Chang K. Mathematical Structures of Quantum Mechanics 2011
Chang K. Mathematical Structures of Quantum Mechanics 2011
STRUCTURES
of
QUANTUM
MECHANICS
QUANTUM
MECHANICS
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI
For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.
ISBN-13 978-981-4366-58-8
ISBN-10 981-4366-58-7
Printed in Singapore.
During the past few years, after a couple of weeks of lecturing the
course of quantum mechanics that I offered at the Physics Department,
National Taiwan University, some students would usually come to ask
me as to what extent they had to refurbish their mathematical back-
ground in order to follow my lecture with ease and confidence. It was
hard for me to provide a decent and proper answer to the question, and
very often students would show reluctance to invest extra time on sub-
jects such as group theory or functional analysis when I advised them
to take some advanced mathematics courses. All these experiences that
I have encountered in my class eventually motivated me to write this
book.
The book is designed with the hope that it might be helpful to those
students I mentioned above. It could also serve as a complementary text
in quantum mechanics for students of inquiring minds who appreciate
the rigor and beauty of quantum theory.
Taipei, Taiwan
March, 2011 Kow Lung Chang
vii
This page intentionally left blank
Contents
ix
x Contents
1.6.2 Intrinsic compatibility of the dynamical observables and the
direct product space . . . . . . . . . . . . . . . . . . . . . . . 19
1.6.3 3rd postulate of quantum mechanics and commutator algebra 20
1.7 Non-commuting operators and the uncertainty principle . . . . . . 22
1.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Bibliography 189
Index 191
Chapter 1
1
2 Chapter 1
ψ + φ = φ + ψ, (1.1a)
ψ + (φ + χ) = (ψ + φ) + χ, (1.1b)
ψ + 0 = ψ, 0 is null vector, (1.1c)
a(ψ + φ) = aψ + aφ, (1.1d)
(a + b)ψ = aψ + bψ, (1.1e)
a(bψ) = (ab)ψ, (1.1f)
1 · ψ = ψ, (1.1g)
0 · ψ = 0, (1.1h)
where if a, b are real numbers, we call this vector space the real vector
space, and denote it by Vr . On the other way, complex vector space Vc
means a, b are complex numbers.
Example
We take n-dimensional Euclidean space, Rn -space, as an exmple. It
is a vector space with the vectors ψ and φ specified as ψ = (x1 , x2 , . . . ,
xi , . . . , xn ) and φ = (y1 , y2 , . . . , yi , . . . , yn ), where xi and yi (i = 1, 2, . . . ,
n) are all taken as real numbers. The sum of ψ and φ becomes (x1 +
y1 , x2 +y2 , . . . , xi +yi , . . . , xn +yn ) and aψ = (ax1 , ax2 , . . . , axi , . . . , axn ).
If a and xi are taken as complex numbers, then ψ is a vector in C n -space;
a n-dimensional complex vector space.
It is easily understood that a set of the continuous functions f (x)
for a 6 x 6 b forms a vector space, namely L2 (a, b)-space.
Before leaving this section, we also introduce some terminologies
in the following subsections that will be frequently referred to in later
chapters.
m
X
ai ψi . (1.2)
i=1
1.2 Inner product 3
n
X
(ψ, φ) = x∗i yi = x∗1 y1 + x∗2 y2 + · · · + x∗n yn . (1.4)
i=1
Proof
Since (ψ + αφ, ψ + αφ) > 0, where α = ξ + iη is a complex number.
Regard this inner product (ψ + αφ, ψ + αφ) = f (ξ, η) as a function of
two variables ξ and η. Then
f (ξ, η) = (ψ, ψ) + |α|2 (φ, φ) + α(ψ, φ) + α∗ (φ, ψ), (1.6)
∂f (ξ, η) ∂f (ξ, η)
= = 0, (1.7)
∂ξ ∂η
ξ0 ,η0 ξ0 ,η0
and we obtain
Therefore
(ψ, φ)(φ, ψ)
f (ξ0 , η0 ) = (ψ, ψ) − > 0, (1.9)
(φ, φ)
and we obtain α = −(ψ2 , ψ10 )∗ /(ψ10 , ψ10 ) = −(ψ10 , ψ2 )/(ψ10 , ψ10 ), hence
(ψ10 , ψ2 )
ψ20 = ψ2 − ψ10 . (1.11)
(ψ10 , ψ10 )
0 ,ψ )
(ψi−1 0 ,ψ )
(ψi−2 0
i i 0 (ψ1 , ψi )
ψi0 = ψi − ψi−1
0
0 , ψ0 ) − ψ 0
i−2 0 , ψ0 ) − · · · − ψ1 .
(ψi−1 i−1 (ψi−2 i−2 (ψ10 , ψ10 )
(1.12)
The set of orthogonal basis {ψ10 , ψ20 , . . . , ψn0 } can be normalized im-
mediately by multiplying the inverse square root of the corresponding
inner product, i.e.
ψi0
ψ̃i = p , (1.13)
(ψi0 , ψi0 )
and {ψ˜1 , ψ˜2 , . . . , ψ˜n } becomes the orthonormal set of the basis in the
vector space. From now on we shall take the basis to be orthonormal
without mentioning it particularly.
6 Chapter 1
Example
Consider the following set of continuous functions in C(−∞, ∞)
2
x
fn (x) = xn exp − , n = 0, 1, . . . . (1.14)
2
f 0 (f 0 , f1 ) x2
f10 (x) = f1 − 0 0 0 0 = f1 (x) = x exp , (1.16)
(f0 , f0 ) 2
f 0 (f 0 , f2 ) f 0 (f 0 , f2 )
2
1 x
f20 (x) = f2 − 1 0 1 0 − 0 0 0 0 = 2
x − exp − . (1.17)
(f1 , f1 ) (f0 , f0 ) 2 2
Similarly we have f30 (x) = (x3 − 3x/2) exp(−x2 /2). The orthonormal
functions can be calculated according to
fn0 (x)
2
1 x
f˜n (x) = p =p √ exp − Hn (x), (1.18)
0 0
(fn (x), fn (x)) n
2 n! π 2
where Hn (x) are called Hermite polynomials. One also recognizes that
f˜n (x) are in fact, the eigenfunctions of the Schrödinger equation for
one-dimension harmonic oscillation.
1.3.1 Norm
A norm on a vector space is a non-negative real function such that,
if ψ, φ are vectors, the norm of ψ is written as kψk, satisfying:
1.3 Completeness and Hilbert space 7
Example
Z b
2
kf (x)k = (f (x), f (x)) = |f (x)|2 dx.
a
lim ψn = lim ψm ,
n→∞ m→∞
kAψk
kAk = sup , ψ 6= 0 , (1.21)
kψk
Proof
k(A + B)ψk kAψ + Bψk
kA + Bk = sup , ψ 6= 0 = sup , ψ 6= 0
kψk kψk
kAψk kBψk
6 sup , ψ 6= 0 + sup , ψ 6= 0 = kAk + kBk.
kψk kψk
(Uχ, Uχ) = (ψ + φ, ψ + φ)
= (χ, χ) = (ψ, ψ) + (ψ, φ) + (φ, ψ) + (φ, φ)
= kψk2 + kφk2 + 2<{(ψ, φ)}.
Since kUψk = kψk, kUφk = kφk, we have <{(Uψ, Uφ)} = <{(ψ, φ)}.
Similarly if χ0 = ψ + iφ, we obtain (Uχ0 , Uχ0 ) = (χ0 , χ0 ), that implies
={(Uψ, Uφ)} = ={(ψ, φ)}, therefore (Uψ, Uφ) = (ψ, φ).
1.4 Linear operator 11
Consider the inner product of (ψ, Aφ) where A is a given linear op-
erator of interest. This numerically scalar quantity certainly is a func-
tion of operator A and the pair of vectors ψ and φ, namely (ψ, Aφ) =
F (A, ψ, φ) is a scalar quantity.
Instead of performing the above inner product straightforwardly, we
shall obtain the very same scalar of (ψ, Aφ) by forming the following
inner product (A† ψ, φ) such that (ψ, Aφ) ≡ (A† ψ, φ). The operator A†
is called the adjoint operator of A. The following relations can be
easily established (proofs left to readers):
(A + B)† = A† + B† , (1.23a)
(αA)† = α∗ A† , (1.23b)
(AB)† = B† A† , (1.23c)
(A† )† = A, (1.23d)
PM + PM⊥ = I.
Proposition 1.
The eigenvalues for a Hermitian operator are all real.
Let Aψa = aψa and A† ψa = aψa , and consider the inner product
hAiψa = (ψa , Aψa ) = (ψa , aψa ) = a(ψa , ψa ) = a. On the other hand, we
have hAiψa = (A† ψa , ψa ) = (Aψa , ψa ) = a∗ (ψa , ψa ) = a∗ which implies
a = a∗ if ψa is not a null vector.
Proposition 2.
Two eigenvectors of a Hermitian operator are orthogonal to each
other if the corresponding eigenvalues are unequal.
Proposition 3.
If the eigenvalue c of the operator C is degenerate, any linear
combination of the linearly independent eigenvectors is also an
eigenvector.
P
Due to the linearity of operator C, the linear combination αi ψci is
also the eigenvector of C with the eigenvalue c. By means of the Gram-
Schmidt orthogonalization process, one is able to easily construct a new
set of orthonormal vectors {ψ̃c1 , ψ̃c2 , . . . , ψ̃cm } out of the previous m
linearly independent set {ψc1 , ψc2 , . . . , ψcm }, such that
Proposition 4.
If two observables are compatible, their corresponding operators
R and S commute, i.e. [R, S] = 0.
which lead to
(RS − SR)ψa = [R, S]ψa = 0 ,
Proposition 5.
If R and S are operators corresponding to two compatible ob-
servables, and if ψr are eigenvectors of R, then
(ψr , Sψr0 ) = 0, for r 6= r0 . (1.29)
Proposition 6.
If Pr is the projection operator onto subspace with vectors ψr ,
then
[Pr , S] = 0.
we have
[Pr , S] = 0. (1.31)
Proposition 7.
If R and S are two commuting Hermitian operators, there exists
a complete set of states which are simultaneously eigenvectors of
R and S.
(s)
Let us construct a vector φr , projected by Ps upon the vector ψr
which is the eigenvector of the Hermitian operator R, i.e.
φ(s)
r = Ps ψr . (1.32)
(s)
It is obvious that φr is automatically the eigenvector of the Her-
mitian operator S with eigenvalue s, namely
Sφ(s) (s)
r = sφr . (1.33)
(s)
On the other hand, Proposition 6 ensures that φr is also an eigen-
vector of the operator R , because
Rφ(s) (s)
r = RPs ψr = Ps Rψr = rPs ψr = rφr . (1.34)
Proposition 8.
Let A and B be two operators corresponding respectively to two
intrinsic compatible observables, and let ψai and ϕbj be the eigen-
vectors of A and B with the eigenvalues ai and bj respectively,
then the direct product of ψai and ϕbj , denoted by ψai ⊗ ϕbj is the
eigenvector of the operator F(A, B) with the eigenvalue F (ai , bj ).
then we have
1 2 1
eA = I + A + A + A3 + . . . . (1.37)
2! 3!
1 1
eA Be−A ≡ B + [A, B] + [A, [A, B]] + . . . . (1.38)
1! 2!
Proof
Let f (λ) = eλA Be−λA and expand the function f (λ) in terms of
power series of λ at λ = 0, i.e.
λ 0 λ2
f (λ) = f (0) + f (0) + f 00 (0) + . . . .
1! 2!
Since,
f (0) = B,
f 0 (0) = [A, B],
f 00 (0) = [A, [A, B]],
and thus
λ 0 λ2
f (λ) = f (0) + f (0) + f 00 (0) + . . .
1! 2!
22 Chapter 1
becomes
λ2
f (λ) = eλA Be−λA = B + λ[A, B] + [A, [A, B]] + . . . .
2!
1
f (1) = eA Be−A ≡ B + [A, B] + [A, [A, B]] + . . . . (1.39)
2!
where ψ is the quantum state of interest, hQiI and hPiI are both con-
stant operators obtained by means of multiplying respectively the ex-
pectation value hQi and hPi by a unit operator.
Let us introduce deviation operators Qd and Pd as follows:
1.7 Non-commuting operators and the uncertainty principle 23
Qd = Q − hQiI, (1.42)
Pd = P − hPiI, (1.43)
and denote
φ1 = Qd ψ, (∆q)2 = (φ1 , φ1 ), (1.44)
1 1 i 1
(ψ, Qd Pd ψ) = (ψ, [Qd , Pd ]ψ) + (ψ, {Qd , Pd }) = ~ + α. (1.49)
2 2 2 2
~2 α2 ~2 1
(∆q)2 (∆p)2 > + > , or ∆q∆p > ~. (1.50)
4 4 4 2
(b) α = 0.
24 Chapter 1
1
λ(∆q)2 + (∆p)2 = 0.
λ
i~
λ= . (1.51)
2(∆q)2
1.8 Exercises
Ex 1.8.1
Ex 1.8.2
kψ + φk 6 kψk + kφk.
Ex 1.8.3
Prove the law of parallelogram holds in Hilbert space, i.e.
kψ + φk2 + kψ − φk2 = 2(kψk2 + kφk2 ).
Ex 1.8.4
Prove that every finite dimensional vector space is complete. (Hint:
since the real and complex numbers are complete.)
Ex 1.8.5
Show that the kABk 6 kAkkBk if A, B ∈ bounded operator.
Ex 1.8.6
Show that the expectation value of any dynamical observable in the
physical system is always real.
Ex 1.8.7
Prove that the adjoint conjugate can equivalently be defined as
(Aψ, φ) = (ψ, A† φ) if (A† ψ, φ) = (ψ, Aφ).
Ex 1.8.8
Let {φ1 , φ2 , . . . , φn } be a set of an orthonormal basis. Prove that
operator U is unitary if {Uφ1 , Uφ2 , . . . , Uφn } is also a set of an or-
thonormal basis.
Ex 1.8.9
Prove that if A and B are two operators that both commute with
their commutator, then
eA eB = eA+B+[A,B]/2 .
df (λ)
(Hint: let f (λ) = eλA eλB e−λ(A+B) , and obtain dλ = λ[A, B]f (λ)
by Eq. (1.39). Then integrate it.)
26 Chapter 1
Ex 1.8.10
Consider a Hilbert space spanned by a Hermitian operator A.
Y
(a) Prove that (A − aI) is a null operator if Aψa = aψa .
a
Y A − aI
(b) What is the significance of the following operator, ?
0
a0 − a
a6=a
Ex 1.8.11
If two observables A1 and A2 are not compatible, but their corre-
sponding operators both commute with the Hamiltonian operator H,
i.e.
[A1 , H] = [A2 , H] = 0.
Show that the energy eigenstates are in general degenerate.
Ex 1.8.12
Consider a one-dimensional Hamiltonian
1 2
H= P + V (Q)
2m
and use the fact that the commutator of Q and [Q, H] is a constant
operator, to show that
X ~2
(Ek − Es )|Qsk |2 = ,
2m
k
x1
x2
..
.
x̄ = xi
, (2.1)
xi+1
.
.
.
xn
24:22.
28 Chapter 2
0
0
..
.
e¯i =
1 ,
(2.2)
0
..
.
0
that allows one to express the vector x̄ as the following linear combina-
tion
n
X
x̄ = xi e¯i . (2.3)
i=1
which is exactly the adjoint conjugate, also called the Hermitian con-
jugate of vector x̄ in C n -space, namely by taking the complex conjugate
of each entry and transposing the column matrix into a row matrix in
∗ C n -space. One justifies immediately that ∗ C n is also a vector space.
24:22.
2.1 Vector space and dual vector space 29
y1
y2
x̃ȳ = x∗1 x∗2 ∗
. . . xn .
. (2.4)
..
yn
on the left hand side of Eq. (2.5), while on the right hand side of
Eq. (2.5), we have merely a complex number, because its adjoint conju-
gate is just the complex conjugate. Hence we have the following relation
24:22.
30 Chapter 2
or more specifically,
0 0 ... 0 0 ... 00
.
0 . . ..
0
.
.. .. . ..
. . .
. .
Pi = 1 0 0 . . . 1 0 . . . 0 = 0 .
1 0
0
.
0 0 0
. . . ..
. . . . .
.
0 0 0 ... 0 0 ... 0
n
X
It is obvious that the sum of the projection operators Pi over
the basis in the entire space is an identity operator, i.e. i=1
n
X
Pi = I. (2.12)
i=1
24:22.
2.1 Vector space and dual vector space 31
Py = ȳ ỹ. (2.13)
24:22.
32 Chapter 2
The set of eigenvectors {|a1 i, |a2 i, . . . , |ai i, . . .} forms the basis of the
Hilbert space H.
The projection operator onto the subspace |αi is designed by putting
a ket vector and a bra vector together in successive order, i.e.
Pα = |αihα|. (2.21)
Projection of the vector onto the subspace spanned by the basis |ai i
is then read as
24:22.
2.2 q-representation and p-representation in quantum mechanics 33
For the case that the eigenvalues become continuous, the eigenvec-
tors |ξi are no longer denumerable. The basis vector is characterized
by the eigenvalue ξ, which is a continuum and the inner product is
normalized to be a delta function, i.e.
Z Z
Pξ dξ = |ξidξhξ| = I. (2.29)
24:22.
34 Chapter 2
where
ξ = real parameter, namely the distance of the spatial translation,
P = momentum operator,
Proposition 1.
U(P; ξ)|xi is an eigenvector of operator X with eigenvalue x + ξ.
and
" #
X 1 i m
[X, U] = i~ X, − ξ Pm
m
m! ~
i m−1 m−1
X 1
=ξ − ξ P = ξU, (2.33)
m
(m − 1)! ~
or
XU = U(X + ξI). (2.34)
24:22.
2.2 q-representation and p-representation in quantum mechanics 35
|αi = U|xi.
Since
Proposition 2.
A state function or wave function as it is generally called, is
defined as the inner product of the vector |xi and the state vector
|ψi of interest, i.e. ψ(x) = hx|ψi.
i
U(P; ξ) = 1 − ξP,
~
and
24:22.
36 Chapter 2
i
hx|U(P; ξ)|x i = hx|x + ξi = hx| 1 − ξP |x0 i,
0 0
~
i
δ(x − x0 ) − δ(x − x0 − ξ) = ξhx|P|x0 i,
~
~ d ~
hx|P|x0 i = δ(x − x0 ) = δ 0 (x − x0 ),
i dx i
d d
δ 0 (x − x0 ) = δ(x − x0 ) = δ(x − x0 ),
dx d(x − x0 )
Z
f (x)δ 0 (x − a)dx = −f 0 (a) (2.36)
for any function f (x). With this property, we are able to derive that
Z Z
0 0 0 ~ 0
hx|P|ψi = hx|P|x idx hx |ψi = δ (x − x0 )ψ(x0 )dx0
i
Z
~ d ~ d
=− 0
δ(x0 − x)ψ(x0 )dx0 = ψ(x). (2.37)
i dx i dx
24:22.
2.2 q-representation and p-representation in quantum mechanics 37
ψ(x) in the above table is called the wave function or the state
function, and the absolute square of ψ(x), i.e. |ψ(x)|2 is commonly
referred to as the probability density of the system.
We particularly consider the eigenvalue equation for the energy of
the system. The Hamiltonian operator reads
1
H= P + V (X), (2.38)
2m
~2 d2
− + V (x) ψ(x) = Eψ(x). (2.41)
2m dx2
24:22.
38 Chapter 2
Proposition 3.
Q-representation of the eigenstate of the momentum operator is
a plane wave, namely
1 i
hx|pi = √ e ~ px
2π~
where |pi is the eigenstate of P, i.e,with the eigenvalue p, i.e.
with P|pi = p|pi.
or
~ d
phx|pi = hx|pi.
i dx
i 1 i
hx|pi = Ae ~ px = √ e ~ px . (2.42)
2π~
Z
i 0
= |A| 2
e ~ p(x−x ) dp = 2π~|A|2 δ(x − x0 ).
24:22.
2.2 q-representation and p-representation in quantum mechanics 39
Z Z
1 i 0
ψ(p) = hp|ψi = hp|x0 idx0 hx0 |ψi = √ e ~ px dx0 ψ(x0 ). (2.44)
2π~
1 2
hp|H|ψi = hp| P + V (X)|ψi. (2.45)
2m
1 2 1 2
hp| P |ψi = p ψ(p) (2.46)
2m 2m
Z
hp|V (X)|ψi = hp|V (X)|p0 idp0 hp0 |ψi,
where
ZZ
0
hp|V (X)|p i = hp|x0 idx0 hx0 |V (X)|x00 idx00 hx00 |p0 i
ZZ
1 i 0 00 −px0 )
= e ~ (p x V (x0 )δ(x0 − x00 )dx0 dx00
2π~
Z
1 i 0
= e ~ (p −p)x V (x)dx = Ṽ (p − p0 ), (2.47)
2π~
24:22.
40 Chapter 2
p2
Z
ψ(p) + Ṽ (p − p0 )ψ(p0 )dp0 = Eψ(p). (2.48)
2m
i~
Pd |ψi = Qd |ψi. (2.49)
2(∆q)2
~ ∂ i~
− hpi ψ(x) = (x − hxi)ψ(x), (2.50)
i ∂x 2(∆x)2
(x − hxi)2 ihpix
1
ψ(x) = exp + . (2.51)
[2π(∆x)2 ]1/4 4(∆x)2 ~
There exists very rich structure in dealing with the harmonic oscil-
lation mathematically on a quantum level. We shall attack the problem
by means of the number operator through the creation operator and
the annihilation operator rather than solving the Schrödinger equa-
tion in the q-representation, in which the quantization of the energy
levels is derived from the boundary conditions of the wave functions.
24:22.
2.3 Harmonic oscillator revisited 41
The technique developed in this section can even be applied to the sys-
tem of linear oscillators in the limit of continuum, which gives rise to
the quantization of the field.
1 2 1
H= P + mω 2 X2 . (2.52)
2m 2
r
1 mω
pξ = √ P, ξ= X. (2.53)
mω~ ~
1 2 1 2
H = ~ω p + ξ = ~ωH, (2.54)
2 ξ 2
1 1
a = √ (ξ + ipξ ), and a† = √ (ξ − ipξ ), (2.56)
2 2
or reversely
24:22.
42 Chapter 2
1
ξ = √ (a† + a), (2.57)
2
i
pξ = √ (a† − a). (2.58)
2
Proposition 4.
If ν and |νi are respectively the eigenvalue and eigenvector of
number operator N , then ν is positive definite.
24:22.
2.3 Harmonic oscillator revisited 43
Proposition 5.
The state with zero eigenvalue i.e. |0i, then |0i will be annihi-
lated into the null vector by a, the annihilation operator, namely
a|0i = 0 .
Proposition 6.
a† |νi is an eigenvector of N with eigenvalue ν + 1.
[N , a† ] = a† , or N a† = a† (N + I). (2.62)
Proposition 7.
For |νi 6= 0 , then a|νi is the eigenvector of N with eigenvalue
ν − 1.
24:22.
44 Chapter 2
Proposition 8.
The eigenvalues of the number operator N take only non-negative
integers, i.e. ν = n = 0, 1, 2, . . . .
so that the action of a† will increase the n-particle system into the
(n+1)-particle system or will promote the n-th excited state
√ into (n+1)-
th excited state. The constant c is determined to be n + 1 when all
eigenstates are orthonormalized, i.e.
√
a† |ni = n + 1|n + 1i. (2.65)
24:22.
2.3 Harmonic oscillator revisited 45
Similarly operator a acts upon state |ni and it becomes the state
c0 |n − 1i, i.e.
a|ni = c0 |n − 1i.
√
a|ni = n|n − 1i. (2.66)
With the help of the creation operator, one is able to construct the
orthonormal basis of the Hilbert space spanned by the eigenvectors of
the number operator N . We start from the ground state or zero particle
states and build the higher excited states or many particle state by
acting the creation operator upon it successively. Then we have
|0i,
|1i = a† |0i,
1
|2i = √ (a† )2 |0i,
2!
..
.
1
|ni = √ (a† )n |0i. (2.67)
n!
24:22.
46 Chapter 2
|0i = a|1i,
1
|1i = √ a|2i,
2
1
|2i = √ a|3i,
3
..
.
1
|ni = √ a|n + 1i. (2.68)
n+1
d
hξ|pξ |ξ 0 i = −i δ(ξ − ξ 0 ),
dξ
ψn (ξ) = hξ|ni.
24:22.
2.4 N-representation and the Rodrigues formula of Hermite polynomials 47
1
Since a† = √ (ξ − ipξ ), the wave function ψn (ξ) becomes
2
1 n
1 † n 1
ψn (ξ) = hξ|ni = √ hξ|(a ) |0i = √ √ hξ|(ξ − ipξ )n |0i.
n! n! 2
By Proposition 2, we have
d n d n
n
hξ|(ξ − ipξ ) |0i = ξ − hξ|0i = ξ − ψ0 (ξ).
dξ dξ
1
hξ| √ (ξ + ipξ )|0i = 0,
2
or
d
ξ+ ψ0 (ξ) = 0. (2.70)
dξ
Therefore we can solve the equation and obtain the normalized wave
function as
1 − 1 ξ2
ψ0 (ξ) = √
4
e 2 , (2.71)
π
24:22.
48 Chapter 2
d 1 2 d 1 2
ξ− ≡ (−1)e 2 ξ e− 2 ξ ,
dξ dξ
we conclude that
d n
n
n 21 ξ 2 d 1 2
ξ− = (−1) e e− 2 ξ .
dξ dξ
Then
n 1 2
(−1)n e− 2 ξ
1 2 d −ξ 2
ψn (ξ) = p √ e2ξ e =p √ Hn (ξ), (2.73)
2n n! π dξ 2n n! π
n
n ξ2 d 2
Hn (ξ) = (−1) e e−ξ . (2.74)
dξ
1 m
H= (P2 + P2y ) + ω(X2 + Y2 ). (2.75)
2m x 2
24:22.
2.5 Two-dimensional harmonic oscillation and direct product of vector spaces 49
1
a† = √ (ξ − ipξ ), (2.76a)
2
1
a = √ (ξ + ipξ ), (2.76b)
2
1
b† = √ (η − ipη ), (2.76c)
2
1
b = √ (η + ipη ), (2.76d)
2
24:22.
50 Chapter 2
or H = Ha + Hb if we put
† 1 1
Ha = a a + I ~ω = Na + I ~ω,
2 2
† 1 1
Hb = b b + I ~ω = Nb + I ~ω.
2 2
1 1
Na + I ~ω + Nb + I ~ω |ψi = E|ψi, (2.80)
2 2
H = Ha ⊗ H b . (2.83)
24:22.
2.5 Two-dimensional harmonic oscillation and direct product of vector spaces 51
24:22.
52 Chapter 2
R ⊗ O = O ⊗ S = O, (2.87a)
I ⊗ I = I, R ⊗ I = R, I ⊗ S = S, (2.87b)
R ⊗ (S1 + S2 ) = R ⊗ S1 + R ⊗ S2 , (2.87c)
(R1 + R2 ) ⊗ S = R1 ⊗ S + R2 ⊗ S, (2.87d)
αR ⊗ βS = αβR ⊗ S, (2.87e)
−1 −1 −1
(R ⊗ S) =R ⊗S , (2.87f)
R1 R2 ⊗ S1 S2 = (R1 ⊗ S1 )(R2 ⊗ S2 ), (2.87g)
~ y) = 1 H0 (−y î + xĵ).
A(x,
2
1 q 2
H= P− A
2m c
( 2 2 )
1 1q 1q
= Px + H0 Y + Py − H0 X . (2.88)
2m 2c 2c
The sum of the two square terms in the Hamiltonian of last equation
ensures that the total energy of the system is always positive definite.
Namely when we solve the following Schrödinger equation
24:22.
2.5 Two-dimensional harmonic oscillation and direct product of vector spaces 53
1 2 m ωc 2 2 1 2 m ωc 2 2 1
H= Px + X + P + Y − ωc Lz , (2.90)
2m 2 2 2m y 2 2 2
where ωc is the cyclotron frequency and Lz stands for the 3rd com-
ponent of the angular momentum operator which can be calculated
explicitly as follows:
qH0
ωc = , Lz = XPy − YPx .
mc
m ωc 2
oscillator with the force constant k = , i.e.
2 2
(2) 1 m ωc 2 2
Hh.o. = (P2x + P2y ) + (X + Y2 ). (2.91)
2m 2 2
(2)
It can be easily verified that Hh.o. and Lz commute with each other
i.e.
(2)
[Hh.o. , Lz ] = O. (2.92)
We will run into a dilemma if we apply the method of Proposition
8 to calculate the eigenvalues of the following Hamiltonian
(2) 1
H = Hh.o. − ωc Lz , (2.93)
2
24:22.
54 Chapter 2
(2) 1
Hh.o. |n1 ; n2 i = (n1 + n2 + 1)~ωc |n1 ; n2 i, (2.94)
2
where m takes the integers, while n1 and n2 take only the positive ones.
Of course we have made use of the result that the angular momentum Lz
is quantized as it is expressed in Eq. (2.95), which we shall investigate
in the next chapter.
Should we try to evaluate the total energy of the system by applying
(2)
H = Hh.o. −ωc Lz /2 to act upon the direct product vector |n1 ; n2 i⊗|mi,
we would end up in the following eigenvalue equation
(2) 1
Hh.o. − ωc Lz |n1 ; n2 i ⊗ |mi = E|n1 ; n2 i ⊗ |mi, (2.96)
2
(2)
The contradiction arises from the fact that the observable Hh.o. and
the observable Lz are not intrinsic compatible even though they com-
mute with each other. One needs, in fact, all the three fundamental
(2)
commutation relations in order to prove that the commutator of Hh.o.
and Lz vanishes, i.e.
(2)
[Hh.o. , Lz ] = O. (2.97)
24:22.
2.5 Two-dimensional harmonic oscillation and direct product of vector spaces 55
r r
1 mωc 2
c= √ (X + iY) + i (Px − iPy ) , (2.98a)
2 2~ m~ωc
r r
† 1 mωc 2
c =√ (X − iY) − i (Px + iPy ) , (2.98b)
2 2~ m~ωc
r r
1 mωc 2
d= √ (X − iY) + i (Px + iPy ) , (2.98c)
2 2~ m~ωc
r r
† 1 mωc 2
d =√ (X + iY) − i (Px − iPy ) . (2.98d)
2 2~ m~ωc
(2) 1
Hh.o. − ωc Lz |nc ; nd i = E|nc ; nd i. (2.100)
2
24:22.
56 Chapter 2
(2)
After some algebra, the two terms Hh.o. and Lz in the Hamiltonian
above can be expressed in terms of the new dynamical operators Nc
and Nd as follows:
(2) 1
Hh.o. = (Nc + Nd + 1)~ωc , (2.101)
2
1 1
− ωc Lz = (Nc − Nd )~ωc , (2.102)
2 2
N N N
1 X 2 αX 2 βX
H= pr + (qr − qr+1 ) + qr qr , (2.105)
2m 2 2
r=1 r=1 r=1
24:22.
2.6 Elastic medium and the quantization of scalar field 57
where qr is the displacement of the r-th mass point away from its equi-
librium position, and pr is the corresponding canonical conjugate mo-
mentum. The commutation relations are expressed as
r
~ X 2πi rs
qr = e N Qs . (2.106)
Nm s
(
takes from s = − 21 (N − 1) to s = 12 (N − 1),
s=
they are integers if N is odd, and half odd integer if N is even.
The condition that qr is real will lead us to take Q−s = Q†s . Before
we try to decouple the mixing terms in the Hamiltonian, we establish
the following formula:
24:22.
58 Chapter 2
Proposition 9.
For r = integers, and s − s0 = integers, then
N
1 X 2πi r(s−s0 )
eN = δss0 .
N
r=1
2πi
(s−s0 )
Denoting z = e N , we have
1 X r 1 z(1 − z N )
F (z) = z = (z + z 2 + . . . + z N ) = , (2.107)
N r N N (1 − z)
F (z) = F (1) = 1.
1
F (z) = (1 + z + . . . + z N −1 ) = 0,
N
r
m X − 2πi rs
Qs = e N qr . (2.109)
N~ r
24:22.
2.6 Elastic medium and the quantization of scalar field 59
r
m~ X 2πi rs
pr = e N Ps , (2.110)
N s
with
P−s = Ps† ,
and
1 X 2πi
Ps = √ e− N rs pr . (2.111)
mN ~ r
It also leaves you as an exercise to show that each term in the Hamil-
tonian operator can be calculated as follows,
1 X 2 ~X
p = Ps Ps† , (2.112)
2m r r 2 s
αX α~ X 2πs
(qr − qr+1 )2 = 4 sin2 Qs Q†s , (2.113)
2 r 2m s N
βX β~ X
qr qr = Qs Q†s . (2.114)
2 r 2m s
~X
H= [Ps Ps† + ω 2 (s)Qs Q†s ], (2.115)
2
2 4α 2πs β
ω (s) = sin2 + , (2.116)
m N m
24:22.
60 Chapter 2
ω(−s) = ω(s).
∂H
= Q̇†s , or Ps = Q̇†s ,
∂Ps†
∂H
= −Ṗs† , or ω 2 (s)Qs = −Ṗs† .
∂Q†s
Hence we have
1
Qs = p as e−iω(s)t + a†−s eiω(s)t , (2.119)
2ω(s)
where as / 2ω(s) and a†−s / 2ω(s), playing the roles of the constants
p p
r
ω(s)
Ps = i −as e−iω(s)t + a†−s eiω(s)t . (2.120)
2
24:22.
2.6 Elastic medium and the quantization of scalar field 61
Let us define bs = as e−iω(s)t and b†−s = a†−s eiω(s)t . Then the normal
coordinate and the normal momentum become as follows
1
Qs = p (bs + b†−s ),
2ω(s)
r
ω(s)
Ps = i (−bs + b†−s ),
2
~X X 1
H= ω(s)(bs b†s + b†s bs ) = ~ ω(s)(b†s bs + I)
2 s s
2
X 1
=~ ω(s)(Ns + I), (2.121)
s
2
l
xr = r . (2.122)
N
2π
ks = s, (2.123)
l
24:22.
62 Chapter 2
and rescale the displacement away from the r-th equilibrium position
at the time t defined as
r r r
mN mN ~ X 2πi rs
Φ(xr , t) = qr = e N Qs
l l mN s
r
~ X 2πi rs 1
= eN p as e−iω(s)t + a†−s eiω(s)t
l s 2ω(s)
r
~X 1
= p as ei(ks xr −ω(s)t) + a†s e−i(ks xr −ω(s)t) ,
l s 2ω(s)
(2.124)
r r
~ X ω(s)
Π(xr , t) = i −as ei(ks xr −ω(s)t) + a†s e−i(ks xr −ω(s)t) ,
l s 2
(2.125)
where the last term in Eqs. (2.124) and (2.125) are obtained by making
use of ks = −k−s and ω(s) = ω(−s).
The commutation relation between Φ(xr , t) and Π(xr0 , t) can be
evaluated as
~
[Φ(xr , t), Π(xr0 , t)] = N δrr0 I.
l
l
x = lim xr = lim r . (2.126)
N →∞ N →∞ N
24:22.
2.6 Elastic medium and the quantization of scalar field 63
r
~X 1
= p ak ei(kx−ω(k)t) + a†k e−i(kx−ω(k)t) , (2.127)
l 2ω(k)
k
r r
~X ω(k)
= i −ak ei(kx−ω(k)t) + a†k e−i(kx−ω(k)t) ,
l 2
k
(2.128)
N
= i~ lim δrr0 I = i~δ(x − x0 )I. (2.129)
N →∞ l
r
~X 1
Φ(r, t) = p ak ei(kr−ω(k)t) + a†k e−i(kr−ω(k)t) , (2.130)
l 2ω(k)
k
r r
~X ω(k)
Π(r, t) = i −ak ei(kr−ω(k)t) + a†k e−i(kr−ω(k)t) ,
l 2
k
(2.131)
24:22.
64 Chapter 2
The universe evolves in time and so does the physical system. Any
quantum state in the quantum mechanical system always changes in
time, just like the motion of the particles in classical mechanics changes
its configuration according to laws of motion. All the physical states as
well as the dynamical observables we have dealt with in the first three
postulates of quantum mechanics were expressed in such a way that
time is totally absent from the equations. In fact, the physical quantities
we had met or dealt with are time dependent implicitly, namely that
we set our observation of the system at a convenient time scale, say
t = 0. Choosing the time scale is entirely arbitrary . We would like to
express the physical state at the particular time t = t0 to perform the
observation or to make the measurement. Then the state will be denoted
by |ψ; t0 i. All the contents in our previous discussion do not change and
all the theories and conclusions we have derived so far remain the same
if we replace |ψi by |ψ; t0 i.
24:22.
2.7 Time evolution operator and the postulate of quantum dynamics 65
hψ; t|ψ; ti = hψ; t0 |U† (t, t0 )U(t, t0 )|ψ; t0 i = hψ; t0 |ψ; t0 i (2.135)
or
Z Z Z
2
hψ; t|xidxhx|ψ, ti = |ψ(x; t)| dx = |ψ(x; t0 )|2 dx.
Therefore
Since U(t, t) = I, the time evolution operator U(t + δt, t), of course
is very close to identity operator, can be approximated up to first order
in δt by the following expression
or Ω† = Ω.
It is the operator Ω that dictates the change of the state, and this
forms the last postulate of quantum mechanics, namely the law of quan-
tum dynamics, in which Ω is taken to be H(P, Q; t)/~.
24:22.
66 Chapter 2
In other words, the total energy of the system is the generator that
drives the quantum state into evolution. Therefore the Schrödinger
equation is referred to as the law of quantum dynamics.
Hence we have
i
U(t + δt, t0 ) = U(t + δt, t)U(t, t0 ) = I − H(P, Q; t) δt. (2.142)
~
∂U(t, t0 ) 1 i
= lim (U(t + δt, t0 ) − U(t, t0 )) = − H(P, Q; t)U(t, t0 ).
∂t δt→0 δt ~
(2.143)
24:22.
2.7 Time evolution operator and the postulate of quantum dynamics 67
The solution for U(t, t0 ) is trivial for the conserved system in which the
Hamiltonian operator does not depend upon time explicitly. U(t, t0 )
then takes the following expression
i
U(t, t0 ) = e− ~ H(P,Q)(t−t0 ) . (2.144)
Z t
i
U(t, t0 ) = I + − H(P, Q; t0 )U(t0 , t0 )dt0 ,
~ t0
Z t0
i 2 t 0
Z t Z
i 0 0 0
U(t, t0 ) = I+ − dt H(t )+ − dt H(t ) dt00 H(t00 )+. . . .
~ t0 ~ t0 t0
(
A(t1 )B(t2 ) if t1 > t2 ,
T (A(t1 )B(t2 )) =
B(t2 )A(t1 ) if t2 > t1 ,
then
Z t Z t0 ZZ t
1
dt0 H(t0 ) dt00 H(t00 ) = dt0 dt00 T (H(t0 )H(t00 )),
t0 t0 2! t0
24:22.
68 Chapter 2
i 2
Z t ZZ t
i 0 0 1
U(t, t0 ) = I + − dt H(t ) + − dt0 dt00 T (H(t0 )H(t00 ))
~ t0 2! ~ t0
3 Z Z Z t
1 i
+ − dt0 dt00 dt000 T (H(t0 )H(t00 )H(t000 )) + . . .
3! ~ t0
n i Rt o
− dtH(P,Q;t)
=T e ~ t0 . (2.145)
d d d
hAit = hψ; t|A|ψ; ti = hψ; t0 |U† (t, t0 )AU(t, t0 )|ψ; t0 i
dt dt dt
i i
= hψ; t0 |U† (HA − AH)U|ψ; t0 i = hψ; t|[H, A]|ψ; ti.
~ ~
(2.147)
24:22.
2.8 Schrödinger picture vs. Heisenberg picture 69
d i
hQit = hψ; t|[H, Q]|ψ; ti. (2.148)
dt ~
~ ∂H
[H, Q] = ,
i ∂P
hence
d ∂H
hQit = . (2.149)
dt ∂P t
d ∂H
hPit = − . (2.150)
dt ∂Q t
24:22.
70 Chapter 2
here for simplicity and clarity, we take t0 = 0 and put U(t, t0 ) = U(t).
Since the commutation relations are invariant under the unitary
transformation, the fundamental quantization rules remain the same
as before, provided that all new operators are of at equal time, i.e.
~
[P̃(t), Q̃(t)] = I, (2.152)
i
~ ∂F
[Q̃(t), F (P̃(t), Q̃(t))] = − , (2.153)
i ∂ P̃(t)
~ ∂G
[P̃(t), G(P̃(t), Q̃(t))] = . (2.154)
i ∂ Q̃(t)
∂ Q̃(t) i ∂H
= U† (t)[H, Q]U(t) = , (2.155)
∂t ~ ∂ P̃(t)
∂ P̃(t) i ∂H
= U† (t)[H, P]U(t) = − , (2.156)
∂t ~ ∂ Q̃(t)
24:22.
2.9 Propagator in quantum mechanics 71
∂ Q̃(t) ∂H ∂Hc
= , q̇c (t) = ,
∂t ∂ P̃(t) ∂pc (t)
∂ P̃(t) ∂H ∂Hc
=− , ṗc (t) = − .
∂t ∂ Q̃(t) ∂qc (t)
24:22.
72 Chapter 2
∂ X i 0
i~ K(x, t; x0 , t0 ) = Eα hx|αie− ~ Eα (t−t ) hα|x0 i
∂t α
X ~ ∂ i 0
= H , x hx|αie− ~ Eα (t−t ) hα|x0 i
α
i ∂x
~ ∂
=H , x K(x, t; x0 , t0 ). (2.160)
i ∂x
d
η(t − t0 ) = δ(t − t0 ),
dt
then the time derivative for the product of η(t − t0 ) and K(x, t; z 0 , t0 )
can be calculated as
∂
i~ (η(t − t0 )K(x, t; x0 , t)) = η(t − t0 )HK + i~δ(t − t0 )K
∂t
= η(t − t0 )HK + i~δ(t − t0 )δ(x − x0 ),
(2.161)
24:22.
2.10 Newtonian mechanics regained in the classical limit 73
where the last term is obtained by the property that δ(x − a)f (x) =
δ(x − a)f (a). Hence
∂
i~ − H η(t − t0 )K(x, t; x0 , t0 ) = i~δ(t − t0 )δ(x − x0 ). (2.162)
∂t
Z
i p2 0
0 0
K(x, t; x , t ) = hx|pie− ~ 2m (t−t ) hp|x0 idp
Z
1 i 0 i p2 0
= e ~ p(x−x ) e− ~ 2m (t−t ) dp
2π~
i m(x − x0 )2
r
m
= exp .
2πi~(t − t0 ) 2~ (t − t0 )
r
(x − a)2
1
δ(x − a) = lim exp − , (2.163)
→0 π
24:22.
74 Chapter 2
The time dependent state function and the Schrödinger equation we had
before can be re-expressed respectively as
i
ψ(x; t) = hx, t|ψi = hx|e− ~ Ht |ψi, (2.165)
∂ ∂ i
i~ ψ(x; t) = i~ hx|e− ~ Ht |ψi = Hhx, t|ψi = Hψ(x; t). (2.166)
∂t ∂t
Z
|x0 , t0 idx0 hx0 , t0 | = I,
that can be inserted in between the state hx, t| and the state |x0 , t0 i at
the instant t0 . If we consider the case that the transition takes place in
N steps from the initial spacetime (x0 , t0 ) to the final spacetime (x, t),
and let
24:22.
2.10 Newtonian mechanics regained in the classical limit 75
t − t0 = N , tn = t0 + n, where n = 1, 2, . . . , N − 1,
× hx1 , t1 |x0 , t0 i.
(2.168)
A typical term hxn , tn |xn−1 , tn−1 i in the above formula shall be ex-
plored in detail as follows
i i 1 2 +V
hxn , tn |xn−1 , tn−1 i = hxn |e− ~ H |xn−1 i = hxn |e− ~ ( 2m )P (X))
|xn−1 i
Z
i 2 i
= hxn |e− 2m~ P |ξidξhξ|e− ~ V (X) |xn−1 i. (2.169)
24:22.
76 Chapter 2
m(xn − ξ)2
Z r
1 i
− 2m~ p2 i
p(xn −ξ) m
= e e ~ dp = exp i .
2π~ 2πi~ 2~
The typical term hxn ; tn |xn−1 , tn−1 i can finally be cast into the ex-
pression,
( " #)
i m xn − xn−1 2
r
m
hxn ; tn |xn−1 , tn−1 i = exp − V (xn−1 )
2πi~ ~ 2
r
m i
= exp Ln , (2.170)
2πi~ ~
2
m xn − xn−1
Ln = − V (xn−1 ),
2
m N/2 Z Y
i
hx, t|x0 , t0 i = dxn exp Ln .
2πi~ ~
24:22.
2.10 Newtonian mechanics regained in the classical limit 77
m N/2 Z Y
i
hx, t|x0 , t0 i = lim dxn exp Ln
→0 2πi~ ~
Z Z t
i
= Dx(τ ) exp L(x(τ ), ẋ(τ ))dτ , (2.171)
~ t0
where and xn (or xn−1 ) are replaced by dτ and x(τ ) respectively, and
m xn − xn−1
L(x(τ ), ẋ(τ )) = lim Ln = lim − V (xn−1 ) .
→0 →0 2
Z t
δS = δ L(x(τ ), ẋ(τ ))dτ, (2.172)
t0
d ∂L ∂L
− = 0, (2.173)
dτ ∂ ẋ(τ ) ∂x(τ )
24:22.
78 Chapter 2
2.11 Exercises
Ex 2.11.1
Show that delta function δ(x) can be expressed as
1 sin N x
(a) δ(x) = lim or
π N →∞ x
1 d2
(b) δ(x) = |x|.
2 dx2
Ex 2.11.2
Show that
2
~0
2
hx|P |x i = δ 00 (x − x0 ),
i
0 ~ d
hx|F (P)|x i = F δ(x − x0 ).
i dx
Ex 2.11.3
Consider a physical system of one dimension that is translated by a
distance ξ to the right, the wave function then becomes
hx|U(P; ξ)|ψi.
24:22.
2.11 Exercises 79
Ex 2.11.4
p2
~ d
ψ(p) + V − ψ(p) = Eψ(p).
2m i dp
Ex 2.11.5
Ex 2.11.6
Ex 2.11.7
Ex 2.11.8
2
Show that h0|eikx |0i = exp(− k2 h0|X2 |0i), where |0i is the ground
state of the one dimensional harmonic oscillator, and X is the position
operator.
Ex 2.11.9
Prove that
24:22.
80 Chapter 2
1 X ~X
pr pr = Ps Ps† ,
2m 2 s
βX β~ X
qr qr = Qs Q†s ,
2 2m s
if we define
r
~ X 2πi rs
pr = e N Ps ,
Nm s
r
m~ X 2πi rs
qr = e N Qs .
N s
Ex 2.11.10
Prove that
αX α~ X 2πs
(qr − qr+1 )2 = 4 sin2 Qs Q†s .
2 r 2m s N
Ex 2.11.11
" #
1 x2 + y 2 − 2xyξ 2 +y 2 )
X ξn
exp − = e−(x Hn (x)Hn (y),
2n n!
p p
1 − ξ2 1 − ξ2 n
24:22.
2.11 Exercises 81
prove that the propagator in the case of the one dimensional harmonic
oscillator can be expressed as (Hint: use Eq. (2.159) for |αi = |ni.),
r
mω
K(x, t; x0 , t0 ) =
2πi~ sin ω(t − t0 )
imω 2
× exp [(x + x20 ) cos ω(t − t0 ) − 2xx0 ] .
2~ sin ω(t − t0 )
Ex 2.11.12
For a dynamical observable A(t) which is time dependent explicitly,
the time derivative of the expectation value takes the form as
d D ∂A E
i~ hA(t)it = h[A, H]it + i~ .
dt ∂t t
hpi2
2 1
h∆xi2t = h∆xi20 + hxp + pxi0 − hxi0 hpi0 t + 20 t2 ,
m 2 m
Ex 2.11.13
Prove that the one dimensional Schrödinger equation in free space
∂ ~2 ∂ 2
i~ ψ(x, t) = − ψ(x, t)
∂t 2m ∂x2
24:22.
82 Chapter 2
Ex 2.11.14
Derive explicitly the propagator in the p-representation, i.e.
Z Z t
− ~i H(t−t0 ) 0 i
hx|e |x i = DxDp exp L̄(τ )dτ ,
~ t0
where
Y dxn dpn Z t
DxDp = lim , L̄(τ )dτ = lim L̄n ,
→0
n
2π~ t0 →0
with
L̄n = pn ẋn − H(xn , pn ),
1
ẋn = (xn − xn−1 ).
24:22.
Chapter 3
25:23.
84 Chapter 3
Further translation from the vector x0 to the vector x00 by another dis-
placement η which is characterized by another set of n continuous pa-
rameters η = (η 1 , η 2 , . . . , η n ) is written as
x00 = x0 + η or x00i = x0i + η i = xi + (ξ i + η i ). (3.2)
and the composition rules for the group parameters are just the addition,
i.e.
γ = (γ 1 , γ 2 , . . . , γ n ), γ i = ξi + ηi. (3.3)
ξ −1 · ξ = ξ · ξ −1 = e,
that implies
ξ −1 = (−ξ 1 , −ξ 2 , . . . , −ξ n ). (3.4)
25:23.
3.1 Symmetry and transformation 85
∂ 1 i j ∂ ∂ ξi ∂ i
= F (x) + ξ i F (X) + ξ ξ F (x) + . . . = e ∂x F (x).
∂xi 2! ∂xi ∂xj
(3.5)
One notices that the group elements for translation take a completely
different form.
The group element of the displacement is written as
exp ξ i ∂/∂xi instead of ξ. Therefore the group operation is the usual
multiplication, namely the successive displacementsof ξ and η is written
as the product of two group elements exp η i ∂/∂xi and exp ξ i ∂/∂xi ,
25:23.
86 Chapter 3
The linear transformations form a group provided that all the trans-
formation matrices are of nonvanishing determinants. Consider two suc-
cessive transformations and x0 = Ax, x00 = Bx0 , then we conclude:
or
25:23.
3.2 Lie groups and Lie algebras 87
and hence
cl = f l (ai , bj ) = al + bl + fijl ai bj + O(3),
with
∂ 2 f l (e, e)
fijl = = constant coefficient. (3.10)
∂ai ∂bj
25:23.
88 Chapter 3
G(a)
x 7−→ x0 = h(x, a),
∂h(x0 , e)
dx0 = δa = u(x0 )δa,
∂(δa)
25:23.
3.2 Lie groups and Lie algebras 89
∂f (a, e) ∂f l (a, e) m
da = δa, or dal = l
δa = fm (a)δam , (3.11)
∂b ∂bm
−1
∂f (a, e)
δa = da = ψ(a)da, (3.12)
∂b
where ψ(a) again is a r×r square matrix, the inverse matrix of ∂f (a, e)/∂b.
Therefore
dx0 = u(x0 )ψ(a)da,
∂
Xi = um
i , (3.13)
∂xm
25:23.
90 Chapter 3
T (ξ)
x 7−→ x0 = x + ξ,
or expressed in components
x0m = xm + ξ m .
The generators of the 3-dimensional translational group can be calcu-
lated as follows
∂ ∂x0m ∂ ∂ ∂
Xi = um
i m
= i m
= δim m = ,
∂x ∂ξ ∂x ∂x ∂xi
∂ i
i
= Pi , (3.14)
∂x ~
i ∂ i ~~
which allows us to express U (P, −ξ) = eξ ∂xi = e ~ ξ·P as the group
element for a coordinate transformation to the right by xi → xi + ξ i ,
which corresponds to the transformation of the physical state to the left
by ξ i .
We take the rotation about the azimuthal axis as another example.
It is a one-parameter group defined by
x01 = x1 cos θ − x2 sin θ, (3.15)
25:23.
3.2 Lie groups and Lie algebras 91
∂x01
u11 = = −x2 , (3.17)
∂θ θ=0
∂x02
u21 = = x1 , (3.18)
∂θ θ=0
∂ ∂ ∂ i
X3 = ui1 i
= −x2 1 + x1 2 = L3 , (3.19)
∂x ∂x ∂x ~
∂ m ∂
∂um
j ∂ l
m ∂ui ∂
[Xi , Xj ] = [uli , u ] = ul
− u
∂xl j ∂xm i
∂xl ∂xm j
∂xm ∂xl
∂um m
l j l ∂ui ∂
= ui l
− uj l
. (3.20)
∂x ∂x ∂xm
25:23.
92 Chapter 3
∂um
j
m
l ∂ui
uli − uj = ckij um
k , (3.21)
∂xl ∂xl
∂xm
Since dxm = um k n
k ψn da , we have = um k
k ψn . With the property
∂an
of infinite differentiability of the transformation functions h(x, a) with
respect to the group parameters, namely that
∂ 2 xm ∂ 2 xm
= , (3.22)
∂al ∂an ∂an ∂al
we reach at
∂um
k ∂um
k l α
Putting = u ψ , the above equation takes the expression
∂an ∂xl α n
as
25:23.
3.2 Lie groups and Lie algebras 93
The term on the right hand side of the last equation depends upon
the coordinate only, while the left hand side is product of a coordinate
dependent function umk and a group parameter dependent factor
k k
∂ψl ∂ψn l n
∂an − ∂al fi fj . The condition can only be satisfied if,
∂ψlk ∂ψnk
− fil fjn = ckij = −ckji = constant coefficient.
∂an ∂al
Therefore we have,
∂um
j k ∂um
ui − i
uk = ckij um
k , (3.26)
∂xk ∂xk j
∂um m
j k ∂ui ∂ ∂
[Xi , Xj ] = uki − uj = ckij um
k = ckij Xk , (3.27)
∂xk ∂xk ∂xm ∂xm
and they are called Lie algebras. ckij are called the structure con-
stants.
(a) A Lie group is said to be Abelian if all the structure constants are
zero, i.e.
ckij = 0, i, j, k 6 r,
which imply that the generators of the group commute with each
other.
25:23.
94 Chapter 3
The generators of the group also satisfy the Jacobi identity, i.e.
[[Xi , Xj ], Xk ] + [[Xj , Xk ], Xi ] + [[Xk , Xi ], Xj ] = 0, (3.28)
25:23.
3.3 More on semisimple group 95
Proposition 1.
A group G is semisimple if and only if det|g| 6= 0.
det|g| = 0, (3.31)
gij̄ = clim cm l m̄ l̄ m̄
j̄l = cim̄ cj̄l = cim̄ cj̄ l̄ = 0, (3.32)
which implies that the elements of the whole column of the g tensor
equal to zero. Hence the determinant of the g tensor is zero, i.e.
det|g| = 0,
gij xj = 0. (3.33)
25:23.
96 Chapter 3
Proposition 2.
If xj are the nontrivial solutions of the following simultaneous
homogeneous equations,
gki xi = 0
and Xi is the generators of the group G, then the generators xi Xi
form an invariant subalgebra of G.
where y k = xi ckij .
which implies that y k is also the solution of Eq. (3.33), and hence
the generators xi Xi form an invariant subalgebra. While in reaching
Eq. (3.35), we have made use of the cyclic properties of the indices in
clji = cjil which will be shown later in Proposition 3.
g il glj = δji ,
C = g ij Xi Xj , (3.36)
25:23.
3.3 More on semisimple group 97
which shall commute with any generator of the group as can be verified
by taking the commutator of C with, say, the generator Xk in the
following expression
[C, Xk ] = g ij [Xi Xj , Xk ] = g ij clik (Xl Xj + Xj Xl ). (3.37)
Let us define a 3rd rank tensor through the following relation
clik = g lj cjik .
We shall prove the anti-symmetric property of cijk in the following
proposition.
Proposition 3.
The 3rd rank tensor cijk is anti-symmetric with respect to the
interchange of any pair of the indices.
= −cljm cnli cm l n m l n m l n m
kn − cmi clj ckn = cjm cil ckn + cmi clj cnk . (3.38)
By summing over the repeated indices, the last line of the above
equation becomes invariant under the cyclic permutation of i, j and k.
Therefore the conclusion of Proposition 3 is reached. Eqation (3.37) can
then be expressed as
[C, Xk ] = g ij g lm cmik (Xl Xj + Xj Xl )
25:23.
98 Chapter 3
A = ai Xi , (3.41)
X = xi Xi , (3.42)
For the r-parameter Lie group, there are r roots of the solution of
ρ in Eq. (3.44). We shall state the conclusions from Cartan’s without
providing the argument in details that if ai is chosen such that the
secular equation has the maximum numbers of distinct roots, then only
the eigenvalue ρ = 0 becomes degenerate. Let l be the multiplicity of
the degenerate roots of the secular equation, then l is said to be the
rank of the semisimple Lie algebra. The eigengenerator corresponding
to the degenerate eigenvalue ρ = 0 is denoted by Hi (i = 1, 2, . . . , l),
which commutes with each other, i.e.
[Hi , Hj ] = 0, (3.45)
25:23.
3.4 Standard form of the semisimple Lie algebras 99
A = λi Hi . (3.47)
[Hi , Eα ] = αi Eα , (3.49)
where the structure constant cβiα = αi δαβ , with α and β taking all the
distinct nonzero eigenvalues. One also proves readily that
α = λ i αi (3.50)
25:23.
100 Chapter 3
i, j 6 l,
α, β takes the values, r − l in total numbers, corresponding to
the roots of non-zero eigenvalues.
25:23.
3.5 Root vector and its properties 101
= g ij cαjα · 1 = g ij αj = αi . (3.59)
Hence we can rewrite Eq. (3.53) as follows
[Eα , E−α ] = αi Hi . (3.60)
Here we summarize the algebra of the semisimple group in the fol-
lowing standard form
[Hi , Hj ] = 0, (3.45)
[Hi , Eα ] = αi Eα , (3.49)
i
[Eα , E−α ] = α Hi , (3.60)
[Eα , Eβ ] = Nαβ Eα+β , if α + β 6= 0. (3.54)
25:23.
102 Chapter 3
Proposition 4.
If α and β are two roots, then
(α, β) 1
= integers. (3.64)
(α, α) 2
25:23.
3.5 Root vector and its properties 103
where the structure constant µj+1 can not be normalized further for
the reason that all eigengenerators corresponding to the roots in α-
string have been rescaled already. Applying the Jacobi identity for the
generators Eα , E−α and Eγ−jα , which leads to the following relation
1
µj = j(γ, α) − j(j − 1)(α, α). (3.71)
2
1
(γ, α) = g(α, α). (3.72)
2
Let us now take β = γ −jα as any root in the α-string, and eliminate
γ in Eq. (3.72), then we reach
1
(α, β) = (g − 2j)(α, α). (3.73)
2
25:23.
104 Chapter 3
1
(α, β) = m(α, α), (3.74)
2
1
(α, β) = n(β, β), (3.75)
2
(α, β)(β, α) 1
cos2 ϕ = = mn, (3.76)
(α, α)(β, β) 4
The length of the root vector α and β can then be calculated according
to the various combinations in m and n as follows:
r
p 2(α, β)
|α| = (α, α) = ,
m
r
p 2(α, β)
|β| = (β, β) = .
n
25:23.
3.6 Vector diagrams 105
A. ϕ = π/6
25:23.
106 Chapter 3
B. ϕ = π/4
They are called B2 group and C2 group, which has 2 null root vectors
and 8 root vectors and are associated with SO(5) group and Sp(4) group,
the symplectic group in 4-dimension, respectively.
C. ϕ = π/3
25:23.
3.7 PCT: discrete symmetry, discrete groups 107
D. ϕ = π/2
We shall consider another one called D2 group, with two pairs of
mutually orthogonal root vectors. D2 group is commonly referred to as
SO(4), which is isomorphic to the direct product of two SO(3) groups.
25:23.
108 Chapter 3
or in general,
P.T.
P : f (~x) 7−→ f (~xp ) = f (−~x). (3.78)
It is obvious that
P 2 = I, (3.79)
PVP −1 = Vp . (3.80)
{P, V} = 0, (3.81)
The position operator and the momentum operator are of this kind, i.e.
25:23.
3.7 PCT: discrete symmetry, discrete groups 109
[P, A] = 0, (3.84)
or
PAP −1 = Ap = A. (3.85)
One can easily prove that the angular momentum L is an axial vector
operator.
It is of interest to investigate a quantum system with symmetry
under parity transformation. We shall summarize a few remarkable
results in the following proposition.
Proposition 5.
If a dynamical operator S is invariance under the parity trans-
formation, then there exists a pair of vectors which are the si-
multaneous eigenvectors of S and P with the eigenvalues s and
±1 respectively.
S|si = s|si.
25:23.
110 Chapter 3
This implies that |sip is also the eigenvector of S with the eigenvalue
s, namely that the eigenvectors are of 2-fold degeneracy. Therefore it
allows one to construct a pair of vectors
1
|s± i = √ (|si ± |sip ), (3.87)
2
1 ~2 ~
S=H= P + V (|X|). (3.89)
2m
d
i~ hPi = h[P, H]i = 0. (3.90)
dt
25:23.
3.7 PCT: discrete symmetry, discrete groups 111
or {C, Q} = 0. (3.93)
∂ ∗ ~
i~ ψ (~x, −t) = H − ∇, ~x ψ ∗ (~x, −t). (3.95)
∂t i
25:23.
112 Chapter 3
tion becomes
∂ ∗ ~
i~ ψ (~x, t) = H ∇, ~x ψ ∗ (~x, −t). (3.96)
∂t i
This implies that both ψ(~x, t) and ψ ∗ (~x, −t) are solutions of the
Schrödinger equation.
The appearance of the complex conjugate as well as the replacement
of t by −t in the wave function of the last equation allows one to con-
struct the time reversal operator which bears some sort of property in
antilinearity. If we decompose T into the product of K and T , respec-
tively standing for an antilinear operator and a temporal reflection
operator, then we reach the conclusions in the following proposition.
Proposition 6.
If the time reversal operator T takes as product of K and T , i.e.
T = KT ,
(
K = antilinear operator,
where
T = temporal reflection operator,
then both ψ(~x, t) and Kψ(~x, −t) are solutions of the time depen-
dent Schrödinger equation as long as KHK −1 = H. Moreover,
K 2 ψ = ηψ for any wave function ψ and unimodulus factor η.
−1 ~
T HT = KH ∇, ~x, −t K −1 ,
i
25:23.
3.7 PCT: discrete symmetry, discrete groups 113
∂
i~ Kψ(~x, −t) = (KHK −1 )Kψ(~x, −t). (3.97)
∂t
∂ ~
i~ Kψ(~x, −t) = H ∇, ~x Kψ(~x, −t). (3.98)
∂t i
Example
25:23.
114 Chapter 3
3.8 Exercises
Ex 3.8.1
x0 = a1 x + a2 .
(b) Denote the group element by a = (a1 , a2 ). Find the inverse ele-
ment of a, namely, a−1 = ((a−1 )1 , (a−1 )2 ).
(c) What are the composition rules of the group parameters, i.e. try
to find c = ba for cl = cl (ai , bj ) = f l (ai , bj )?
Ex 3.8.2
x03 = x3 ,
25:23.
3.8 Exercises 115
2 ∂ 1 ∂
exp(θX3 )F (x) = exp θ −x +x F (x1 , x2 , x3 )
∂x1 ∂x2
Ex 3.8.3
A conjugate subgroup is defined as a−1 Ha if H is a subgroup, where
a ∈ G. Show that the self-conjugate subgroup can alternatively be
defined as the invariant subgroup.
Ex 3.8.4
Verify that the Jacobi identity leads to following relation among the
structure i.e.
clij cm l m l m
lk + cjk cli + cki clj = 0.
Ex 3.8.5
Show that the elements giα of the g tensor in the standard form of
Lie algebra vanish.
Ex 3.8.6
Give the argument to verify the elements gαβ of the g tensor in the
standard form of Lie algebra also vanish if α + β 6= 0.
Ex 3.8.7
Let the parity operator be defined as (Hint: by using Eq. (1.39).)
25:23.
116 Chapter 3
π ~ ~ ~ ~
P = e 2 (P ·X+X·P ) .
Show that
~ −1 = −X,
P XP ~
P P~ P −1 = −P~ .
25:23.
Chapter 4
Angular Momentum
25:31.
118 Chapter 4
that allows us to calculate the matrix elements uli in the rotational group
O(3), i.e.
∂xl
uli = = lik xk .
∂θi
∂
Since Xi = uli , we obtain that
∂xl
∂
Xi = lik xk , (4.4)
∂xl
25:31.
4.1 O(3) group, SU(2) group and angular momentum 119
m ∂ k ∂ ∂
[Xi , Xj ] = lim njk x ,x = (kim ljk + kmj lik )xm l
∂xl ∂xn ∂x
∂
= kij lkm xm = −kij Xk , (4.5)
∂xl
where the fourth term of the above equation is obtained by the Jacobi
identity,
kim ljk + kmj lik + kji lmk = 0. (4.6)
Hence we have the following Lie algebra of the rotational transfor-
mation in 3-dimensional vector space,
[Xi , Xj ] = −kij Xk , (4.7)
that reproduce the commutation relations for the quantum operators of
the angular momentum, i.e.
[Li , Lj ] = i~kij Lk , (4.8)
i
if we put Xi = Li .
~
There are various representations for the generators of the symme-
try transformation as well as for the Lie algebras. The representations
of the group generators and the Lie algebras we adopted so far are ex-
pressed with the purpose to investigate the properties in the symmetry
transformation of the functions. We refer this representation as the
canonical formulation. If we focus our attention on the transformation
of the vectors directly, we attain completely different representations.
Take the rotational transformation of a vector in matrix notation as
follows
x01
x1
x0 = Rx
02 2
or x = R x , (4.9)
x03 x3
25:31.
120 Chapter 4
x0T x0 = xT RT Rx or RT R = I. (4.10)
R(θ1 , θ2 , θ3 ) = I + A(θ1 , θ2 , θ3 ),
A(0, 0, 0) = 0. (4.12)
RT R = (I + AT )(I + A) ' I + AT + A = I,
25:31.
4.1 O(3) group, SU(2) group and angular momentum 121
∂
Xi = A(θ1 , θ2 , θ3 ) , (4.13)
∂θi θi =0
0 − sin ϕ 0 0 −ϕ 0
R = I + A(0, 0, ϕ) = I + sin ϕ 0 0 ' I + ϕ 0 0 ,
0 0 0 0 0 0
0 −1 0
∂A
X3 = = 1 0 0 . (4.14)
∂ϕ
0 0 0
cos ϕ − sin ϕ 0
R(0, 0, ϕ) = eϕX3 = sin ϕ cos ϕ 0 . (4.15)
0 0 1
25:31.
122 Chapter 4
0 0 1 0 0 0
X2 = 0 0 0 , X1 = 0 0 −1 , (4.16)
−1 0 0 0 1 0
which takes only the positive root in the equation (det R)2 − 1 = 0.
This can be understood that the matrix R(ϕ, θ, χ) is reached from the
identity element I = R(0, 0, 0) with the determinant being equal to one,
by varying the group parameters continuously away from the origin of
25:31.
4.1 O(3) group, SU(2) group and angular momentum 123
25:31.
124 Chapter 4
!
ξ1
ξ= . (4.22)
ξ2
! ! !
0 ξ 01 α β ξ1
ξ = Mξ, or = , (4.23)
ξ 02 γ δ ξ2
ξ 0† ξ 0 = ξ † M† Mξ = ξ † ξ, (4.24)
25:31.
4.1 O(3) group, SU(2) group and angular momentum 125
! !
α∗ β ∗ 1 δ −β
= . (4.26)
γ ∗ δ∗ det M −γ α
α∗ = δ, β = −γ ∗ , (4.27)
!
α β
M= ,
−β α∗
∗
!
ξ β
M=I+ = I + D.
−β ξ ∗
∗
M† M = (I + D† )(I + D) ' I + D† + D = I,
!
i i
2c 2 (a− ib)
M=I+ i
,
2 (a + ib) − 2i c
25:31.
126 Chapter 4
where a, b and c, all real, are the three parameters of the SU(2) group.
The generators of the group are obtained as follows
! !
i 1
∂M 0 2 ∂M 0 2
Y1 = = i
, Y2 = = ,
∂a 2 0 ∂b − 12 0
!
i
∂M 2 0
Y3 = = . (4.28)
∂c 0 − 2i
i i i
Y1 = σ 1 , Y2 = σ2 , Y3 = σ3 . (4.30)
2 2 2
The identical Lie algebras for both groups O(3) and SU(2) do not
necessary have the same domain in the group manifold. Let us construct
the group elements, i.e. the matrix M by exponentiating the group
generators as follows
~
M = eχn̂·Y , (4.31)
where n̂ = n̂(θ, ϕ) has the same definition as that of O(3) group, yet the
parameter χ takes different domain in the group manifold for the reason
that both matrix M and matrix −M fulfill the unitarity condition of
the transformation. Therefore the group elements of SU(2) will fill the
sphere of radius 2π, namely it takes all the points in a domain of the
sphere with radius 2π in the parameter space to exhaust the group
elements. The parameters ϕ, θ and χ range as follows
25:31.
4.1 O(3) group, SU(2) group and angular momentum 127
with again the points of antipode on the surface of the sphere are iden-
tified.
It will be left to you as an exercise to show that
~ ) = cos χ χ
M(χn̂ · Y + in̂ · ~σ sin , (4.33)
2 2
~ ) = (−1)m M(χn̂ · Y
M((χ + 2mπ)n̂ · Y ~ ). (4.34)
!
i x3 x1 − ix2
X = x σi = . (4.35)
x1 + ix2 −x3
The above transformation leaves the norm of the vector invariant and is
justified by taking the determinant on both sides of the above equation,
i.e.
−x0i x0i = det X0 = (det M† )(det X)(det M) = −xi xi |det M|2 = −xi xi .
(4.37)
25:31.
128 Chapter 4
Let us take the rotation about the 3rd axis with angle ϕ as an
example and evaluate the matrix M = eϕY3 explicitly as follows
!
x03 x01 − ix02
X0 =
x01 + ix02 −x03
i i
! ! !
e− 2 ϕ 0 x3 x1 − ix2 e2ϕ 0
= i i
0 e2ϕ x1 + ix2 −x3 0 e− 2 ϕ
!
x3 e−iϕ (x1 − ix2 )
= . (4.38)
eiϕ (x1 + ix2 ) −x3
x03 = x3 , (4.41)
Let us consider the Lie algebras of the O(3) group or the SU(2)
group, in which we take the generators as follows
~ ~
Ji = Xi , or Ji = Yi , (4.42)
i i
25:31.
4.2 O(3)/SU(2) algebras and angular momentum 129
A = ai Ji , X = xi Ji , (4.45)
−ρ −i~a3 i~a2
i~a3 −ρ −i~a1 = 0. (4.49)
−i~a2 i~a1 −ρ
µ0 = 0, (4.50a)
p
µ+ = ~ (a1 )2 + (a2 )2 + (a3 )2 , (4.50b)
p
µ− = −~ (a1 )2 + (a2 )2 + (a3 )2 . (4.50c)
25:31.
130 Chapter 4
ai Ji A
J0 = p =p , (4.51)
1 2 2 2 3
(a ) + (a ) + (a )2 (a ) + (a2 )2 + (a3 )2
1 2
[[J+ , J− ], J0 ] = 0, (4.52)
where we take α to be one due to the fact that the eigenvalue equation is
linear in X on both sides of the equation, that allows the eigengenerators
J+ and J− to absorb any arbitrary factor. Henceforth the Lie algebra
of the O(3)/SU(2) group is summarized as follows
J0 = J0† , †
J+ = J− , †
J− = J+ . (4.56)
C = 2g ij Ji Jj . (4.57)
The factor 2 is introduced with the purpose to identify C with the total
angular momentum operator J 2 , i.e.
25:31.
4.2 O(3)/SU(2) algebras and angular momentum 131
C = J 2 = J+ J− + J− J+ + J02 . (4.58)
Since C commutes with the generators of the group, it is then invariant
under the transformation. Furthermore the eigenvalues of the Casimir
operator will be used to label the dimension of the irreducible represen-
tation of the group.
Let us construct the eigenvectors of both the operators C and J0 ,
and denote it as |c, νi, referred to commonly as the irreducible repre-
sentations of the O(3)/SU(2) group, hence we have
C|c, νi = c|c, νi, (4.59a)
J0 |c, νi = ν~|c, νi. (4.59b)
With the commutation relations,[J0 , J± ] = ±~J± , one is now ready
to show the following proposition.
Proposition 1.
If |c, νi is the eigenvector of the Casimir operator C and the op-
erator J0 with the eigenvalues c and ν~ respectively, then J± |c, νi
are also the eigenvectors of operators C and J0 with the eigen-
values c and (ν ± 1)~ respectively.
25:31.
132 Chapter 4
|c, ν− i, . . . , |c, ν−qi, . . . , |c, ν−1i, |c, νi, |c, ν+1i, . . . , |c, ν+pi, . . . , |c, ν+ i.
1
hc, ν+ |J− J+ |c, ν+ i = hc, ν+ |(C − J02 − ~J0 )|c, ν+ i = 0,
2
or
c − ~2 ν+ (ν+ + 1) = 0, (4.61)
c − ~2 ν− (ν− − 1) = 0. (4.62)
r
1 c
ν+ − ν− = 2 + 2 − 1 = 2j, (4.63)
4 ~
ν+ = −ν− = j. (4.65)
25:31.
4.2 O(3)/SU(2) algebras and angular momentum 133
~ p
J+ |j, mi = √ (j − m)(j + m + 1)|j, m + 1i, (4.66a)
2
~ p
J− |j, mi = √ (j + m)(j − m + 1)|j, m − 1i. (4.66b)
2
25:31.
134 Chapter 4
! ! !
(1) ~ 0 1 (1) ~ 0 0 (1) ~ 1 0
J+2 =√ , J−2 =√ , J0 2 = ,
2 0 0 2 1 0 2 0 −1
(4.67)
0 1 0
~
Jx(1) = √ 1 0 1 , (4.69a)
2
0 1 0
0 −i 0
~
Jy(1) = √ i 0 −i , (4.69b)
2
0 i 0
1 0 0
Jz(1) = ~ 0 0 0 , (4.69c)
0 0 −1
25:31.
4.3 Irreducible representations of O(3) group and spherical harmonics 135
Proposition 2.
Let Ylm and Ylm−1 be two spherical harmonics of the same degree
l with the order m and m−1 respectively, then they are connected
by the following relation
Ylm (θ,ϕ)
√
2 ∂ ∂
=p eiϕ
+ i cot θ Ylm−1 (θ, ϕ).
(l + m)(l − m − 1) ∂θ ∂ϕ
(4.71)
25:31.
136 Chapter 4
lowing equation
√
2
hθ, ϕ|l, mi = p hθ, ϕ|L+ |l, m − 1i,
(l + m)(l − m − 1)~
Proposition 3.
All the spherical harmonics takes the form of the product of eimϕ ,
an exponential function of ϕ and another function of f (θ), i.e.
Ylm (θ, ϕ) = eimϕ f (θ). (4.73)
∂ m
Y (θ, ϕ) = imYlm (θ, ϕ),
∂ϕ l
25:31.
4.3 Irreducible representations of O(3) group and spherical harmonics 137
Proposition 4.
Function f (θ) satisfies the associated Legendre’s differential
equation, i.e.
m2
d 2 d
(1 − µ ) f (µ) + l(l + 1) − f (µ) = 0, (4.74)
dµ dµ 1 − µ2
or more explicitly by writing f (θ) in terms of the associated
Legendre polynomial of degree l and order m, i.e.
f (θ) ∝ Plm (cos θ) = Plm (µ), (4.75)
where we put µ = cos θ, and identify f (θ) = f (µ).
1 ∂2
1 ∂ ∂
L2 = −~2 sin θ +
sin θ ∂θ ∂θ sin2 θ ∂ 2 ϕ
1 ∂2
1 ∂ ∂
− sin θ + Ylm (θ, ϕ) = l(l + l)Ylm (θ, ϕ).
sin θ ∂θ ∂θ sin2 θ ∂ 2 ϕ
m2
d 2 d
(1 − µ ) f (µ) + l(l + 1) − f (µ) = 0,
dµ dµ 1 − µ2
25:31.
138 Chapter 4
Hence A reads as
r
(2l + 1)! 1
A= .
4π 2l l!
25:31.
4.3 Irreducible representations of O(3) group and spherical harmonics 139
r
(2l + 1)! 1 −ilϕ l
Yl−l (θ, ϕ) = e sin θ.
4π 2l l!
s l+m
(l − m)! ∂ ∂
Ylm (θ, ϕ) = eiϕ
+ i cot θ Yl−l (θ, ϕ)
(2l)!(l + m)! ∂θ ∂ϕ
s l+m
1 (2l + 1)(l − m)! ∂ ∂
= l iϕ
e + i cot θ e−ilϕ sinl θ.
2 l! 4π(l + m)! ∂θ ∂ϕ
Proposition 5.
∂ ∂
Acting the operator eiϕ ∂θ + i cot θ ∂ϕ n times repeatedly upon
the function eipϕ f (θ) will result in another function of the fol-
lowing form, i.e.
n
∂ ∂
eiϕ + i cot θ eipϕ f (θ)
∂θ ∂ϕ
n i(p+n)ϕ p+n d −p
= (−1) e sin θ sin θ f (θ).
d cos θ
(4.76)
25:31.
140 Chapter 4
iϕ ∂ ∂
e + i cot θ eipϕ f (θ)
∂θ ∂ϕ
i(p+1)ϕ d −p
= (−1)e sinp+1 θ sin θ f (θ). (4.77)
d cos θ
d −p
If we regard sinp+1 θ sin θ f (θ) as a new function of g(θ),
d cos θ
i.e.
d −p
g(θ) = sinp+1 θ sin θ f (θ),
d cos θ
iϕ ∂ ∂
e + i cot θ ei(p+1)ϕ g(θ)
∂θ ∂ϕ
i(p+2)ϕ p+2 d −p−1
= (−1)e sin θ sin θ g(θ)
d cos θ
d2
2 i(p+2)ϕ p+2 −p
= (−1) e sin θ 2 sin θ f (θ) .
d cos θ
n
iϕ ∂ ∂
e + i cot θ eipϕ f (θ)
∂θ ∂ϕ
n i(p+n)ϕ p+n d −p
= (−1) e sin θ sin θ f (θ),
d cos θ
25:31.
4.4 O(4) group, dynamical symmetry and the hydrogen atom 141
s l+m
(2l + 1)(l − m)! 1 ∂ ∂
Ylm (θ, ϕ) = eiϕ + i cot θ e−ilϕ sinl θ
4π(l + m)! 2l l! ∂θ ∂ϕ
s
(2l + 1)(l − m)! eimϕ m dl+m
= (−1) l+m
sin θ sin2l θ
4π(l + m)! 2l l! dl+m cos θ
s
(2l + 1)(l − m)! eimϕ 2 m dl+m
= (−1)l+m (1 − µ ) 2 (1 − µ2 )l
4π(l + m)! 2l l! dl+m µ
s
m (2l + 1)(l − m)! eimϕ m
= (−1) P (µ),
4π(l + m)! 2l l! l
where
(−1)l 2 m dl+m 2 m dm
Plm (µ) = (1 − µ ) 2 (1 − µ 2 l
) = (−1)l
(1 − µ ) 2 Pl (µ),
2l l! dl+m µ dm µ
1 dl
Pl (µ) = (1 − µ2 )l ,
2l l! dl µ
25:31.
142 Chapter 4
∂ ∂
xi j
− xj i , for i, j = 1, 2, 3, 4.
∂x ∂x
If we denote
∂ ∂ ∂ ∂ ∂ ∂
M1 = x 2 − x3 2 , M2 = x3 − x1 3 , M3 = x1 − x2 1 ,
∂x3 ∂x ∂x1 ∂x ∂x2 ∂x
and
∂ ∂ ∂ ∂ ∂ ∂
N1 = x1 4
− x4 1 , N 2 = x2 4
− x4 2 , N 3 = x3 4
− x4 3 ,
∂x ∂x ∂x ∂x ∂x ∂x
1 1
Ai = (Mi + Ni ), Bi = (Mi − Ni ), (4.79)
2 2
Among the Lie algebra of the O(4) group, we find that there exist
two sets of generators Ai and Bi whose commutators respectively form
25:31.
4.4 O(4) group, dynamical symmetry and the hydrogen atom 143
a sub-algebra of the O(3) group. The O(4) algebra becomes the direct
sum of two O(3) algebras and the O(4) group is then locally isomorphic
onto O(3) ⊗ O(3) group.
We are now in position to discuss the hydrogen-like atoms from
the dynamical symmetry properties. The Hamiltonian operator for the
hydrogen-like atoms takes the following expression
1 2 Ze2 1 2 k
H= P − = P − , (4.81)
2m r 2m r
where we have fixed the center of the Coulomb potential at the origin.
The space described by this coordinate system is isotropic with respect
to the rotation about any axis through the origin. It implies that the
angular momentum operators are conserved because they commute with
the Hamiltonian operator. Furthermore, the attractive Coulomb poten-
tial enables us to introduce another three conserved operators, called
Lenz operators. Classically, it is the vector, called the Lenz vector or
the Runge-Lenz vector, that originates from the center of the force and
points to the aphelion, as shown in Figure 4.2. It is a particular feature
that the orbit is closed and fixed in space for the case of the Coulomb
potential. If we construct a plane perpendicular to the constant angular
momentum through the origin, the closed orbit lies on the plane without
25:31.
144 Chapter 4
precession of the major axis. Then the Lenz vector can be constructed
by means of the following elliptical orbit equation, i.e.
1 mk
= 2 (1 − cos θ), (4.82)
r l
l2 l2
~ = rR cos θ = r 1
~r · R 1− = αr2 − βl2 ,
mk(1 − ) mkr
or
l2 l2
αr − r − l2 β − 2 2
= 0,
mk(1 − ) m k (1 − )
1 l2 l2
α= , β= ,
r mk(1 − ) m2 k 2 (1 − )
l2
~ = ~r 1 ~
R + l × p~ . (4.83)
mk(1 − ) r mk
25:31.
4.4 O(4) group, dynamical symmetry and the hydrogen atom 145
2H 1
[Ri , Rj ] = − i~ijk Lk .
mk (1 − )2
2 2
r
mk 2
R̃i = (1 − )Ri ,
−2H
1 1
Ji = (Li + R̃i ), Ki = (Li − R̃i ), (4.86)
2 2
and these operators are the generators of O(3) ⊗ O(3) group, namely,
25:31.
146 Chapter 4
mk 2
2 1 2 2
J +K =− +~ , (4.88)
2 2H
or
mk 2
1
H=− . (4.89)
2 2(J 2 + K 2 ) + ~2
mk 2 1 mk 2 1
E=− 2 2
= − , (4.92)
2 4j(j + 1)~ + ~ 2 (2j + 1)2 ~2
2
mk 2 1 mc2 e2 mc2 2 1
1
En = − 2 2 = − = − α 2,
2~ n 2 ~c n2 2 n
25:31.
4.5 Exercises 147
4.5 Exercises
Ex 4.5.1
Ex 4.5.2
Taking σ i = σi and making use of the following properties of Pauli
matrices,
(a) hermiticity σ i† = σ i ,
(b) traceless Tr σ i = 0,
(c) {σ i , σ j } = 2δ ij I,
(d) Tr σ i σ j = 2δ ij ,
1
Rji = Tr (σ i M † σj M ).
2
Ex 4.5.3
Construct explicitly the matrix representations of the generators in
two dimensions.
25:31.
148 Chapter 4
Ex 4.5.4
Denote the 2 × 2 traceless matrices A and B by arbitrary parameters
as A = ~a · ~σ , B = ~b · ~σ . Show that AB = ~a · ~bI + i(~a × ~b) · ~σ .
Ex 4.5.5
Show that
ϕ)~σ U† (~
U(~ ϕ) = (n̂ · ~σ )n̂ + n̂ × ~σ sin ϕ − n̂ × (n̂ × ~σ ) cos ϕ,
ϕ) = exp(i ϕ2 n̂ · ~σ ).
where U(~
Ex 4.5.6
Show that the Casimir operator of the rotational group takes the
expression
(
Ji Ji in the Cartesian basis,
C= 2
J+ J− + J− J+ + J0 in the standard form.
Ex 4.5.7
Find the unitary matrix U which diagonalizes
0 −1 0 1 0 0
~ (1)
Jz(c) = 1 0 0 into J0 = ~ 0 0 0 ,
i
0 0 0 0 0 −1
0 1 0 0 −i 0
~ ~
Jx(1) = √ 1 0 1 , Jy(1) = √ i 0 −i .
2 2
0 1 0 0 i 0
25:31.
4.5 Exercises 149
Ex 4.5.8
Show that [H, Ri ] = O for the hydrogen atom.
Ex 4.5.9
Prove that
2H 1
[Ri .Rj ] = − i~ijk Lk .
mk (1 − )2
2 2
Ex 4.5.10
Prove that
mk 2
2 1 2 2
J +K =− +~ .
2 2H
25:31.
This page intentionally left blank
25:31.
Chapter 5
Space and time had been treated separately since the birth of physics
as an experimental science in mediaeval time. As Newton stated in his
work, The Principia, absolute space, in its own nature, without relation
to anything external, remains always similar and immovable. Absolute,
true, and mathematical time, of itself, and from its own nature, flows
equably without relation to anything external. Physicists had been
confined by the concept of the absoluteness of space and time such
that the coordinate transformations from one inertial frame to another,
always kept time as an invariant parameter.
All this concept of absolute space and time resulted in the formu-
lation of Galilean relativity which has not been challenged until the
end of 19th century and the beginning of the 20th century. It was
the triumphal results that related the wave phenomena of light to the
electromagnetic theory in Maxwell equations, that space and time were
interwoven into the formulation of the theory and were treated as equals.
The development in special relativity afterwards led naturally to gen-
eralize the 3-dimensional coordinate space to the 4-dimensional one by
including the time-axis, an extra dimension. And so it is no longer the
Euclidean vector space, but a four-dimensional Minkowski space with
metric tensor defined as
35:17.
152 Chapter 5
35:17.
5.1 Space-time structure and Minkowski space 153
namely that
ΛT gΛ = g or ΛT g = gΛ−1 . (5.7)
This is the crucial difference between the Euclidean space and the
Minkowski space because the g matrix is involved in the latter case.
If det Λ 6= 0, then the transformation forms a group, called the ho-
mogeneous Lorentz group. If we take the determinant on both sides
of Eq. (5.7), we obtain (det Λ)2 = 1. We shall restrict ourselves to the
case det Λ = 1 with which the group elements can be reached continu-
ously from the identity element. The length invariance will provide 10
equations of constraints among the matrix elements of Λ, namely
35:17.
154 Chapter 5
p
where β = v/c and γ = 1/ 1 − β 2 .
It is interesting to visualize the above boost transformation as a
rotation on the (x0 , x1 ) plane with an imaginary angle, or hyperbolic
angle ξ defined by
ξ = tanh−1 β. (5.11)
cosh ξ − sinh ξ 0 0
− sinh ξ cosh ξ 0 0
Λ=
0
. (5.12)
0 1 0
0 0 0 1
0 −1 0 0
−1 0 0 0
B1 =
. (5.13a)
0 0 0 0
0 0 0 0
0 0 −1 0 0 0 0 −1
0 0 0 0 0 0 0 0
B2 =
−1
, B3 = . (5.13b,c)
0 0 0
0
0 0 0
0 0 0 0 −1 0 0 0
35:17.
5.1 Space-time structure and Minkowski space 155
The remaining three generators of the SO(3,1) group are merely those
of the spatial rotational ones, expressed as:
0 0 0 0
0 0 0 0
A1 =
0
(5.14a)
0 0 1
0 0 −1 0
0 0 0 0 0 0 0 0
0 0 0 −1 0 0 1 0
and A2 =
0
, A3 =
0 −1 0
. (5.14b,c)
0 0 0 0
0 1 0 0 0 0 0 0
but the commutators among the boost generators Bi take the different
sign to close the algebra, i.e.
cosh ξ − sinh ξ 0 0
− sinh ξ cosh ξ 0 0
eξB1 =
0
. (5.17)
0 1 0
0 0 0 1
35:17.
156 Chapter 5
∂ ∂
M µν = xµ − xν , (5.19)
∂xν ∂xµ
[M µν , M αβ ] = −g νβ M µα − g µα M νβ + g να M µβ + g µβ M να . (5.20)
1
Bi = gij M 0j , Ai = ijk M jk . (5.21)
2
As in the case of the SO(4) group, let us take the linear combinations
of the generators of the SO(3,1) group and denote that
1 Ai 1 Ai
Li = + Bi , and Ri = − Bi . (5.22a,b)
2 i 2 i
Then the Lie algebra of the SO(3,1) group takes the expressions as those
of the SO(3) ⊗ SO(3) group, namely, i.e.
35:17.
5.2 Irreducible representation of SO(3,1) and Lorentz spinors 157
[Li , Ri ] = 0, (5.23b)
Yet they are not exactly parallel to the case of the SO(4) group,
because the generators Li and Ri of the non-compact Lorentz group are
not Hermitian. The immediate consequence of these properties gives
rise to two options we have to face. Either we choose the finite dimen-
sional representations, in which case the unitarity condition shall be
abandoned or we choose to preserve the unitarity of the representation
by accepting the infinite dimensional representations.
Let us consider the finite dimensional irreducible representation, and
take the direct product |l, mil ⊗ |r, nir ≡ |l, m; r, ni as the bases, where
|l, mil is the eigenvector of the commuting operators L2 and L3 with the
eigenvalues l(l + 1) and m respectively, while |r, nir is the eigenvector
of another pair of commuting operators R2 and R3 with the eigenvalues
r(r + 1) and n respectively, such that
35:17.
158 Chapter 5
( 1 ,0) 1 ( 1 ,0)
Li 2 = σi , Ri 2 = 0, (5.26a)
2
(0, 12 ) (0, 12 ) 1
Li = 0, Ri = σi , (5.26b)
2
( 1 ,0) i ( 1 ,0) 1
Ai 2 = σi , Bi 2 = σi , (5.27a)
2 2
(0, 12 ) i (0, 12 ) 1
Ai = σi , Bi = − σi , (5.27b)
2 2
or
(0, 21 )† ~ ~ i ~ ~ 1
~ −1 ,
~ ξ)
D (θ, ξ) = exp − ~σ · (θ − iξ) 6= D(0, 2 ) (θ,
2
35:17.
5.2 Irreducible representation of SO(3,1) and Lorentz spinors 159
! !
1 1 1 1 1 0
| , il 7−→ = e1 , | , − il 7−→ = e2 , (5.29a,b)
2 2 0 2 2 1
space, and construct the left-handed Lorentz spinor, or simply the left-
handed spinor as follows
!
ψl1 (x)
ψl (x) = ψla (x)ea = , (5.30)
ψl2 (x)
where ψl1 (x) or ψl2 (x) are all complex functions of space-time coordinate
x = (x0 , x1 , x2 , x3 ). Under the Lorentz transformation, the left-handed
spinor is transformed by applying the (1/2, 0) matrix representation of
the SO(3,1) group, i.e. Eq. (5.28a) upon the spinor as follows
1
ψl (x) 7−→ ψl0 (x0 ) = D( 2 ,0) (θ, ~ l (Λ−1 x0 ),
~ ξ)ψ (5.31)
! ! !
1 0 ψr1 (x)
ψr (x) = ψra (x)fa = ψr1 (x) + ψr2 (x) = , (5.32)
0 1 ψr2 (x)
35:17.
160 Chapter 5
where ψra (x)(a = 1, 2) are the complex functions of the space-time co-
ordinates, and fa (a = 1, 2) are the bases column matrix of the right-
handed spinor space. Similar to the transformation in the left-handed
spinor space, the Lorentz transformation for ψr (x) can then be written
as
1
ψr (x) 7−→ ψr0 (x0 ) = D(0, 2 ) (θ, ~ r (Λ−1 x0 )).
~ ξ)ψ (5.33)
1
~ ξ)
It is interesting to observe that both matrices D( 2 ,0) (θ, ~ and
1
~ ξ)
D(0, 2 ) (θ, ~ play more roles than just performing the Lorentz trans-
formation upon the spinor wave functions. Let us consider the linear
transformation on a two-dimensional complex vector space, in which a
vector ξ is acted upon by a 2×2 matrix L with complex matrix elements
as follows
! ! ! !
0 ξ 01 ξ1 a b ξ1
ξ = =L 2 = . (5.34)
ξ 02 ξ c d ξ2
a b
det L = = ab − bc = 1, (5.35)
c d
L = eA , (5.36)
35:17.
5.3 SL(2,C) group and the Lorentz transformation 161
Proposition 1.
If a matrix L can be expressed as L = eA , then
det L = eTr A
. (5.37)
and similarly
~ ξ)~ ~ ξ)~
= eTr
1
det D(0, 2 ) (θ, ~ = det e 2i ~σ·(θ+i
~ ξ) i
2
σ ·(θ+i
~
= 1. (5.42)
35:17.
162 Chapter 5
i 1 i 1
Ai = σi , Bi = σi or Ai = σi , Bi = − σi , (5.43a,b)
2 2 2 2
!
µ −x0 + x3 x1 − ix2
X = x σµ = , (5.45)
x1 + ix2 −x0 − x3
where σ 0 = −σ0 = I, a unit matrix, and σi are the Pauli matrices. The
length of the space-time vector is related the determinant of the matrix,
i.e.
35:17.
5.4 Chiral transformation and spinor algebra 163
1 1
! !
0 1
σ ξ 1
σ ξ e2ξ 0 e2ξ 0
X =e 2 3 Xe 2 3 = − 12 ξ
X − 12 ξ
. (5.49)
0 e 0 e
x01 = x1 , (5.50b)
x02 = x2 , (5.50c)
35:17.
164 Chapter 5
Yet the new set of the group generators Li and Ri are transformed
according to the following relations
−1 1 Ai
KLi K = K + Bi K−1 = −Ri , KRi K−1 = −Li , (5.54a,b)
2 i
Proposition 2.
The vector KLjm is the eigenvector of R2 and R3 with the eigen-
values j(j + 1) and −m respectively. While the vector KRkn is
the eigenvector of L2 and L3 with the eigenvalues k(k + 1) and
−n respectively.
35:17.
5.4 Chiral transformation and spinor algebra 165
The proof goes as follows: let us use the brief notations Ljm , Rkn
instead of |j, mil and |k, nir respectively, and consider R2 KLjm =
KL2 Ljm = j(j + 1)KLjm , and R3 KLjm = −KL3 Ljm = −mKLjm
which implies that KLjm = γ(m)Rj,−m . Similarly that KRkn = δ(n)×
Lk,−n .
The antiunitarity of the operator K allows us to find the coefficients
γ(m) and δ(n) as follows
(KLjm , KLjm0 ) = γ ∗ (m)γ(m0 )(Rj,−m , Rj,−m0 ) = (Ljm0 , Ljm ), (5.55)
2
( 1 ,0)
representation as an example by considering the matrix elementDmm 0
( 1 ,0) θ·A+ξ·B ~ ~ ~ ~ ~ ~ ~ ~
2
Dmm 0 = (L 1 ,m , e L 1 ,m0 ) = (Keθ·A+ξ·B L 1 ,m0 , KL 1 ,m )
2 2 2 2
~ ~ ~ ~ (0, 1 )∗
= γ ∗ (m0 )γ(m)(eθ·A+ξ·B R 1 ,−m0 , R 1 ,−m ) = γ(m)D−m,−m
2 ∗ 0
0 γ (m ),
2 2
! !
( 12 ,0) 0 γ( 12 ) 1 0 γ ∗ (− 21 )
D = D(0, 2 )∗ . (5.56)
1
γ(− 2 ) 0 γ ∗ ( 12 ) 0
! !
(0, 12 ) 0 δ( 12 ) 1 0 δ ∗ (− 21 )
D = D( 2 ,0)∗ . (5.57)
1
δ(− 2 ) 0 δ ∗ ( 12 ) 0
35:17.
166 Chapter 5
1 1
D( 2 ,0) = D0,( 2 )∗ −1 , (5.59b)
or in matrix notation as
ψ̇ = ψ T T = (ψ̇1 , ψ̇1 ) = (ψ 1 , ψ 2 )T = (ψ 1 , ψ 2 )−1 . (5.63)
35:17.
5.4 Chiral transformation and spinor algebra 167
1 1
= ψ̇(D( 2 ,0)∗ −1 )† = ψ̇D(0, 2 )† .
ϕ̇ = ϕT T . (5.64)
!
( 12 , 1̇2 ) u11̇ u12̇
U = . (5.66)
u21̇ u22̇
!
( 1̇2 , 12 ) u1̇1 u2̇1
U = . (5.67)
u1̇2 u2̇2
35:17.
168 Chapter 5
1 1̇ L.T. 0 ( 1 , 1̇ ) 1 1
U ( 2 , 2 ) 7−→ U 2 2 = D( 2 ,0) U D( 2 ,0)† , (5.68a)
1̇ 1 L.T. 0 ( 1̇ , 1 ) 1 1
U ( 2 , 2 ) 7−→ U 2 2 = D(0, 2 ) U D(0, 2 )† . (5.68b)
1
(dτ )2 = − gµν dxµ dxν . (5.70)
c2
35:17.
5.5 Lorentz spinors and the Dirac equation 169
Yet this invariant quantity when measured in the initial reference frame,
namely O-system, is given by
r
v2
1 µ ν 1
dτ = − 2 gµν dx dx 2 = 1 − 2 dt. (5.72)
c c
pµ = m0 uµ . (5.73)
dx0 m0 c
p0 = m 0 =p , (5.74a)
dτ 1 − β2
d~x m0 cβ~
p~ = m0 =p . (5.74b)
dτ 1 − β2
One can easily verify that the scalar product, or the length of the 4-
vector pµ is characterized by the rest mass of the moving particle, an
invariant under the Lorentz transformation, namely
pµ pµ = −m20 c2 . (5.75)
!
−p0 + p3 p1 − ip2
P = p µ σµ = . (5.76)
p1 + ip2 −p0 − p3
35:17.
170 Chapter 5
1
L.T.
P 7−→ P 0 = D( 2 ,0) (θ, ~ D( 12 ,0)† (θ,
~ ξ)P ~ ξ).
~ (5.77)
or in matrix notation,
1 1
! ! !
0 e2ξ 0 −m0 c 0 e2ξ 0
P = − 21 ξ 1
0 e 0 −m0 c 0 e− 2 ξ
!
−m0 ceξ 0
= . (5.79)
0 −m0 ce−ξ
We expect that the energy and the momentum of the mass point, when
measured in the O0 -system, will be respectively given as follows:
m0 c
p00 = m0 c cosh ξ = p , (5.80a)
1 − β2
−m0 v
p03 = −m0 c sinh ξ = p . (5.80b)
1 − β2
One recognizes that X and Xc belong respectively to the ( 21 , 1̇2 ) and the
35:17.
5.5 Lorentz spinors and the Dirac equation 171
Proposition 3.
35:17.
172 Chapter 5
L.T. 1 1 1 1
ξ 7−→ ξ 0 = A0c η 0 = D(0, 2 ) Ac D(0, 2 )† D( 2 ,0) ξ = D(0, 2 ) ξ, (5.85)
Pc ψl = m0 cψr , (5.86a)
P ψr = m0 cψl , (5.86b)
! ! !
0 Pc ψr ψr
= m0 c , (5.87)
P 0 ψl ψl
or
!
0 Pc
ψd (x) = m0 cψd (x), (5.88)
P 0
~ ∂ ~
Pµ = µ
= ∂µ . (5.89)
i ∂x i
35:17.
5.5 Lorentz spinors and the Dirac equation 173
" ! #
0 iσcµ m0 c
∂µ + ψd (x) = 0, (5.90)
iσ µ 0 ~
where the 4×4 matrices γ µ are called Dirac matrices with the explicit
expression given as follows:
! ! !
0 σcµ 0 I 0 −σ i
γµ = or γ0 = , γi = . (5.92)
σµ 0 I 0 σi 0
{γ µ , γ ν } = γ µ γ ν + γ ν γ µ = −2g µν I, (5.93)
we are able to convert the Dirac equation which is of the first order
derivative in space and time, into the second order differential equation
of the Klein-Gordon one. Applying the factor iγ ν ∂ν once more upon
the Dirac equation, we then obtain the following relation
m c 2
ν µ m0 c ν µν 0
iγ ∂ν iγ ∂µ + iγ ∂ν ψd (x) = g ∂µ ∂ν − ψd (x) = 0,
~ ~
or
m c 2
µ 0
∂µ ∂ − ψd (x) = 0, (5.94)
~
which implies that each component in the Dirac spinor satisfies the
Klein-Gordon equation.
35:17.
174 Chapter 5
The covariant formulation of the Dirac equation does not imply that
γ µ ∂µ is Lorentz invariant, namely that iγ µ ∂µ in the O-Lorentz frame
will not take the expression iγ 0µ ∂µ0 in the O0 -Lorentz frame. Once the
Dirac matrices are chosen in one Lorentz frame, they will in fact be
valid in all of the Lorentz frames. The universality of the γ µ matrices
is the unique feature of the Dirac equation. What has to be modified
in the Dirac equation under Lorentz transformation are the space-time
coordinates as well as the Dirac spinor components, namely the Lorentz
transformed Dirac equation in the O0 -system will take the following
form
m0 c 0 0
iγ µ ∂µ0 + ψd (x ) = 0, (5.95)
~
Proposition 4.
The gamma matrices γ µ are universal in all Lorentz frame,
namely the Dirac equation in another Lorentz frame, i.e. the
O0 -system always takes the same gamma matrices γ µ used in O-
system. The equation in O0 -system is expressed as
m0 c 0 0
iγ µ ∂µ0 + ψd (x ) = 0.
~
1
representation D(0, 2 ) (θ, ~ and D( 12 ,0) (θ,
~ ξ) ~ ξ),
~ i.e.
~ iγ µ ∂µ + m0 c ψd (x) = 0.
~ ξ)
D(θ,
~
35:17.
5.5 Lorentz spinors and the Dirac equation 175
~ ξ)ψ
by identifying ψd0 (x0 ) ≡ D(θ, ~ d (x) = D(θ,
~ ξ)ψ
~ d (Λ−1 x0 ).
!
1 I I
S=√ , (5.98)
2 −I I
and denote the transformed matrices by γ̃µ . Then the new set of Dirac
matrices reads as follows
γ̃ µ = Sγ µ S −1 , (5.99)
or explicitly:
! !
I 0 0 −σ i
γ̃ 0 = , γ̃ i = , (5.100)
0 I σi 0
35:17.
176 Chapter 5
this enables us to show that each component in Dirac spinor ψ̃d also
satisfies the Klein-Gordon equation.
Let us return to the p-representation of the Dirac equation expressed
in Eq. (5.88), i.e.
!
0 Pc
ψd (x) − m0 cψd (x) = 0. (5.102)
P 0
Pc ψl = 0, P ψr = 0. (5.103)
35:17.
5.6 Electromagnetic interaction and gyromagnetic ratio of the electron 177
The equations imply that the left-handed spinor and the right-handed
spinor are of opposite helicities.
Consider for the case of pure magnetic field, namely Aµ = (0, A),
then the stationary state solution of the last equation takes a much
35:17.
178 Chapter 5
( 2 )
E
(σ · π)2 − − (m0 c)2 ψr = 0, (5.108)
c
q~
= π2 − σ · (∇ × A), (5.109)
c
2
E
and the second term − (m0 c)2 can be approximated as
c
2
E E
− (m0 c)2 ' 2m0 c − m0 c = 2m0 E 0 , (5.110)
c c
1 q 2 q~
P− A − σ · H ψr = E 0 ψr . (5.111)
2m0 c 2m0 c
The second term of the last equation stands for the energy of the
intrinsic magnetic moment of the charged spin 1/2 particle µe coupled
to the magnetic field H, namely
q~ q
− σ·H =− S · H = −µe · H. (5.112)
2m0 c m0 c
35:17.
5.7 Gamma matrix algebra and PCT in Dirac spinor system 179
ge = 2. (5.114)
{γ µ , γ ν } = −2g µν I
(1) I,
(2) γ µ ,
(3) γ 5 = γ 0 γ 1 γ 2 γ 3 ,
(4) γ 5 γ µ ,
(5) σ µν = 21 [γ m u, γ ν ].
35:17.
180 Chapter 5
Class 1
It contains only the identity matrix, obtained, up to a sign, by mak-
ing the square of any of γ µ matrix. It commutes with all the gamma
matrices in 5 classes.
Class 2
It contains 4 traceless matrices with the square equaling to ±I.
Class 3
It is a traceless, diagonal matrix obtained by successive multiplica-
tion of the 4 Dirac matrices. The matrix is denoted by γ 5 with the
following properties,
(γ 5 )2 = −I,
{γ 5 , γ µ } = 0.
Class 4
It contains 4 traceless matrices with the following properties,
{γ 5 γ µ , γ 5 γ ν } = 2g µν I.
Class 5
It contains 6 matrices, also traceless, by taking different values on µ
and ν. The square of the matrix is, up to a sign, a unit matrix.
35:17.
5.7 Gamma matrix algebra and PCT in Dirac spinor system 181
or P : PS µ P −1 = S pµ = (S 0 , S), (5.116)
when S µ is an operator.
It is of the outmost interest to look into the parity transformation of
a physical quantity S which is invariant under Lorentz transformation,
namely a Lorentz scalar. We name S a scalar if it is also invariant under
parity transformation, i.e.
P.T.
P: S −→ S p = S,
or if it is an operator, then
P: PSP −1 = S p = S. (5.117)
PP P −1 = −P, (5.118)
(γ µ Pµ − m0 c)ψd = 0, (5.119a)
or (γ 0 P0 + γ · P − m0 c)ψd = 0. (5.119b)
Applying the parity operator P upon the Dirac equation above, and
inserting the identity I = P −1 P right after the momentum operator Pµ ,
we obtain
35:17.
182 Chapter 5
Proposition 5.
Let ψd (x) be the solution of the Dirac equation, and Pψd (x) be
the parity transformed Dirac spinor, i.e. ψdp (x) = Pψd (x) =
ψ(x0 , −x), then γ 0 ψ p (x) is the solution of the same Dirac equa-
tion.
35:17.
5.7 Gamma matrix algebra and PCT in Dirac spinor system 183
35:17.
184 Chapter 5
iq
iγ ∂µ + Aµ − m0 c ψ c (x) = 0,
µ
(5.123)
~c
iq
iγ µ∗ ∂µ − Aµ + m0 c ψ c∗ (x) = 0,
~c
iq
µ
iγ ∂µ − Aµ − m0 c Cψ c∗ (x) = 0,
~c
Cγ µ∗ C−1 = −γ µ ,
35:17.
5.8 Exercises 185
n m0 c o −1
T iγ µ ∂µ − T T ψ(x) = 0,
~
n m0 c o t
or −iγ 0 ∂0 + iγ · ∂ − ψ (x) = 0, (5.125)
~
where one takes the time t into −t in the equation, T ψ(x) = ψ t (x), i.e.
T.R.
T : ∂0 7−→ ∂00 = −∂0 .
Equation (5.125) can easily be converted back into the Dirac equa-
tion of the following expression
n m0 c o t
iγ 0 ∂0 + iγ · ∂ − Tψ (x) = 0,
~
Tγ 0 T−1 = −γ 0 ,
TγT−1 = γ.
5.8 Exercises
Ex 5.8.1
Show that a space-like vector is orthogonal to a time-like vector.
(Hint: if x0 > |~x|, then y 0 < |~y | for x · y = 0.)
Ex 5.8.2
Show that the sum of two time-like vectors in the same light cone is
also a time-like vector.
(Hint: make use of the Schwarz inequality.)
35:17.
186 Chapter 5
Ex 5.8.3
β1 + β2
β= .
1 + β1 β2
Ex 5.8.4
2
( 1 ,0)
2 ∗ 0 (0, 1 )∗
Verify that Dmm 0 = γ(m)D−m−m0 γ (m ) can be cast in the matrix
form as follows
! !
( 12 ,0) 0 γ( 12 ) 1 0 γ ∗ (− 21 )
D = D(0, 2 )∗ .
1
γ(− 2 ) 0 γ ∗ ( 12 ) 0
Ex 5.8.5
Verify that
! !
0 1 0 −1
σi∗ −1 = σi∗ = −σi .
−1 0 1 0
Ex 5.8.6
35:17.
5.8 Exercises 187
Ex 5.8.7
Show that the Lorentz transformation matrix element Λµν can be
expressed as follows (Hint: by Eqs. (5.77) and (5.82).)
1 n 1 1
o 1 n 1 1
o
Λµν = Tr σcµ D( 2 ,0) σν D( 2 ,0)† = Tr σ µ D(0, 2 ) σcν D(0, 2 )† .
2 2
Ex 5.8.8
Show that Dγ µ D −1 = Λµν γ ν where
1
! !
D(0, 2 ) 0 0 σcν
D= ( 12 ,0)
and γ ν = .
0 D σν 0
Ex 5.8.9
Define a new Dirac spinor as
iq
ψd0 (x) = e ~c α(x) ψd (x).
Show that the Dirac equation for a charge particle with EM interac-
tion is invariant under such gauge transformation. (Hint: with a new
vector potential A0µ (x) = Aµ (x) − ∂µ α(x).)
35:17.
This page intentionally left blank
35:17.
Bibliography
189
This page intentionally left blank
Index
191
192 Index
index
ket vector, 31
Lenz vector, 143
linear dependent, 2
linear independent, 2
null vector, 3
orthogonal vectors, 6
orthonormal set of, 5
root vector, 101, 104
state vector, 13
vector space, 1, 27
vector diagram, 104
wave
relativistic wave equation, 152,
159, 176
wave function, 13, 35, 37, 39, 46,
71
wave number, 61, 63