Symmetry

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

1

1 Symmetries in Quantum Mechanics


We work in the Schrödinger picture (state vectors change with time). In this picture

d
ih̄ |ψ(t)i = H|ψ(t)i. (1)
dt
The total derivative is used since the dependence on the coordinate variables does not appear
explicitly. Using H † = H
d
−ih̄ hψ(t)| = hψ(t)| H. (2)
dt
The solutions are straightforward:

|ψ(t)i = e−iHt/h̄ |ψ(0)i; hψ(t)| = hψ(0)| eiHt/h̄ . (3)

H is hermitian, so e±iHt/h̄ is unitary. Thus, the norm of the bra or the ket remains unchanged.
This operation performs a generalized rotation in Hilbert space.
In QM, almost all symmetry operations (the operations which leave the system invariant) are as-
sociated with some unitary operators; one of the most notable exceptions is time-reversal symmetry
(not time-translation) which involves an antiunitary operator. We will discuss that later.
Suppose Ω is a dynamical variable. How does its matrix element between two states |ψi and
hφ| changes with time? We write

d d d ∂Ω
   
hφ(t)|Ω|ψ(t)i = hφ(t) |Ω|ψ(t)i + hφ(t)|Ω| ψ(t) i + hφ(t)| |ψ(t)i
dt dt dt ∂t
1 ∂Ω
= hφ(t)|[Ω, H]|ψ(t)i + hφ(t)| |ψ(t)i. (4)
ih̄ ∂t
If Ω does not depend on t explicitly, and if [Ω, H] = 0, the l.h.s. is zero; then we say that Ω is a
constant of motion (since the result is true for any two states).
Symmetry of a system has important consequences, e.g.,
homogeneous space −→ translational symmetry −→ conservation of linear momentum;
isotropic space −→ rotational symmetry −→ conservation of angular momentum.

Consider the spatial displacement caused by the unitary operator Ur :


U
r
ψ(r) −→ ψ 0 (r) = ψ(r + a)
U
r
or ψ(r) −→ ψ(r0 ) = ψ(r − a). (5)

The first equation is from the active viewpoint where the wavefunction is changed (ψ → ψ 0 ), and
the second one is from the passive viewpoint where the coordinates are changed (r → r0 ). It does
not matter which one you choose, but you should stick to one convention consistently. We will stick
to the passive convention.
Thus, we define Ur — again, a dynamical variable producing a generalized rotation of the Hilbert
space — as follows:

Ur (a)ψ(r) = ψ(r − a)
3 3 X 3
X ∂ 1 X ∂ ∂
= ψ(r) − axi ψ(r) + ax ax ψ(r) − · · ·
i=1
∂xi 2! i=1 j=1 i j ∂xi ∂xj
−a.∇
= e ψ(r). (6)

Putting the momentum operator p = −ih̄∇, we get

Ur (a) = e−ia.p/h̄ . (7)


Quantum Mechanics II – Calcutta Univ. 2

This equation is valid for all state vectors and in all representations (and not in coordinate repre-
sentation only). Space displacements form a continuous abelian group and the components of p
are the generators for displacement along different axes. The operator Ur is unitary since p is a
hermitian operator and a is real.
Does the space-displaced state represent a proper motion of the system? In other words, does
Ur (a)ψ(r) = ψ(r − a) satisfy the corresponding Schrödinger equation? To see this, note that

d d
ih̄ ψ(r − a) = ih̄Ur (a) ψ(r) = Ur (a)Hψ(r)
dt dt
= Ur (a)HUr† (a)ψ(r − a), (8)

so the Schrödinger equation is satisfied if and only if Ur is a constant of motion:

Ur (a)HUr† (a) = H =⇒ [Ur (a), H] = 0 =⇒ [p, H] = 0. (9)

Thus, both p and H can be diagonalized simultaneously. Therefore a single state can have well-
defined eigenvalues for energy and momentum. Obviously, if a particle moves in a force field (like
the electron in an atom) its momentum is not a constant of motion, and the wavefunction does not
possess a space-displacement symmetry.
The result is true for any pair of generalized coordinate and momentum. If H is independent
of qi (i.e., there is a translational symmetry qi → qi + δqi ), then dpi /dt = 0 as ∂H/∂qi = 0. This
means that pi is a constant of motion. Thus, there is nothing inherently quantum-mechanical in it.

1.1 Time displacement


Suppose a ket ψ(t) is displaced in time so that

Ut (τ )ψ(t) = ψ(t − τ ) (10)

which gives
d
Ut (τ ) = e−τ dt . (11)
Since we are dealing with Schrödinger kets, d|ψ(t)i/dt = (1/ih̄)H|ψ(t)i. But the second derivative
is proportional to H 2 only if H is independent of time. With this assumption,

Ut (τ ) = eiτ H/h̄ . (12)

Since Ut (τ ) commutes with H, both the original and the time-displaced wavefunctions satisfy
Schrödinger equation (show explicitly).
There is absolutely no contradiction with the usual time evolution of exp(−iHt/h̄) (eq. 3).
Here the time coordinate changes, not the wavefunction, so by the application of Ut (τ ) we look
at the wavefunction at −τ (remember the basic distinction between the active and the passive
viewpoints). This explains the absence of the usual minus sign in the exponential.

1.2 Rotation and angular momentum


A rotation is defined by a linear operator R which takes a vector r to another vector r0 in such a
way that
ri0 = Rij rj (13)
where i and j can run from 1 to 3 (summation over repeated index implied from now on). Thus,
R is a 3 × 3 matrix. Since all components of r and r0 are real, R must be real.
The scalar product of two vectors r and s remains unchanged when both of them are rotated
by the same operator R; thus

r0 .s0 = ri0 s0i = Rij Rik rj sk =⇒ Rij Rik = δjk (14)


Quantum Mechanics II – Calcutta Univ. 3

where δjk is the 3 × 3 unit matrix (also known as Krönecker delta) which forces j = k. In other
words, R is orthonormal — RT R = 1, or R−1 = RT . (We will only consider proper rotations for
which det R = +1; thus, R−1 exists.)
Since you have some familiarity with the concept of Lie groups, you know that rotation group
is a Lie group. Any finite rotation can be built out of infinitesimal rotations, which means that all
elements are continuously connected to identity, i.e., no rotation. A two-dimensional rotation has
only one generator, while a three-dimensional Eulerian rotation has three generators. The latter
forms the group SO(3): the special orthogonal group of 3 × 3 matrices. All elements are real, and
orthogonality has been discussed above.
Rotation is non-abelian; two finite rotations about different axes do not commute. (The two-
dimensional rotation group is abelian; it is nothing but a U (1) symmetry.) However, rotation by an
infinitesimal amount is abelian; i.e., order of two such rotations is irrelevant as long as we neglect
terms which are more than one order in smallness. If the vector φ is of infinitesimal length, the
relation r0 = Rr may be written as
r0 = r + φ × r (15)
with
1 −φz φy
 

R =  φz 1 −φx  . (16)
−φy φx 1
(We are looking from a passive viewpoint, that’s why there is an extra minus sign compared to the
usual rotation matrix.) We now wish to find a transformation UR (φ) that changes ψ(r) to ψ 0 (r),
or, from passive viewpoint, ψ(r) to ψ(R−1 r):

UR (φ)ψ(r) = ψ(R−1 r)
≈ ψ(r − φ × r)
≈ ψ(r) − (φ × r).∇ψ(r)
i
= ψ(r) − (φ × r).pψ(r). (17)

Thus, we may put
i
UR (φ) ≈ 1 − φ.J (18)

where J, the angular momentum of the particle about the origin, is defined as r × p. (This is,
strictly speaking, not fully correct. In quantum mechanics we do not define the angular momentum
as r × p. Spin is also angular momentum, but that has nothing to do with spatial coordinates or
momenta. It is better to define angular momentum through eq. 18 so that the generator of rotation
is J.)
The three operators Jx , Jy , Jz are called generators of the infinitesimal rotations about the three
coordinate axes through the angles φx , φy , φz respectively (just like p and H are the generators for
infinitesimal space- and time-displacements). Components of J do not commute with each other;
mathematically this means that the rotation group is non-abelian. What should we do for finite
rotations? Suppose we wish to find UR (φ) for some finite rotation φ. Let us choose the axes in
such a way that the z axis is along φ. Then the effect of an infinitesimal change from φ to φ + ∆φ
can be written as
i
 
UR (φ + ∆φ) ≈ 1 − ∆φ Jz UR (φ) (19)

which, in the limit ∆φ → 0, gives
dUR (φ) i
= − Jz UR (φ) =⇒ UR (φ) = e−iφJz /h̄ , (20)
dφ h̄
since UR (0) = 1. For a general finite rotation,

UR (φ) = e−iφ.J/h̄ . (21)


Quantum Mechanics II – Calcutta Univ. 4

1.3 Commutation relations


Using [xi , pj ] = ih̄δij , [xi , xj ] = 0 and [pi , pj ] = 0, it is straightforward to show that

[Ji , Jj ] = ih̄ijk Jk (22)

where ijk = +1(−1) for any even (odd) permutation of 1,2,3, and ijk = 0 if two or more indices
are same. Eq. 22 can also be written as

J × J = ih̄J. (23)

Eq. 22 is the defining equation for angular momentum algebra.


Again, if I cannot define J = r × p, how can I get the commutation relations? Well, consider
two infinitesimal rotations, both by , but about two different axes, say x and y. I can write
2
 
1 0 0 1 − 2
 
0 
 0 1 − 2 −  , Ry () =  0
Rx () =  1 0 , (24)

2
2 2
0  1 − 2 − 0 1 − 2

so that
0 −2 0
 

Rx ()Ry () − Ry ()Rx () = 2


 0 0  = Rz (2 ) − 1. (25)
0 0 0
But the form of Rx () is known: it is exp(−iJx /h̄). Putting these forms of Rx , Ry and Rz , and
equating terms of the order of 2 , we get [Jx , Jy ] = ih̄Jz . Similarly other commutation relations
follow.
One can construct J2 , defined as

J2 = Jx2 + Jy2 + Jz2 . (26)

(Q. Show that J2 commutes with all Ji .)


Thus, J2 and one of the Ji s, say Jz , can be diagonalized simultaneously. We will work in such a
representation. There is nothing special in choosing Jz ; you could have chosen Jx or Jy , but not
more than one of them.
Define the raising and lowering operators

J+ = Jx + iJy , J− = Jx − iJy . (27)

Q. Show that
[Jz , J± ] = ±h̄J± , [J+ , J− ] = 2h̄Jz . (28)

1.4 Matrix elements


Let J2 and Jz be diagonal matrices (we will always talk about matrix representations of the rotation
group). Denote the eigenvalues of Jz by mh̄ where m is a set of dimensionless real numbers. Let
the eigenvalues of J2 be f (j)h̄2 , where f (j) is a dimensionless function of j 1 . Denoting a particular
set by j and m,

hjm|J2 |j 0 m0 i = f (j)h̄2 δjj 0 δmm0 ; hjm|Jz |j 0 m0 i = mh̄δjj 0 δmm0 , (29)

(satisfy yourself about the applicability of the δ-functions). Since J± commutes with J2 , they must
be diagonal in j, but not necessarily in m.
1
To avoid confusion, j labels the angular momentum of a state, while J is the angular momentum operator.
Quantum Mechanics II – Calcutta Univ. 5

Now, [Jz , J+ ] = h̄J+ can be written as (summation over j 0 and m0 implied)

hjm|Jz |j 0 m0 ihj 0 m0 |J+ |j 00 m00 i − hjm|J+ |j 0 m0 ihj 0 m0 |Jz |j 00 m00 i = h̄hjm|J+ |j 00 m00 i. (30)

We must have j = j 0 = j 00 ; otherwise, both sides are identically zero. Then

(m − m00 − 1)h̄ hjm|J+ |j 00 m00 i = 0. (31)

The matrix element is nonvanishing if and only if m = m00 + 1. So, the effect of J+ is raising the
value of m by 1 (and that’s why it is a raising operator):

hj m + 1|J+ |j mi = λm h̄ (32)

where λm is a function of j and m. Taking the hermitian conjugate, we get,

hj m|J− |j m + 1i = λ∗m h̄. (33)

Substituting these two equations in the last relation of eq.28, we see that the off-diagonal elements
vanish, and a typical diagonal element is

|λm−1 |2 − |λm |2 = 2m. (34)

This is a difference equation with general solution

|λm |2 = C − m(m + 1). (35)

|λm |2 is non-negative, but for sufficiently large values of m (positive or negative), the r.h.s. is
negative. The way out is to take a highest possible value of m, say m1 , for which J+ |jm1 i = 0.
This is the highest value of m in the sense that it cannot be raised any further; the ket |jm1 i is
annihilated by J+ . For this ket, λm1 = 0, and m1 is a solution of the equation
1 1
m21 + m1 − C = 0 =⇒ m1 = − + (1 + 4C)1/2 . (36)
2 2
Similarly, we can think of a ket |jm2 i (m2 is negative) which cannot be lowered any further by J− ;
it just gets annihilated. In this case, |λm2 −1 | = 0, and m2 is a solution of

(m2 − 1)m2 + C = 0 =⇒ m22 − m2 + C = 0. (37)

Since m2 is negative, m1 = |m2 |. What do we get from that?

• m can range from +m1 to −m1 , with unit step. This means that m can only be an integer
or a half-integer. Possible values are 0, ±1/2, ±1, · · ·

• C = m1 (m1 + 1). Let us rename the largest value as j; so for a given j, m can run from +j
to −j, and there are 2j + 1 such values. Essentially, m denotes the orientation of the angular
momentum vector in space w.r.t. the chosen set of axes. Thus, while j is an intrinsic property
of the wavefunction, m is not, and will in general change if we choose a different set of axes.

• Writing J2 = 21 (J+ J− + J− J+ ) + Jz2 , and taking the matrix element between hjm| and |jmi,
we get
1 1 
f (j)h̄2 = |λm−1 |2 + |λm |2 + m2 h̄2
2 2
1 
= {j(j + 1) − m(m − 1) + j(j + 1) − m(m + 1) + 2m2 } h̄2 (38)
2
which gives f (j) = j(j + 1).
Quantum Mechanics II – Calcutta Univ. 6

• That the eigenvalue of J2 is j(j + 1)h̄2 and not j 2 h̄2 is a direct consequence of the uncertainty
relation. Suppose f (j) = j 2 , and m = j. Then J2 = Jx2 + Jy2 + Jz2 leads to hjj|Jx,y |jji = 0.
Thus both Jx and Jy must be zero for such a state. But we know that this violates the
uncertainty principle since not more than one component of J may be precisely measured.
The matrix elements are now easy to write:

hj m + 1|J+ |j mi = [j(j + 1) − m(m + 1)]1/2 h̄ = [(j − m)(j + m + 1)]1/2 h̄


hj m − 1|J− |j mi = [j(j + 1) − m(m − 1)]1/2 h̄ = [(j + m)(j − m + 1)]1/2 h̄
hj m|Jz |j mi = mh̄. (39)

Q. Find hj 0 m0 |Jx |jmi and hj 0 m0 |Jy |jmi.

1.5 A brief digression


Recall Bohr’s angular momentum postulate: angular momentum of the electron around the nucleus
must be an integral multiple of h̄. However, that is not the only angular momentum of the electron.
Every particle has an intrinsic angular momentum (think of it as spinning around its own axis, like
the diurnal motion of the earth, but be careful, since this picture, though easy to understand, is not
correct in the strict quantum-mechanical sense), which is called spin, and which must be added to
the orbital angular momentum using the usual vector addition law. That is, if J, the total angular
momentum, consists of the orbital part L and the spin part S, it satisfies

J = L + S. (40)

Spin of a particle can be 0 or half-integer multiples of h̄. Particles having integral spins
(0, h̄, 2h̄, ...) are called bosons and particles having half-integer spins (h̄/2, 3h̄/2, ...) are called
fermions. Electrons, protons, neutrons all have spin h̄/2; so they are all fermions. Photon, on
the other hand, has spin h̄, so it is a boson. There is a short-lived particle created in nuclear
interactions; this is called the π-meson (there are three such mesons, with charge +e, 0, −e) which
has spin 0.
Since spin is nothing but angular momentum, it satisfies everything that the ordinary angular
momentum algebra satisfies. However, there is one point worth remembering: a fermion wavefunc-
tion, under a rotation of 2π, does not return to itself; it returns to the negative of the original
wavefunction. There is no such subtleties with bosonic wavefunctions.

1.6 Matrix representation


Jz and J2 must be diagonal matrices. For j = 1/2 (a convenient shorthand for h̄/2 — fermionic
spin), it is easy to show that
1 3
   
1 0 1 0
Jz = h̄ ; J = h̄2
2
.A (41)
2 0 −1 4 0 1
Jx and Jy can be obtained if you solve the problem given above:

1 1 −i
   
0 1 0
Jx = h̄ ; Jy = h̄ . (42)
2 1 0 2 i 0

The three matrices for Jx , Jy and Jz , apart from the factor of (1/2)h̄ in front, are known as Pauli
spin matrices, and commonly denoted by σx , σy and σz .
Q. Write down Jx , Jy , Jz , J2 for j = 1 and j = 3/2.
Q. For j = 1/2, explicitly verify
• [Ji , Jj ] = ijk Jk ;
Quantum Mechanics II – Calcutta Univ. 7

• Jx , Jy and Jz have eigenvalues ±h̄/2.

• J+ annihilates |1/2 1/2i and J− annihilates |1/2 − 1/2i.

• Jx2 + Jy2 + Jz2 = J2 .


P3
Q. Show that any 2 × 2 hermitian traceless matrix can be expressed as i=1 ai σi , where ai s are
real numbers and σi s are the Pauli spin matrices.

1.7 Unitary and Special unitary groups


Consider J = h̄/2. We have already shown

UR (φ) = e−iφ.J/h̄ . (43)

So if J is hermitian, UR is unitary (prove this).


All 2 × 2 unitary matrices form a group, matrix multiplication being the CP. This group is
known as U (2) (unitary group in 2-dimension).
As an extension of the last problem, try to show that any 2 × 2 unitary matrix (not necessarily
traceless) can be expressed as 3a=0 ai σi , where all ai s are real and σ0 is the 2 × 2 unit matrix.
P

The group of matrices UR (φ) constructed from three Pauli spin matrices is smaller than U (2),
since it contains only traceless matrices. Now,

det(eA ) = eTr A (44)

(prove this). Since the trace of each σi is zero, UR -matrices have determinant +1. This subset of
U (2) also forms a group — you try to prove it — which is called SU (2), the special unitary group
in 2-dimension.
Any n × n unitary matrix U has n real elements along its diagonal (why real?). The off-diagonal
pair of elements, symmetrically placed about the diagonal, are complex conjugates of each other.
There are (n2 − n)/2 independent off-diagonal elements, which, being complex, generate n2 − n
independent parameters. Thus, U has n2 independent parameters, and U (n) is an n2 -parameter
Lie group. There is one more constraint for SU (n) — the matrices must be traceless (so only n − 1
independent parameters along the diagonal). Thus, SU (n) is an (n2 − 1)-parameter Lie group.
Any s-parameter Lie group has s generators λi ; they specify an element, infinitesimally close
to the identity element, in terms of s infinitesimal real parameters φi :
X
1+i φi λi . (45)
i

For SU (2), the three generators are the three Pauli matrices. SU (3) has 8 generators. The
maximum number of mutually commuting generators is known as the rank of the group; for U (n)
is is n, and for SU (n) it is n − 1. SU (2) is rank-1 (only σz is diagonal); SU (3) is rank-2.
According to a theorem, the number of independent operators that can be constructed from the
generators (like J2 ) and that commute with all the generators is equal to the rank of the group.
Such operators can always be formed by taking suitable bilinear combinations of the generators —
they are known as the Casimir operators. For SU (2), the only Casimir is J2 .

1.8 Addition of angular momentum states


Let us see how the angular momenta associated with two parts of a system (e.g., two electrons in
an atom) add to form the total angular momentum of the system. The initial system is specified
by two individual angular momenta j1 and j2 and their projections are m1 and m2 . This state can
be written as |j1 m1 j2 m2 i which is a direct product of the two kets |j1 m1 i and |j2 m2 i.
Quantum Mechanics II – Calcutta Univ. 8

Clearly, all components of J1 commutes with all components of J2 ; operators related with J2
have no effect on |j1 m1 i and vice versa. The set of kets |j1 m1 j2 m2 i is taken to be orthonormal.
The total angular momentum j, being a vector, is given by

j = j1 + j2 (46)

which follows the standard rules of vector addition:

|j1 − j2 | < j < j1 + j2 . (47)

Jz , on the other hand, is a scalar operator

Jz = J1z + J2z (48)

so that the projection of angular momentum of the combined system, m, is given by

m = m1 + m2 . (49)

The maximum value of m is j1 + j2 , sum of the maximum possible values of m1 and m2 . This
occurs only for j = j1 + j2 and obviously for no other values of j. Similarly, the minimum value of
m, −j1 − j2 , occurs only for j = j1 + j2 . However, m = j1 + j2 − 1 can occur both for j = j1 + j2
and j = j1 + j2 − 1. Similarly, m = j1 + j2 − 2 can occur for three values of j. Thus, initially the
dimension of the product space was (2j1 + 1)(2j2 + 1) and finally it is 2j + 1, with j ranging from
j1 + j2 to |j1 − j2 |. It is easy to show that they are equal (prove it).
We will work with definite values of j1 and j2 and will indicate the direct product ket by |m1 m2 i
only. The final ket, |jmi, can be written as
j1
X j2
X
|jmi = |m1 m2 ihm1 m2 |jmi. (50)
m1 =−j1 m2 =−j2

Our task is to find the numbers hm1 m2 |jmi, which are called the Clebsch-Gordan (CG) coefficients.
The inverse expansion of (50) is
j1X
+j2 j
X
|m1 m2 i = |jmihjm|m1 m2 i. (51)
j=|j1 −j2 | m=−j

Q. Show that

hm1 m2 |jmihjm|m01 m02 i = δm1 m01 δm2 m02 ,


X

j,m

hjm|m1 m2 ihm1 m2 |j 0 m0 i = δjj 0 δmm0 ,


X

m1 ,m2
hjm|m1 m2 i = hm1 m2 |jmi∗ . (52)

By convention, we take the CG coefficients to be real, so that one can drop the asterisk in the
third equation of (52).
It is possible to obtain general formulae for the CG coefficients. However, a more useful pro-
cedure is to calculate them explicitly for given j1 and j2 , particularly if none of them are not too
large. We show such an explicit construction for j1 = j2 = 1.
Step 1.
|22i = |11i (53)
where the l.h.s. is for |jmi and the r.h.s. is for |m1 m2 i.
Step 2.
Quantum Mechanics II – Calcutta Univ. 9

Apply J− = J1− + J2− on both sides:


√ √
2|21i = 2|01i + 2|10i
1 1
or |21i = √ |01i + √ |10i. (54)
2 2
Step 3.
We continue applying J− on both sides. The next few steps yield
r
1 1 2
|20i = √ |1 − 1i + √ | − 1 1i + |00i,
6 6 3
1 1
|2 − 1i = √ |0 − 1i + √ | − 1 0i,
2 2
|2 − 2i = | − 1 − 1i. (55)

Note that J1− annihilates | − 1 1i and J2− annihilates |1 − 1i.


Step 4.
m = 1 can come from j = 1. Thus, the ket |j = 1, m = 1i will be a combination of |10i and
|01i, orthogonal to |j = 2, m = 1i. One finds
1 1
|11i = √ |10i − √ |01i. (56)
2 2
By convention, we take the first CG to be real and positive. Applying J− ,
1 1
|10i = √ |1 − 1i − √ | − 1 1i. (57)
2 2
Note that there is no |00i state on the r.h.s. Finally,
1 1
|1 − 1i = √ |0 − 1i − √ | − 1 0i. (58)
2 2
There is one more state, |j = 0, m = 0i, which is orthogonal to both |20i and |10i, and can be
written as
1 1 1
|00i = √ |1 − 1i + √ | − 1 1i − √ |00i. (59)
3 3 3
This completes the construction.
Q. Construct the coupled states for (i) j1 = 1/2, j2 = 1/2; (ii) j1 = 1, j2 = 1/2; (iii) j1 = 2,
j2 = 1/2.

1.9 Recursion relation of CG coefficients


Take eq. (50): |jmi = hm1 m2 |jmi|m1 m2 i, with sum over m1pand m2 . Apply J+ = J1+ + J2+
on both sides to get (with a shorthand notation f1 (j, m) = j(j + 1) − m(m + 1), f2 (j, m) =
p
j(j + 1) − m(m − 1)):
X
J+ |jmi = f1 (j, m)|j, m+1i = hm1 m2 |jmi {f1 (j1 , m1 )|m1 + 1, m2 i + f1 (j2 , m2 )|m1 , m2 + 1i} .
m1 ,m2
(60)
There are two terms on the right hand side. In the first term, replace m1 by = m1 + 1. The m01
sum over m01 will still be from −j1 to j1 , since J1+ does not change the j-value. Similarly, change
m2 → m02 = m2 + 1 in the second term, so the right hand side becomes

hm01 − 1, m2 |jmif2 (j1 , m01 )|m01 m2 i + hm1 , m02 − 1|jmif2 (j2 , m02 )|m1 m02 i.
X X
(61)
m01 m2 m1 m02
Quantum Mechanics II – Calcutta Univ. 10

Now we can safely drop the primes, and equate the coefficients of the ket |m1 m2 i on both sides
(using the fact that the kets are orthonormal):

f1 (j, m)hm1 m2 |j, m + 1i = f2 (j1 , m1 )hm1 − 1, m2 |jmi + f2 (j2 , m2 )hm1 , m2 − 1|jmi. (62)

Note that due to the shift of the summed variables, we have now m1 + m2 = m + 1. The procedure
may be repeated for J− to get

f2 (j, m)hm1 m2 |j, m − 1i = f1 (j1 , m1 )hm1 + 1, m2 |jmi + f1 (j2 , m2 )hm1 , m2 + 1|jmi. (63)

2 Symmetry and degeneracy


Suppose Ω is some symmetry operation, continuous or discrete. So [H, Ω] = 0. This means that
if H|αi = Eα |αi, H(Ω|αi) = Eα Ω|αi. So if Ω|αi is independent of |αi, the energy eigenvalue is
degenerate. If the symmetry is continuous, there are infinite energy-degenerate states. For example,
space-displacement symmetry tells that momentum eigenfunctions are degenerate (this is expected,
since the energy of a free particle depends on |p|, not p). The exception is a state which is constant
in space; here p = 0 and the state is nondegenerate. Similarly, [UR , H] = 0 says that [J, H] = 0 and
[J2 , H] = 0, so that simultaneous eigenstates of H, Jz and J2 can be formed; we have denoted them
by |njmi. It is evident that all states of the form UR |njmi are degenerate. In general, UR |njmi is
a linear combination of 2j + 1 independent states, so the degeneracy is 2j + 1-fold (possible number
of m-values). By continuously changing the rotation parameters, we get different combinations of
such 2j + 1 basis vectors.
As an example, consider an atomic electron in the potential V (r) + VLS (r)L.S. Since both
these terms are rotationally invariant (how do you show that for the second term?) we expect
a 2j + 1-fold degeneracy for each atomic level. Apply an external electric or magnetic field; the
rotational symmetry is broken, and states with different m are no longer degenerate.

3 Irreducible Tensor Operators


If the system is rotationally symmetric, [H, J2 ] = 0, and rotation cannot change the j-value of the
ket |jmi. Of course, it can change m. So if we start from a ket |jmi, what is the amplitude of the
j
state to be found in the state |jm0 i after rotation? This is given by the rotation matrix Dm 0 m (R),

defined as
j 0 −iJ.φ/h̄
Dm 0 m (R) = hjm |e |jmi. (64)
j
The matrix Dm 0 m (R) is a (2j + 1) × (2j + 1) dimensional one; this is nothing but the 2j + 1-

dimensional irreducible representation of the rotation operator UR (φ). One can form a reducible
representation by putting several IRs in bolck-diagonal form. It is a trivial exercise to check that
j
the rotation matrices Dm 0 m (R) form a group (do check it; what is the identity? inverse?).

One can write


j
|jm0 ihjm0 |UR (φ)|jmi = 0
X X
|jmi → UR (φ)|jmi = Dm 0 m (R)|jm i. (65)
m0 m0

Here we have used the fact that UR connects only states with same j and hence the completeness
j
relation can be safely used. Thus, the rotation matrix Dm 0 m (R) is the amplitude for the system to

be in the state |jm0 i when the unrotated state is |jmi.


Consider a spin-1/2 system, and a rotation by an angle α about the y-axis. Then Jy = 21 h̄σy ,
and
1 0 −iσy α/2 1 cos α2 − sin α2
 
1/2
D (α) = h , m |e | , mi = , (66)
2 2 sin α2 cos α2
since e−iσy α/2 = cos α2 1 − iσy sin α2 .
Quantum Mechanics II – Calcutta Univ. 11

If the rotation be about the z-axis, the form is straightforward (justify this):
 −iβ/2 
1/2 e 0
D (β) = , (67)
0 eiβ/2
so that different m-values do not mix. Is this expected?
We won’t construct the rotation matrices for higher j; the interested reader may look at Sakurai,
p. 194. However, for integer values of j, there is an interesting relationship between the rotation
matrices and the spherical harmonics Ylm (θ, φ). Let us explore that.
Suppose our initial ket is along the positive z direction. Call it |zi. We apply a rotation on it
such that the final ket |ni is along some arbitrary direction specified by θ, φ. But we know how to
get |ni: first rotate around the y-axis by θ and then around the z-axis by φ. So |ni = UR (φ, θ)|zi.
This we can write as XX
|ni = UR (φ, θ)|lmihlm|zi. (68)
l m

This involves a double sum, but now let us take hlm0 |ni:

hlm0 |ni =
X
l
Dm 0 m (φ, θ)hlm|zi. (69)
m
∗ (θ, φ) evaluated at θ = 0 with φ undetermined (the asterisk comes
Now hlm|zi is just a number: Ylm
because we have a bra and not a ket). At θ = 0, Ylm vanishes for all m except m = 0, so
s s
∗ 2l + 1 2l + 1
hlm|zi = Ylm (θ = 0)δm0 = Pl (cos 0)δm0 = δm0 . (70)
4π 4π
So s
∗ 2l + 1 l
Ylm 0 (θ, φ) = Dm0 0 (φ, θ). (71)

l (θ) = P (cos θ).
The m = 0 case is particularly simple: D00 l

3.1 Irreducible spherical tensors


What do we mean by the statement that V is a vector operator? (We have already seen a number
of examples: r, p, L, J.) V is a vector operator if its expectation value in any state transforms like
a classical vector Vi → Vi0 = Rij Vj . Under rotation the ket |αi changes to UR |αi, so we expect

hα|Vi |αi → hα|UR† Vi UR |αi = Rij hα|Vj |αi, (72)

but if this is true for any arbitrary ket, we must have

UR† Vi UR = Rij Vj . (73)

Consider an infinitesimal rotation about the z-axis, so that UR = 1 − iJz /h̄. We can write eq. (73)
as

Vi + [Vi , Jz ] = Rij ()Vj , (74)
ih̄
but
1 − 0
 

R() =   1 0  , (75)
0 0 1
so
  
Vx + [Vx , Jz ] = Vx − Vy , Vy + [Vy , Jz ] = Vy + Vx , Vz + [Vz , Jz ] = Vz . (76)
ih̄ ih̄ ih̄
This means that V must satisfy the commutation relation

[Vi , Jj ] = iijk h̄Vk . (77)


Quantum Mechanics II – Calcutta Univ. 12

Under a finite rotation, V should transform as exp(iJj φ/h̄)Vi exp(−iJj φ/h̄), so our only task is
to calculate commutators of Vi with Jj . Eq. (77) can be taken as the definition of a vector oper-
ator. (Show that [y, Lz ] = ih̄x or [px , Lz ] = −ih̄py are special cases of eq. (77); also check these
commutators explicitly.)
A vector is a tensor of rank 1; so how should higher rank operators transform? It is obvious
that the transformation property will define a tensor operator, but a cartesian tensor transforms
in a mixed way. Consider the 9-component cartesian tensor Ai Bj , which can be written as

1 Ai B j − Aj B i Ai B j + Aj B i 1
 
Ai Bj = A.Bδij + + − A.Bδij . (78)
3 2 2 3

The first term is a scalar, the second one is a vector (A × B), and the last one is a rank-2 symmetric
traceless tensor. They transform differently, so it is not a good idea to use cartesian tensors to
study the transformation properties. We will use the spherical tensors instead. What is a spherical
tensor of rank k? It is analogous to the spherical harmonic Ylm with l = k, and has 2k + 1 different
projections. We know how Ylm ’s transform; this allows us to write in an identical way the defining
transformation property of a spherical tensor of rank k:
k

UR† Tqk UR =
X
k k
Dqq 0 (R)Tq 0 . (79)
q 0 =−k

The angular momentum operators can be written in the spherical polar coordinate as
∂ ∂ ∂ ∂ ∂
   
Jz = −ih̄ , J+ = Jx +iJy = h̄eiφ + i cot θ , J− = Jx −iJy = h̄e−iφ − + i cot θ .
∂φ ∂θ ∂φ ∂θ ∂φ
(80)
(Strictly speaking, they hold for the orbital angular momentum components Li , but the spin com-
ponents do nothing on the spherical harmonics, so we can just add them up and construct the total
Ji .)
Thus, Jz Ylm (θ, φ) = mh̄Ylm (θ, φ), and so

[Jz , Ylm (θ, φ)]f (θ, φ) = mh̄Ylm (θ, φ)f (θ, φ). (81)

But this holds for any arbitrary function f , so

[Jz , Ylm ] = mh̄Ylm . (82)

Similarly,
q q
[J+ , Ylm ] = l(l + 1) − m(m + 1)h̄Yl,m+1 , [J− , Ylm ] = l(l + 1) − m(m − 1)h̄Yl,m−1 . (83)

In analogy, we can write for the tensor operator Tqk ,

[Jz , Tqk ] = qh̄Tqk ,


q
[J+ , Tqk ] = k
k(k + 1) − q(q + 1)h̄Tq+1 ,
q
[J− , Tqk ] = k
k(k + 1) − q(q − 1)h̄Tq−1 . (84)

From eq. (79), it is clear that under a general rotation all 2k + 1 components mix with each
other (that is why they are called irreducible operators) but they do not mix with other tensor
operators (e.g., a rank-2 tensor will not mix with a scalar or a vector). There are a number of
physical observables that can be related with the expectation values of irreducible spherical tensor
operators, the most well-known example is the 2l electric or magnetic multipole moment, which is
a rank-l tensor operator.
Quantum Mechanics II – Calcutta Univ. 13

3.2 Product of tensor operators


The close similarity between the angular momentum commutation relations and eq. (84) suggests
that tensor operators can be combined in accordance with the same rule as angular momentum
eigenstates. The product is, of course, reducible.
Let us start with two irreducible spherical tensor operators T1 kq11 and T2 kq22 . The product is a
(2k1 + 1)(2k2 + 1) dimensional reducible space. Now let us define
k1
X k2
X
Tqk = T1 kq11 T2 kq22 hq1 q2 |kqi, (85)
q1 =−k1 q2 =−k2

where the CG coefficient hq1 q2 |kqi corresponds to j1 = k1 and j2 = k2 . Let us now apply differnt
components of J on both sides of the above equation, but as usual, we will apply J on the left hand
side and J1 + J2 on the right hand side. These are operators, not kets, but our logic is the identity
[J, T1 T2 ] = [J, T1 ]T2 + T1 [J, T2 ]. Taking the commutator with Jz ,
X
[Jz , Tqk ] = {[Jz , T1 ]T2 + T1 [Jz , T2 ]} hq1 q2 |kqi
q1 ,q2
X
= (q1 + q2 )h̄T1 T2 hq1 q2 |kqi = qh̄Tqk , (86)
q1 ,q2

where we have used the fact that the CG coefficient is zero unless q = q1 + q2 . Next, we take the
commutator with J+ :

{f1 (k1 , q1 )h̄T1 kq11+1 T2 kq22 + f1 (k2 , q2 )h̄T1 kq11 T2 kq22+1 }hq1 q2 |kqi.
X
[J+ , Tqk ] = (87)
q1 ,q2

(For a definition of f1 and f2 , see eq. 60). In an analogous way to the derivation of eq. (62), we
shift q1 → q10 = q1 + 1 in the first term and q2 → q20 = q2 + 1 in the second term on the right hand
side, keeping the summation limits unchanged, and then drop the primes to get
X
[J+ , Tqk ] = h̄{f2 (k1 , q1 )T1 kq11 T2 kq22 hq1 − 1, q2 |kqi + f2 (k2 , q2 )T1 kq11 T2 kq22 hq1 , q2 − 1|kqi}. (88)
q1 ,q2

Using eq. (62) for the numbers f1 , f2 , and the CG coefficients, we get
X
[J+ , Tqk ] = f1 (k, q)h̄T1 kq11 T2 kq22 hq1 q2 |k, q + 1i
q1 ,q2
k
= f1 (k, q)h̄Tq+1 . (89)

In the last step, we have used the fact that the CG coefficient is zero unless q1 + q2 = q + 1 and
k . Similarly, we can
hence the double sum ensures that the product of T1 and T2 is nothing but Tq+1
show
[J− , Tqk ] = f2 (k, q)h̄Tq−1
k
. (90)
Thus we have shown that the 2k + 1 operators Tqk constitute an irreducible tensor operator.
One can invert eq. (85):
k
X k1X
+k2
T1 kq11 T2 kq22 = Tqk hq1 q2 |kqi. (91)
q=−k k=|k1 −k2 |

Remember that the CG coefficients are real, so hkq|q1 q2 i = hq1 q2 |kqi. The second sum (over k)
shows the reducibility of the product.
There is another (and probably more elegant) way to prove this, using eq. (79), see Sakurai, p.
236.
Quantum Mechanics II – Calcutta Univ. 14

3.3 The Wigner-Eckart theorem


In the theory of atoms and nuclei, we may have to calculate matrix elements of tensor operators with
respect to angular momentum eigenstates, like hα0 , j 0 m0 |Tqk |α, jmi. Here α and α0 denote all other
coordinates and quantum numbers, like the radial coordinates. Only the angular informations go
into jm and j 0 m0 . We have a m-selection rule: the above matrix element is zero unless m0 = q + m.
The proof follows from the relation [Jz , Tqk ] = qh̄Tqk :

hα0 , j 0 m0 |[Jz , Tqk ] − qh̄Tqk |α, jmi = (m0 − m − q)h̄hα0 , j 0 m0 |Tqk |α, jmi = 0, (92)

so the matrix element is zero unless m0 = q + m.


Now the Wigner-Eckart theorem: the matrix elements of tensor operators with respect to
angular momentum eigenstates satisfy

hα0 , j 0 ||T k ||α, ji


hα0 , j 0 m0 |Tqk |α, jmi = hjk; mq|jk; j 0 m0 i √ , (93)
2j + 1

where the reduced (or double-bar) matrix element is independent of m, m0 , and q.


This is, as far as practical applications go, one of the most important theorems in QM. Let us
first try to understand it physically.
The matrix element is a product of two factors. The first factor is a CG coefficient; it just
tells you how to add j and k to get j 0 , and depends only on the geometry of the system (how it is
oriented about the electric field, for example). It does not tell you anything about the dynamics.
The second factor contains all the dynamics, coming from the tensor operator as well as from
the radial integrals. But it is independent of the magnetic quantum numbers, i.e., it does not
depend upon the geometry of the system! So to know the dynamics, it is enough to calculate the
reduced matrix element between any two suitable states and then scale it, if necessary, with the
corresponding CG coefficients. If we want the CG coefficients to be nonzero, we immediately get
the m-selection rule discussed above, as well as the angular momentum selection rule:

|j − k| ≤ j 0 ≤ j + k. (94)

For example, with a scalar operator (k = q = 0), we must have j = j 0 , m = m0 ; it cannot change j
or m values. For a vector operator (like the electric dipole operator), we get

∆m = m0 − m = ±1, 0; ∆j = j 0 − j = ±1, 0; (95)

but the 0 → 0 transition is forbidden (you cannot add 0 and 1 to get 0). This selection rule is of
fundamental importance in the theory of radiation.
P
Proof: First note that if we have two sets of linear homogeneous equations of the form j aij xj =
P
0 and j aij yj = 0, with i < j, then we cannot solve for individual xj and yj , but we can say that
these two sets of variables are related, i.e., for all j, xj = cyj where c is a universal constant.
Now take the matrix element of the second equation of eq. (84) between hα0 , j 0 m0 | and |α, jmi.
From the application of J+ , we get

f2 (j 0 , m0 )h̄hα0 , j 0 m0 |Tqk |α, jmi = f1 (j, m)h̄hα0 , j 0 m0 |Tqk |α, jmi + f1 (k, q)hα0 , j 0 m0 |Tq+1
k
|α, jmi. (96)

Compare this with eq. (63); the coefficients are identical, with some change in the notation. To be
precise, hm1 , m2 + 1|jmi corresponds to hα0 , j 0 m0 |Tq+1
k |α, jmi. So we must have

hα0 , j 0 m0 |Tq+1
k
|α, jmi = Chm, q + 1|j 0 m0 i, (97)

where C is a universal proportionality constant independent of m, m0 and q. This constant contains



the dynamics and conventionally written as a double-bar matrix element divided by 2j + 1. This
proves the theorem.
Quantum Mechanics II – Calcutta Univ. 15

4 Discrete Symmetry: Parity


Parity or space inversion is a discrete symmetry that changes, in the passive viewpoint, a right-
handed coordinate system into a left-handed one. In the active viewpoint, the ket |αi is transformed
under the parity operator P as |αi → P |αi, so that the expectation value of r with respect to these
two states are opposite in sign:
hα|P † rP |αi = −hα|r|αi, (98)
which can be accomplished if P † rP = −r or {r, P } = 0.
Consider a position eigenket |r0 i. P reverses the position coordinates, upto a phase factor eiδ
(δ real), so
P |r0 i = eiδ | − r0 i. (99)
To prove this assertion, note that

rP |r0 i = −P r|r0 i = (−r0 )P |r0 i. (100)

So P |r0 i is an eigenket of the position operator r with eigenvalue −r0 , so it must be the position
eigenket | − r0 i upto a phase factor. By convention we take eiδ = 1; so P 2 |r0 i = |r0 i, or P 2 always
has eigenvalue +1. Thus, P −1 = P † = P , with possible eigenvalues +1 and −1 (remember that
all kets are eigenkets of P 2 , but need not be of P ; an example is the momentum eigenket |pi,
whose wavefunction is just the plane wave: the reason is that P flips momentum). Just like r, the
momentum operator p also anticommutes with P . Physically, this means that translation (whose
generator is momentum) followed by P is P followed by translation in the opposite direction.
Thus, r and p are odd under parity. Angular momentum, defined as L = r × p, is even:
[P, L] = 0. In quantum mechanics, we postulate that [P, UR ] = 0, so [P, J] = 0 since J is the
generator of rotation. But J = L + S, so spin is also even under parity. From the discussion of the
previous section, we saw that r, p, and J are all vectors, or spherical tensors of rank 1. But the
first two are polar vectors while the last one is an axial vector.
The operators L.S or r.p are ordinary scalars. But the operator S.r picks up a minus sign
under parity:
P −1 S.rP = −S.r. (101)
This transformation property defines a pseudoscalar operator.

4.1 Wavefunctions under parity


The wavefunction ψ(r) can be even or odd under P depending on whether its state ket |αi is
even or odd (remember ψ(r = hr|αi). To prove this, suppose that |αi is an eigenket of parity, so
that P |αi = ±|αi. Its corresponding wavefunction will be hr|P |αi = ±hr|αi. But we also have
hr|P |αi = h−r|αi, so |αi is even under parity if ψ(r) = ψ(−r) and odd if ψ is odd.
What about the orbital angular momentum eigenkets? Note that hr|α, lmi = Rα (r)Ylm (θ, φ).
Under P , r → r, θ → π − θ, φ → π + φ, so that Ylm → (−1)l Ylm . Thus, orbital angular momentum
wavefunctions (i.e., spherical harmonics) are even or odd depending on whether l is even or odd.
Now we state an important theorem. Suppose that the Hamiltonian is invariant under parity
(most of the physical Hamiltonians are, except that of the weak interaction) so that [H, P ] = 0, and
let |ni be a nondegenerate eigenket with energy eigenvalue En . Then |ni is also a parity eigenket.
The proof is extremely simple and goes like this.
Consider the ket |n± i = 12 (1 + P )|ni. This is evidently a parity eigenket, with eigenvalue ±1.
This is also an energy eigenket with eigenvalue En . But |ni is nondegenerate, so |n± i must be
equal to |ni upto a multiplicative constant, and |ni has parity eigenvalue ±1.
Consider the harmonic oscillator whose energy levels are nondegenerate. The ground state has
even parity, and the n-th excited state has parity (−1)n . The nonrelativistic hydrogen atom, whose
energy levels depend only on the principal quantum number n and hence degenerate, does not
satisfy this theorem: one can construct a state c1 |2si + c2 |2pi, which has the same energy E2 but
Quantum Mechanics II – Calcutta Univ. 16

is not an eigenfunction of parity. Similar considerations apply for momentum eigenkets too. They
are not parity eigenkets, since |pi and | − pi have same energy and hence degenerate.

4.2 Selection rule


The rule says that parity-odd (even) operators always connect states of opposite (same) parity.
Thus, if |αi and |βi are parity eigenkets, hα|r|βi = 0 unless they have opposite parities. (Try
to prove this, by writing hα|r|βi = hα|P −1 P rP −1 P |βi.) This is important in studying radiative
transitions. For example, an electric dipole transition between two same-parity states is forbidden.
(From this, and the theorem that we have just discussed, show that if H is invariant under parity,
nondegenerate eigenstates cannot have a permanent electric dipole moment.)

5 Discrete Symmetry: Time reversal


Time reversal, or reversal of motion to be precise, is nothing but running a film backward. If
the Hamiltonian is symmetric under time reversal, you cannot distinguish between forward and
backward runnings of the film. Mathematically speaking, if mr̈ = −∇V (r) has a solution r(t),
then r(−t) is also a solution. Note that the force should not be a dissipative one, like friction!
For classical electrodynamics, time reversal means E → E, B → −B, ρ → ρ, j → −j, and
v → −v. This keeps Maxwell’s equations as well as Lorentz force law invariant. This is also
intuitively clear: since t → −t, both current and velocity flip sign, and magnetic field is caused by
tiny currents, so that also flips. (How should φ and A change?)
Consider the Schrödinger equation:
!
∂ψ h̄2 2
i = − ∇ +V ψ, (102)
∂t 2m

and let ψ(r, t) be a solution. Obviously, ψ(r, −t) is not a solution, since we have a first derivative in t.
But take the complex conjugate; then ψ ∗ (r, −t) is a solution. This is also intuitively obvious: both
ψ(r, t) and ψ ∗ (r, −t) go like exp(−iEt/h̄). Thus, time reversal has something to do with complex
conjugation. The time-reversed wavefunction corresponding to ψ(r, t) = hr|αi is hr|αi∗ = hα|ri.
For any symmetry operation, we expect the inner product to be preserved: if |αi → |α̃i,
|βi → |β̃i, then hα̃|β̃i = hα|βi. This follows from the property of unitary operators: hα̃|β̃i =
hα|U † U |βi = hα|βi. However, for time reversal we cannot take the symmetry operation to be
unitary (we’ll see later why), and we must impose a weaker condition: |hα̃|β̃i| = |hα|βi|. This
admits another possibility, namely, hα̃|β̃i = hα|βi∗ = hβ|αi (i.e., ket and bra are interchanged).
Consider an operator θ. If this is antiunitary, then for the transformations |αi → |α̃i = θ|αi,
|βi → |β̃i = θ|βi, we must have

hα̃|β̃i = hα|βi∗ , θ (c1 |αi + c2 |βi) = c∗1 θ|αi + c∗2 θ|βi. (103)

The last equation alone defines an antilinear transformation.


Any antiunitary operator can be written as θ = U K where U is a unitary operator and K is
the complex conjugation operator defined as Kc|αi = c∗ K|αi.
The proof is easy. Putting θ = U K in the second equation of (103), we get

U K (c1 |αi + c2 |βi) = c∗1 U K|αi + c∗2 U K|βi = c∗1 θ|αi + c∗2 θ|βi. (104)

Before proving the first relation of (103), let us see what θ|αi is. |αi can be expanded in terms of
the basis kets |ii: |αi = i |iihi|αi. But |ii is a column vector with one 1 and rest 0, so K|ii = |ii.
P

Thus
hi|αi∗ U K|ii =
X X
|αi → |α̃i = hα|iiU |ii. (105)
i i
Quantum Mechanics II – Calcutta Univ. 17

Similarly, |β̃i = †.
j hβ|jiU |ji, so hβ̃| = j hj|βihj|U
P P
Thus,

hj|βihj|U † U |iihα|ii = hα|iihi|βi = hα|βi = hβ|αi∗ .


X X
hβ̃|α̃i = (106)
i,j i

5.1 Time reversal operator


Let T be the antiunitary time-reversal operator. Since this changes t to −t, momentum p and
angular momentum J should change sign under T , but position r should not. Now consider the ket
T |αi at t = 0. The application of T changes the momentum and angular momentum of |αi. What
should it be at some later infinitesimal time δt? This is given by (1 − iHδt/h̄)T |αi. If the motion is
symmetric under time-reversal, this should be the same as T (1 − iH(−δt)/h̄)|αi. So −iHT = T iH.
If T is unitary, we can cancel the i’s and get HT = −T H.
This is clearly nonsensical. Why? Well, this means that if |ni is an eigenket of H with energy En ,
so is T |ni with an eigenvalue −En . But a free particle cannot have negative energy eigenvalues!
Moreover, we expect p to change sign under T , but not p2 , but now we face a situation where
T −1 p2 T = −p2 . This tells us that we cannot take T to be unitary. Let us proceed with the
antiunitary option. Then −iHT = T iH means HT = T H (note the antilinear property of T ,
which changes the sign of i).
Let us first see how operators change under T . For the transformations |αi → |α̃i = T |αi,
|βi → |β̃i = T |βi, we first establish an important identity:

hβ|O|αi = hα̃|T O† T −1 |β̃i, (107)

where O is a linear operator. To prove this let us define |γi ≡ O† |βi, so that hγ| = hβ|O. Thus,

hβ|O|αi = hγ|αi = hα̃|γ̃i = hα̃|T O† |βi = hα̃|T O† T −1 T |βi = hα̃|T O† T −1 |β̃i. (108)

For hermitian observables O = O† . An hermitian operator is even (odd) under T if T OT −1 =


O(−O). We expect T pT −1 = −p, and same for J, but T rT −1 = r. This is nice, since the
fundamental commutation relation [xi , pj ] = ih̄δij is unaffected. To see this, apply the commutation
(which is an operator relation) on an arbitrary ket |αi, and then apply T :

T [xi , pj ]T −1 T |αi = T ih̄δij |αi ⇒ [xi , −pj ]T |αi = −ih̄δij T |αi, (109)

so that the commutation relation is unchanged. (Similarly, show that if [Ji , Jj ] = ih̄ijk Jk is to
remain unaffected, J must be odd under T .)

5.2 Wavefunctions under time reversal


Consider a spinless (spin brings extra complications) single-particle system to be in a state |αi at
t = 0. If its wavefunction is hr|αi, then
Z
|αi = d3 r |rihr|αi. (110)

Applying the time-reversal operator,


Z Z
T |αi = d3 r T |rihr|αi∗ = d3 r |rihr|αi∗ , (111)

where we use the convention T |ri = |ri. Thus under T , ψ(r) → ψ ∗ (r). The spherical har-
monic Ylm (θ, φ) gets complex conjugated, but that is just a multiplication by (−1)m , coming from
exp(−imφ).
If the Hamiltonian H is invariant under T and if the energy eigenket |ni is nondegenerate, then
it is real (upto a phase factor that does not depend on the position coordinate r). The proof goes
like this. Since HT = T H, T |ni is also an eigenket with energy En . But |ni is nondegenerate, so
Quantum Mechanics II – Calcutta Univ. 18

|ni and T |ni must represent the same state. The corresponding wavefunctions are hr|ni and hr|ni∗ ,
so the wavefunction must be real, apart from a position-independent phase factor.
Thus a nondegenerate bound state wavefunction is always real. For the hydrogen atom, if l 6= 0,
m 6= 0, the wavefunction is complex, ∼ exp(−imφ), but the state is degenerate. Similarly, the plane
wave is complex, but exp(ip.r/h̄) is degenerate with exp(−ip.r/h̄).

5.3 T for nonzero spin particles: Kramers degeneracy


It is slightly more trickier to discuss the transformation of spin under T . Since spin is nothing but
some angular momentum, we expect T ST −1 = −S. It is possible to construct the representation
matrices of the spin generators in such a way that Sx and Sz are real while Sy is imaginary. We
can write T = U K, but K does nothing on real matrices, so

Sx U = −U Sx , Sz U = −U Sz . (112)

On the other hand, Sy U K = −U KSy = U Sy K, so Sy U = U Sy . Considering the commutation


relation [Si , Sj ] = ih̄ijk Sk , we see that in this particular representation U must be some unitary
operator involving Sy alone. We take U = exp(−iπSy /h̄), so that this is a rotation by π along the
spin y axis, and hence flips the sign of Sx and Sz . Thus,

T = e−iπSy /h̄ K. (113)

We repeat that this is true only in the representation where Sy is imaginary and other components
are real. For spin-1/2 particles, Sy = 12 h̄σy , so

T = −iσy K. (114)

(Prove this, by expanding exp(−iπσy /2) and collecting real and imaginary terms.)
Eq. 114 gives rise to a remarkable representation-independent result:

T 2 = (−iσy K)(−iσy K) = −1, (115)

as K passing through two imaginary numbers brings a factor of (−1)2 and K 2 = 1.


For any general value of S, we get T 2 = e−2iπSy /h̄ (the exponential is purely real and commutes
with K). This is a rotation by an angle 2π; for a spin-1/2 particle, this brings an extra minus sign,
so T 2 = −1, but for integer spin particles, T 2 = 1. In fact, we have a general result: T 2 = (−1)2j .
This is independent of any representation; the system may not even be in an eigenstate of J.
Suppose we have a system where [H, T ] = 0. Let |ni and T |ni be the energy eigenkets of
the original and the time-reversed state. Obviously, both have same energy eigenvalue En . The
question is, are these two states identical? If they are, we can write T |ni = eiδ |ni, where δ is a
r-independent phase factor. Then

T 2 |ni = T eiδ |ni = e−iδ T |ni = e−iδ eiδ |ni = |ni. (116)

But this can be true only for integer-j valued systems! Thus, if the total number of fermions be odd
in a system so that T 2 has eigenvalue −1, the two states |ni and T |ni are bound to be different,
and the system must possess at least a twofold degeneracy, no matter how complicated H might
be. This is known as Kramers degeneracy. For example, a system with an odd number of electrons
in an electric field (however complicated) must have at least a twofold degeneracy. But this can
be lifted if we introduce interactions that do not respect T , i.e., if the original H is even under T
these new terms should be odd, like S.B or L.B, where B is an externally applied magnetic field
2 . With such a T -odd interaction, the two levels will get split.

2
Under T , B does not change; this is an external field, and so we assume that T does not affect the electrons in
the magnet. This is not contradictory to what we said about the invariance of Maxwell’s equations. There we applied
T to the whole universe.

You might also like