |∆|2
and
7
8
8 1 − √ ǫk′ −ǫF
9 (ǫk′ −ǫF )2 +|∆|2
sin (θ k′ ) = , (18.298)
2
7
8
8 1 + √ ǫk′ −ǫF
9 (ǫk′ −ǫF )2 +|∆|2
cos (θ k′ ) = . (18.299)
2
The value of the energy gap |∆| can be determined from Eq. (18.279). Let
nk′ ,σ denotes the expectation value of the operator Nk′ ,σ , i.e.
∆ (1 − n−k′ ,↓ − nk′ ,↑ )
Bk′ = . , (18.302)
2 (ǫk′ − ǫF )2 + |∆|2
gD0 4ǫ2
1= log D , (18.307)
2 ∆20
thus
1
∆0 = 2ǫD exp − . (18.308)
gD0
where CE ≃ 0.577 is Euler’s constant, one finds that [see Eq. (18.308)]
eCE
kB Tc = ∆0 ≃ 0.566 × ∆0 . (18.314)
π
General Temperature. The energy gap |∆| at temperature T can be nu-
merically evaluated from Eq. (18.305). To a good approximation the solution
can be expressed by the following analytical relation
;
3
T
|∆| ≃ ∆0 1 − . (18.315)
Tc
where
bk′ ,↑ Kk′ = eiφ∆ cos θk′ ak′ ,↑ − e−iφ∆ sin θk′ a†−k′ ,↓
b−k,↓ |Ψ0 = 0 .
Alternatively, the ground state |Ψ0 , which is given by Eq. (18.316), can
be expressed as
#
|Ψ0 = C0 1 − γ k′ a†k′ ,↑ a†−k′ ,↓ |0 , (18.323)
k′
(18.325)
2
Furthermore, since a†k′ ,↑ a†−k′ ,↓ = 0 the following holds
! "
|Ψ0 = C0 exp − γ k′ a†k′ ,↑ a†−k′ ,↓ |0 . (18.326)
k′
′ ′′′
= a†k′ ,↑ a†−k′ ,↓ dr′′′ γ (r′′′ ) eik ·r ,
k′
(18.327)
where Ψσ (r′ ) is quantized field operators [see Eq. (16.96)]. In view of the
above result the ground state |Ψ0 , which is given by Eq. (18.326), can be
expressed as
|Ψ0 = C0 exp dr′ dr′′ γ (r′′ − r′ ) Ψ↑† (r′ ) Ψ↓† (r′′ ) |0 , (18.328)
where the function γ (r′′ − r′ ), which is called the pairing wavefunction, sat-
isfies
′ ′′′
dr′′′ γ (r′′′ ) eik ·r = −γ k′ , (18.329)
= −eiΘ e−iφ∆ sin θk′ a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ a†k′ ,↑ a†−k′ ,↓ |0
3 4
= −eiΘ e−iφ∆ sin θk′ a†k′ ,↑ a†−k′ ,↓ 1 − a†k′ ,↑ ak′ ,↑ + a†k′ ,↑ a†−k′ ,↓ 1 − a†−k′ ,↓ a−k′ ,↓ |0
d Bk′
i = 2µ Bk′ . (18.348)
dt
Thus, the complex energy gap ∆, which is given by [see Eq. (18.279)]
g
∆= Bk′ , (18.349)
V
k′
2µt 2eV t
Θ (t) = = . (18.351)
H = H1 + H2 + HT , (18.353)
where
Hk′ ,k′′ = Ak′ ,k′′ b†1,−k′ ,↓ b†2,k′′ ,↑ − b†1,k′ ,↑ b†2,−k′′ ,↓
+A∗k′ ,k′′ (b2,k′′ ,↑ b1,−k′ ,↓ − b2,−k′′ ,↓ b1,k′ ,↑ )
+Bk′ ,k′′ b1,−k′ ,↓ b†2,−k′′ ,↓ + b1,k′ ,↑ b†2,k′′ ,↑
where T = |tk′ ,k′′ |2 and where Θ is the relative phase difference between
the two superconductors. The energy correction δE can be expressed as a
function of Θ as
RN = , (18.368)
4πe2 V 2 D02 T
lead to
∞ ∞
|∆| dθ1 dθ2
EJ = 2
πe RN 0 0 cosh θ 1 + cosh θ 2
∞ ∞
|∆| dθ1 dθ2
=
4πe2 RN −∞ −∞ cosh θ 1 + cosh θ2
∞ ∞
|∆| dθp dθ m
= ,
4πe2 RN −∞ cosh θp −∞ cosh θm
(18.378)
thus
π |∆|
EJ = . (18.379)
4e2 RN
For arbitrary temperature the result is
π |∆| β |∆|
EJ = tanh . (18.380)
4e2 RN 2
As was shown above [see Eq. (18.364)], the energy of a Josephson junction
UJ having phase Θ relative to the energy when the phase vanishes is given
by [compare with Eq. (18.62)]
Let I (t) and V (t) be the current through and voltage across a Josephson
junction, respectively, at time t. Assume that initially at time t = 0 the
phase Θ vanishes. Energy conservation leads to the requirement that
t
UJ = dt′ I (t′ ) V (t′ ) . (18.382)
0
With the help of the second Josephson relation Θ̇ = (2e/ ) V (18.352) and
Eq. (18.381) this becomes
Θ
−EJ cos Θ = dΘ′ I (t′ ) . (18.383)
2e 0
Taking the derivative with respect to Θ leads, in agreement with Eq. (18.56),
to the first Josephson relation
I = Ic sin Θ , (18.384)
where the so-called critical current Ic is given by
2eEJ 2πcEJ
Ic = = , (18.385)
φs
where
hc
φs = (18.386)
2e
is the superconducting flux quantum, which is identical to the superconduct-
ing flux quantum given by Eq. (18.44) provided that the charge qs⋆ is taken to
be 2e. Note also that for the ’normal’ flux quantum φ0 given by Eq. (12.48)
the charge of elementary carrier is e.
18.7 Problems
1. Rotating Superconductor - Consider a superconductor rotating at
angular frequency Ω around the z axis. In the presence of an externally
applied magnetic field B calculate the magnetic field deep inside the
superconductor.
2. Consider a conductor containing charge carriers having charge q and
mass m. The density of charge carriers at point r is n (r) and the current
density is J (r). Contrary to the case of a normal metal, it is assumed
that all charge carriers at point r move at the same velocity v, which is
related to J by the relation [see Eq. (18.227)]
J
v= . (18.387)
qn
Show that in steady state this assumption leads to the 2nd London equa-
tion [see Eq. (18.25)]
1
∇2 H = H, (18.388)
λ2L
where H is the magnetic field and where
;
mc2
λL = . (18.389)
4πnq 2
3. Consider the so-called gradiometer RF SQUID seen in Fig. 18.4. The
junction’s critical current is labeled by Ic . It is assumed that the junc-
tion has capacitance, which is denoted by CJ . Consider the case where a
magnetic flux that is denoted by φe1 (φe2 ) is externally applied to the up-
per (lower) loop. Let Λ1 (Λ2 ) be the self inductance of the upper (lower)
loop. Derive an equation of motion for the system.
4. Consider the so-called DC SQUID device seen in Fig. 18.5. The Joseph-
son junctions on both arms of the DC SQUID have critical currents Ic1
and Ic2 respectively and both have the same capacitance CJ . The self
inductance of the loop is denoted as Λ. A bias current Ib is externally
injected and a magnetic flux φe is the externally applied to the loop. Find
equations of motion that govern the device’s dynamics.
5. Cooper pair box - Find an Hamiltonian for the device seen in Fig. 18.6.
' (
6. Calculate nP and (∆nP )2 with respect to the BCS ground state |Ψ0 ,
where nP is the pairs number operator (18.335).
7. Calculate the energy density of states for elementary excitations in a
superconductor.
8. Find the time evolution of the operators ak,↑ (t) and a†−k,↓ (t) at time
t [see the Hamiltonian (18.274)] for given initial conditions ak,↑ (0) and
a†−k,↓ (0) at time t = 0.
9. Dicke model - Consider a system composed of N TLSs interacting with
a single cavity mode having angular frequency ω e . Assume that all TLSs
have the same energy spacing ωa and the same coupling coefficient to
the cavity mode, which is denoted by gs . In the RWA the Hamiltonian
of the system is taken to be given by [compare with Eq. (18.152)]
−1 1 ωa
HD = ω e A† A + + Σz
2 2
+ gs A† Σ− + AΣ+ ,
(18.390)
where the operators Σ± and Σz are related to the single TLS operators
Σ±,n and Σz,n [see Eqs. (18.147), (18.148) and (18.149)] by
N
Σ± = Σ±,n , (18.391)
n=1
N
Σz = Σz,n . (18.392)
n=1
Assume that the single TLS operators Σ±,n and Σz,n satisfy the com-
mutation relations (18.180), (18.181) and (18.182) (which implies that
the operators Σ± and Σz satisfy the same relations). In the so-called
Holstein-Primakoff transformation the operators Σ± and Σz are ex-
pressed as
1/2
Σ+ = B † N − B † B , (18.393)
1/2
Σ− = N − B † B B, (18.394)
†
Σz = −N + 2B B , (18.395)
where N is a positive constant.
a) Show that the operators Σ± and Σz given by Eqs. (18.393), (18.394)
and (18.395) satisfy the commutation relations (18.180), (18.181) and
(18.182) provided that the operator B satisfies the following commu-
tation relation
, -
B, B † = 1 . (18.396)
18.8 Solutions
FΩ = 2m⋆s v × Ω , (18.402)
1 2m⋆s c
∇2 B = B+ Ω . (18.404)
λ2L qs⋆
is the magnetic energy [see Eq. (14.38)]. With the help of the Maxwell’s
equation (18.196) and Eq. (18.227) E can be expressed in terms of H as
1 3 4
E= λ2L (∇ × H)2 + H2 dV . (18.406)
8π V
With the help of the general vector identity [see Eq. (14.41)]
∇ · (F1 × F2 ) = (∇ × F1 ) · F2 − F1 · (∇ × F2 ) , (18.408)
one finds (for the case where F1 and F2 are taken to be given by F1 =
∇ × H and F2 = δH) that
The volume integral over the second term on the right hand side can be
expressed as a surface integral using the divergence theorem. However,
when boundary conditions of δH = 0 on the surfaces are applied the
surface integral vanishes. Thus Eq. (18.407) becomes
1
0 = δE = −λ2L ∇2 H + H · δH dV . (18.411)
4π V
where
2πφe1
Φe1 = − 2πn1 , (18.418)
φs
2πφe2
Φe2 = − 2πn2 , (18.419)
φs
are the normalized external fluxes. Using the notation
1 1 1
= + , (18.420)
Λ0 Λ1 Λ2
Λ0 Φe1 Λ0 Φe2
Φe0 = − , (18.421)
Λ1 Λ2
the Lagrangian can be expressed as
2
CJ φ2s Φ̇2 φ2s (Φ − Φe0 ) φ Ic
L= 2 2
− 2
+ s cos Φ + CG , (18.422)
8π c 8π cΛ0 2πc
where the constant CG is given by
φ2s Λ0 2
Λ1 Φe1 + Λ0 2
Λ2 Φe2 − Φ2e0
CG = − . (18.423)
8π 2 cΛ0
The resulting Euler - Lagrange equation of motion (1.8) is given by
d ∂L ∂L
= , (18.424)
dt ∂ Φ̇ ∂Φ
thus
CJ φs Φ̈ φ (Φ − Φe0 )
=− s − Ic sin Φ . (18.425)
2πc 2πΛ0
Note that [see Eqs. (18.412) and (18.413)]
2πΛ0 (Is1 − Is2 )
Φ − Φe0 = , (18.426)
φs
thus with the help of Eqs. (18.58) and (18.416) the equation of motion
(18.425) can be expressed as a current conservation law [compare with
Eq. (18.90)]
ΛIs
φ = φe + , (18.428)
2
where φe is the externally applied flux,
Is = I1 − I2 (18.429)
is the circulating current in the loop, and where I1 and I2 are the to-
tal currents flowing in the upper and lower arms respectively. The bias
current is given by Ib = I1 + I2 . In general the Josephson current IJk in
each junction (k = 1, 2) is related to the critical current Ick and to the
Josephson phase γ k by [see Eq. (18.56)]
where n is integer [see Eq. (18.72)]. By using this and Eq. (18.428) one
finds that
φs 2πφe
Is = − γ1 − γ2 + − 2πn . (18.433)
πΛ φs
The total voltage across the DC SQUID, which is denoted as VS is given
by
Λ dI1 Λ dI2
VS = V1 + = V2 + , (18.434)
2c dt 2c dt
and therefore
1 Λ dI1 dI2
VS = V1 + V2 + +
2 2c dt dt
1 d φs ΛIb
= (γ + γ 2 ) + .
2 dt 2π 1 2c
(18.435)
To model the current biasing an inductor having inductance LB is added
in parallel to the SQUID (see Fig. 18.8). As will be shown below, in the
limit where LB ≫ Λ the added inductor allows modeling current biasing.
The requirement that the voltage across the DC SQUID is the same as
the one across the inductor LB yields [see Eq. (18.435)]
dφB 1 d φs ΛIb
= (γ 1 + γ 2 ) + , (18.436)
dt 2 dt 2π 2c
where φB is the magnetic flux in the inductor LB . In what follows the bias
current Ib is assumed to be a constant. Integrating the above equation
yields
φs
φB = (γ + γ 1 ) + φB0 , (18.437)
4π 1
where the constant φB0 is taken to be given by φB0 = −LB Ib . The
Lagrangian of the closed system L = T − U [see Eq. (1.16)] is expresses
as a function of the coordinates γ 1 and γ 2 and their time derivatives γ̇ 1
and γ̇ 2 , where the kinetic energy T is given by
CJ φ2s γ̇ 21 + γ̇ 22
T = , (18.438)
8π 2 c2
and the potential energy U by [see Eq. (18.433)]
φs (Ic1 cos γ 1 + Ic2 cos γ 2 )
U =−
2πc
2
2πφe
φ2s γ 1 − γ 2 + φs − 2πn φ2B
+ + .
8π2 cΛ 2cLB
(18.439)
Fig. 18.8. The DC SQUID. The inductor LB is added to model current biasing.
where
φs Ic
EJ = . (18.448)
2πc
The corresponding Euler - Lagrange equation (1.8)
d ∂L ∂L
= , (18.449)
dt ∂ Φ̇ ∂Φ
which yields
! "
φs Φ̈ φs Φ̈
CJ + Ic sin Φ = Cg V̇g − , (18.450)
2πc 2πc
7. With the help of Eqs. (16.102) and (18.284) one finds that the density of
states D (ǫ) per unit volume (volume is labeled by V) is given by [compare
with Eq. (16.103)]
1
D (ǫ) = δ (ǫ − ηk′ )
V
k′
∞ .
1 2V
= 4π dk k δ ǫ − (ǫk′ − ǫF )2 + |∆|2
′ ′2
.
V 8π3
0
(18.459)
Assuming that the energy ǫk′ of an electron having wave vector k′ is
given by [see Eq. (16.97)]
2 ′2
k
ǫk′ = , (18.460)
2m
one finds that
∞ 0 .
ǫ′
D (ǫ) = DF dǫ ′
δ ǫ− (ǫ′ − ǫF )2 + |∆|2
ǫF
0
.
1/2
ǫ2 − |∆|2 ǫ
= DF 1 + . ,
ǫF
ǫ2 − |∆|2
(18.461)
where
21/2 m3/2 √
DF = ǫF (18.462)
π2 3
is the normal phase density of states per unit volume at the Fermi energy,
which is labeled by ǫF . For the case where ǫ ≪ ǫF and |∆| ≪ ǫF one has
ǫ
D (ǫ) = DF . . (18.463)
ǫ2 − |∆|2
8. The time evolution of the operators bk,↑ (t) and b†−k,↓ (t) is governed by
[see Eqs. (4.37) and (18.291)]
dbk,↑ 3 4
= −i −1 ηk′ bk,↑ , b†k′ ,σ bk′ ,σ , (18.464)
dt ′ k ,σ
db†−k,↓ 3 4
= −i −1
ηk′ b†−k,↓ , b†k′ ,σ bk′ ,σ , (18.465)
dt
k′ ,σ
thus
dbk,↑ 3 4
= −i −1 ηk bk,↑ , b†k,↑ bk,↑ , (18.466)
dt
db†−k,↓ 3 4
= −i −1 η−k b†−k,↓ , b†−k,↓ b−k,↓ . (18.467)
dt
With the help of the identity (16.69) one finds that
dbk,↑
= −i −1 ηk bk,↑ , (18.468)
dt
db†−k,↓
= i −1 η−k b†−k,↓+ , (18.469)
dt
thus
! −1
"
bk,↑ (t) e−i ηk t
0 bk,↑ (0)
= . (18.470)
b†−k,↓ (t) b†−k,↓ (0)
−1
i ηk t
0 e
(18.473)
9. Recall that the commutation relation (18.396) implies that the eigen-
values of the number operator B † B are the non-negative integers (see
chapter 5).
a) By assuming that the commutation relation (18.396) holds one finds
that [see Eqs. (18.393), (18.394) and (18.395)]
, - 1/2
[Σz , Σ+ ] = 2 B † B, B † N − B † B
1/2
= 2B † N − B † B
= 2Σ+ ,
(18.474)
1/2 , -
[Σz , Σ− ] = 2 N − B † B B † B, B
1/2
= −2 N − B † B B
= −2Σ− ,
(18.475)
and
[Σ+ , Σ− ]
1/2 1/2
= B† N − B† B B − N − B†B BB † N − B † B
, - 1/2 , - 1/2
= B † N − B † , B + BB † B − N − B † B B, B † + B † B N − B†B
= −N + 2B † B
= Σz ,
(18.476)
thus the commutation relations (18.180), (18.181) and (18.182) hold.
b) Note that the commutation relations (18.179) and (18.396) imply
that the operators a and b satisfy the same relations [see Eqs. (18.397)
and ,(18.398)]
-
a, a† = 1 , (18.477)
, †-
b, b = 1 . (18.478)
In terms of the operators a and b the Hamiltonian (18.390) is given
by [see Eqs. (18.393), (18.394), (18.395), (18.397) and (18.398)]
−1 1
HD = ω e α∗ + a† (α + a) +
2
ωa , -
+ −N + 2 β ∗ + b† (β + b)
2, -
+ gs α∗ + a† JD (β + b) + (α + a) β ∗ + b† JD ,
(18.479)
where JD is given by
, -1/2
JD = N − β ∗ + b† (β + b) . (18.480)
−1 ωe N ωa
HD = ω e a† a + ω a b† b + geff a† b + ab† + − ,
2 2
(18.481)
where
or in a matrix form
−1 a
HD = a† b† M
b
ωe N ωa
+ − ,
2 2
(18.483)
ω e geff
M= . (18.484)
geff ω a
ωe N ωa
En+ ,n− = n+ ω+ + n− ω − + − , (18.485)
2 2
action, 1 eigenvector, 23
adiabatic approximation, 421 equipartition theorem, 522
Aharonov-Bohm effect, 398 Euler-Lagrange equations, 2
angular momentum, 153 expectation value, 29