QM LecNotes 02
QM LecNotes 02
By
A. K. Kapoor
School of Physics
University of Hyderabad
Hyderabad 500046, INDIA
Version 1.1
November 4, 2010
3 Variation Method 32
§ 1 Variation Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
§ 1.1 Ritz variation method . . . . . . . . . . . . . . . . . . . . . . 33
6 Scattering Theory 73
§ 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
§ 2 Scattering in classical and quantum mechanics . . . . . . . . . . . . . 73
§ 3 Integral Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
§ 4 Integral Equation for Scattering . . . . . . . . . . . . . . . . . . . . . . 79
§ 5 Born Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
§ 6 Problems and Cross Links . . . . . . . . . . . . . . . . . . . . . . . . 85
Version 1.1
⊠2) It is useful to introduce the operators J± = Jx ± iJy and these operators obey
commutation relations
[J+ , J− ] = 2~Jz , (1.4)
[Jz , J+ ] = ~J+ , (1.5)
[Jz , J− ] = −~J− . (1.6)
Also J~2 commute with all the three components of the angular momentum
Jx , Jy , Jz .
⊠3) Some useful relations are
J+ J− = J 2 − Jz2 + ~Jz , (1.7)
J− J+ = J 2 − Jz2 − ~Jz , (1.8)
1
J2 = (J+ J− + J− J+ ) + Jz2 . (1.9)
2
⊠4) One can find simultaneous eigenvectors of J~2 and any one component, Jn =
~ along a fixed direction given by the unit vector n̂.
n̂ · J,
⊠5) The eigenvalues of J~2 are j(j + 1)~2 where j can be integer of half integer.
The eigenvalues of a component of Jn take values from −j to j in step of 1.
⊠6) It is customary to denote the normalised, simultaneous,eigenvectors of J~2 and
Jz by |jmi so that
J~2 |jmi = j(j + 1)~|jmi, (1.10)
Jz |jmi = m~|jmi. (1.11)
⊠8) The operators J± annihilate |j, ±ji because the Jz value cannot be increased
beyond j nor can it decrease beyond −j.
J+ |j, ji = 0 J− |j, −ji = 0. (1.13)
⊠9) The above results are applicable to operators satisfying angular momentum
commutation relations except that the half integral values are ruled out for
the orbital angular momentum, beacause of additional requirement of single
valuedness of the wave function.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : .................................. ◭ QM-08-Sum;
(b) All the three components of J~ commute with J~12 and J~22 .
(c) It follows from (b) that J~ 2 commutes with J~12 and J~22 .
The components of total J~ satisfy angular momentum commutation relations and
hence the operators J 2 and Jz have simultaneuos eigenvectors and have eigenvalues
as shown in table below:
Angular momentum eigenvalues
Operator Eigenvalues Remarks
The vectors |j1 , j2 ; m1 m2 i listed in the second column of the first row are just the
(1)
direct product of vectors |j1 m1 i and |j2 m2 i which are eigenvectors of J~ (1)2 , Jz and
(2)
of J~ (2)2 , Jz , respectively. The set of all the vectors
for different m1 , m2 values form a basis in the tensor product V1 ⊗V2 of spaces V1 , V2
spanned by |j1 m1 i and |j2 m2 i.
The problem of addition of angular momenta consists of the following questions
(i) For a state |j1 j2 ; m1 m2 i, which has definite values j1 , m1 and j2 , m2 for the in-
dividual values of the square, and the z− component of the angular momenta,
what values of square of total angular momentum J~2 and Jz are allowed?
(ii) The second part of the problem is to construct the transformation matrix con-
necting the two sets of simultaneous eigenvectors, one set for each commuting
set.
The states |j1 j2 JMi can be expressed as linear combination of states
|j1 j2 ; m1 m2 i and vice versa
X
|j1 j2 m;1 m2 i = C(JM; j1 j2 m1 m2 )|j1 j2 JMi (1.18)
JM
Conversely
X
|j1 j2 JMi = hj1 j2 ; m1 m2 |j1 j2 , JMi |j1 j2 ; m1 m2 i. (1.21)
m1 ,m2
m1 +m2 =M
Since the values of j1 , j2 are fixed and common for all the states, we shall frequently
suppress these labels and write the kets |j1 j2 JMi and |j1 j2 ; m1 m2 i as |JMi and
as |; m1 m2 i respectively. The coefficients hj1 j2 , JM|j1 j2 ; m1 m2 i , written in short
as hJM|j1 j2 ; m1 m2 i , are known as Clebsch Gordon coefficients and expressions are
tabulated in books. The problem of addition of angular momenta consists of finding
allowed range of values of J, M and obtaining expressions for the Clebsch Gordon
coefficients.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
13001;
⊠2) The total angular momentum J 2 can take J(J + 1)~2 , where the minimum
value of J is |j1 − j2 | and the maximum value is (j1 + j2 ).
⊠3) The allowed values of J are all values in the range |j1 − j2 | to (j1 + j2 ) in steps
of unity.
The above statements are equivalent to saying that the Clebsch Gordon coefficients
are zero when any one of the above conditions is not satisfied. In addition one must
remember the angular momentum results that value of M must always be between
−J and J, and J must be positive.
M = m1 + m2 :
(1) (2)
The vectors |; m1 m2 i are eigenvectors of Jz , Jz . We shall now show that they are
eigenvectors of Jz with eigenvalue m1 + m2 .
Jz |; m1 , m2 i = Jz(1) + Jz(2) |; m1 , m2 i = (m1 + m2 )|; m1 , m2 i. (1.22)
hJM|; m1 m2 i = M 6= m1 + m2 (1.23)
−j2 ≤ J − j1 ≤ j2 (1.24)
−j1 ≤ J − j2 ≤ j1 (1.25)
These two conditions, Eq.(1.24) and Eq.(1.25), are equivalent to the requirement
J ≥ |j1 − j2 |. Since maximum possible value of M = m1 + m2 is j1 + j2 we must
have J ≤ j1 + j2 . Thus the total angular momentum is constrained to lie between
|j1 − j2 | and j1 + j2 . The three numbers J, j1 , j2 should be such that they satisfy
triangle inequalities
|j1 − j2 | ≤ J ≤ j1 + j2 . (1.26)
It should be remarked that Eq.(1.24)-Eq.(1.26), written as conditions on j1 and j2 ,
are equivalent to each of the following two alternate forms
|J − j2 | ≤ j1 ≤ J + j2 and |J − j1 | ≤ j2 ≤ J + j1 . (1.27)
These inequalities, known as triangle inequalities, state that the allowed combina-
tions of values of j1 , j2 and J are precisely those which can represent lengths of three
sides of a triangle.
(b) m1 = j1 , m2 = j2 − 1
What are the corresponding J values? One combination of these two states must be
the M = j1 + j2 − 1 partner of the state Eq.(1.28); and the other linear combination
can only correspond to the next value J = j1 + j2 − 1. Continuing in this way, we
now consider M = j1 + j2 − 2 which will come from three sets of m1 , m2 values, i.e.
(b) m1 = j1 − 1, m2 = j2 − 1, and,
(c) m1 = j1 , m2 = j2 − 2
Two linear combinations of the three states |j1 j2 ; m1 m2 i, corresponding to the set of
above m1 , m2 values, will correspond to the J values j1 +j2 , j1 +j2 −1 already found;
a third linear combination must therefore correspond to the value J = j1 + j2 − 2.
Proceeding in this fashion we see that the successive J values differ by one. How
are we sure that all the J values have been correctly identified ?. We will now count
the number of states in two different ways to confirm the conclusion that J takes
all the values in the allowed range from |j1 − j2 | to (j1 + j2 ).
and n o
|j1 j2 ; JMiJ = j1 − j2 , · · · , j1 + j2 , M = −J, · · · , J (1.30)
It is easily seen that the total number of vectors in two bases are equal. The
number of elements in the set (1.29) is (2j1 + 1)(2j2 + 1) and in the set (1.30) is also
the same
j1 +j2
X
(2J + 1) = (2j1 + 1)(2j2 + 1). (1.31)
J=|j1 −j2 |
The two bases in Eq.(1.29) and Eq.(1.30) are orthonormal and hence the transforma-
tion connecting the two must be a unitary transformation and the Clebsch Gordon
coefficients must satisfy the relations
X
hj1 j2 ; m1 m2 |j1 j2 JMi hJM|j1 j2 ; m′1 m′2 i = δm1 m′1 δm2 m′2 (1.32)
JM
X
hJM|j1 j2 ; m1 m2 i hj1 j2 ; m1 m2 |j1 j2 J ′ M ′ i = δJJ ′ δM M ′ (1.33)
m1 ,m2
These relations can also be seen as a consequence of the completeness formula such
as X
|JMi hJM| = Iˆ (1.34)
JM
10
1
m2 = 2
m2 = − 21
s s
1 1
1
j1 + M + 2
j1 − M + 2
j = j1 + 2
2j1 + 1 2j1 + 1
s s
1 1
1
j1 − M + 2
j1 + M + 2
j = j1 − 2
−
2j1 + 1 2j1 + 1
m2 = 1 m2 = 0 m2 = −1
s s s
(j1 + M)(j1 + M + 1) (j1 − M + 1)(j1 + M + 1) (j1 − M)(j1 − M + 1)
j = j1 + 1
(2j1 + 1)(2j1 + 2) (2j1 + 1)(j1 + 1) (2j1 + 1)(2j1 + 2)
s s
(j1 + M)(j1 − M + 1) M (j1 − M)(j1 + M + 1)
j = j1 − p
2j1 (j1 + 1) j1 (j1 + 1) 2j1 (j1 + 1)
s s s
(j1 − M)(j1 − M + 1) (j1 − M)(j1 + M) (j1 + M)(j1 + M + 1)
j = j1 − 1 −
2j1 (2j1 + 1) j1 (2j1 + 1) 2j1 (2j1 + 1)
11
and using Eq.(1.35) and Eq.(1.37) we get two relations one for J+ and
p
(J − M)(J + M + 1)hj1 j2 ; m1 m2 |JM + 1i
p
= hj1 j2 ; m1 − 1m2 |JMi (j1 + m1 )(j1 − m1 + 1)
p
+hj1 j2 ; m1 m2 − 1|JMi (j2 + m2 )(j2 − m2 + 1) (1.39)
We will make repeated use of the results Eq.(1.39),Eq.(1.40) given above. These
equations can be used successively with M = J, J − 1, . . . to compute the Clebsch
Gordon coefficients. The restrictions ⊠1) − ⊠4), given on page 8, on the allowed
values of the total angular momentum J will be derived by considering the matrix
elements hj1 j2 ; m1 m2 |Jz |JMi,hj1 j2 ; m1 m2 |Jz |JMi and hj1 j2 ; m1 m2 |J± |JMi and by
repeated use of Eq.(1.39) and Eq.(1.40) . We shall suppress j1 j2 and use the
(1) (2)
shorthand notation |; m1 m2 i to denote the eigenvectors |j1 j2 m1 m2 i of Jz , Jz . On
the other hand the eigenvectors of Jz , |j1 j2 JMi, will be written as |JMi.
12
we obtain
hJM|Jz(1) + Jz(2) − Jz |j1 j2 ; m1 m2 i = 0 (1.43)
Therefore,
(m1 + m2 − M)hJM|j1 j2 ; m1 m2 i = 0 (1.44)
Thus if M 6= m1 + m2 , the Clebsch Gordon coefficient hJM|j1 j2 ; m1 m2 i has to be
zero. In other words, a nonzero value of hJM|j1 j2 ; m1 m2 i is possible only when
M = m1 + m2 (1.45)
0 × h; m1 m2 |J(M + 1)i
p
= h; (j1 − 1)m2 |JJi 2j1
p
+h; j1 (m2 − 1)|JMi (j2 + m2 )(j2 − m2 + 1) (1.46)
13
We can continue in this fashion. We see that the Clebsch Gordon coefficient for
different pairs of values of M, m1 are known in terms of a single coefficient for
M = J, m1 = ji as follows
Eq.(1.46) m1 = j1 − 1, M = J m1 = j1 , M = J
Eq.(1.47) m1 = j1 , M = J − 1 m1 = j1 , M = J
Eq.(1.48) m1 = j1 − 1, M = J − 1 m1 = j1 , M = J − 1 and m1 = j1 − 1, M = J
Eq.(1.49) m1 = j1 − 1, M = J − 1 m1 = j1 − 1, M = J
Eq.(1.50) m1 = j1 , M = J − 2 m = j1 , M = J − 1
−j1 ≤ J − j2 ≤ j1 (1.52)
14
b) If We add angular momenta j1 and j2 = 1, the possible values of the total angular
momentum are j1 − 1, j1 , and j1 + 1.
c) The minimum value of j which when added to itself will give rise to a given value J
for the total angular momentum is J2 .
Explicit values of the Clebsch Gordon coefficients can be obtained by following
the procedure in § 5. This is not the only procedure available to get the values of ←
the Clebsch Gordon coefficients. In the solved problem below we show, by means
of an example of adding angular momenta j1 = 1 and j2 = 1/2, how the Clebsch
Gordon coefficients could be obtained using the properties of angular momentum
ladder operators and orthogonality of states with different values of J.
§ 6.2 Problem.
Enumerate all possible states with definite JM values for the case j1 = 1, j2 = 21 .
Using properties of J− and orthogonality, obtain different states |JMi in terms of
the states |j1 j2 ; m1 m2 i.
Solution : Let |1 12 ; m1 m2 i denote the state |1m1 i| 21 m2 i eigenstate of Jz(1) and Jz(2) and
|JM i will be used to denote the states with definite values of total angular momentum.
This completes the construction of the states |JM i for J = 32 .
1
1 Since j1 = 1 and j2 = 2 the only possible values of J are 12 , 32 .
2 The state, |1 12 ; 1 12 i, with highest values for m1 , m2 , corresponds to the highest value
of both J and M . Thus
| 23 32 i = |1 21 ; 1 12 i (1.53)
(1) (2)
3 Next apply J− = J− + J− on Eq.(1.53) and use Eq.(1.35) and Eq.(1.36) to get
q √ q
3 5 3 1 31 1 1 1 3 1 1 1 1
·
2 2 − · |
2 2 22 i = 1.2 − 1.0|1 2 ; 0, 2 i + 2 · 2 − 2 · (− 2 )|1 2 ; 1, − 2 i (1.54)
We thus arrive at q q
| 32 21 i = 2 1
3 |; 0, 2 i +
1 1
3 |; 1, − 2 i (1.55)
(1) (2)
4 Applying J− = J− + J− on Eq.(1.55) again we get the state with total M = − 21 .
hq q i
(1) (2)
J− | 32 12 i = J− + J− 2
3 |; 0, 1
2 i + 1
3 |; 1, − 1
2 i
hq q i
(1) 2 1 1 1
= J− 3 |; 0, 2 i + 3 |; 1, − 2 i
hq q
(2) 2 1 1 1
+J− 3 |; 0, 2 i + 3 |; 1, − 2 i (1.56)
15
where the last term in Eq.(1.56) vanishes because the minimum value of m2 is − 12 .
On simplifying we get
q q
| 32 , − 21 i = 13 |; −1, 21 i + 23 |; 0, − 12 i (1.58)
| 32 , − 32 i = |; −1, − 21 i (1.59)
The above relation can also be written down directly because for the state with J =
M = − 32 would involve only one combination of m1 , m2 , viz., m1 = −1, m2 = − 21
and hence by normalisation we must have Eq.(1.59).
6 We now consider the two states M = 12 for J = 12 . The state | 12 21 i would be linear
combination of the states with (m1 , m2 ) values given by (0, 12 ) and (1, − 21 ) and hence
The same values of (m1 , m2 ) also appear in the state J = 32 , Eq.(1.55). The states
in Eq.(1.55) and Eq.(1.60) being states with different total angular momenta should
be orthogonal. Therefore orthogonality gives
q q
2 1
3α + 3β = 0 (1.61)
We shall now take up an example of constructing the states |JMi using the tables
of Clebsch Gordon coefficients. There two tables of Clebsch Gordon coefficients are
given on page 11. Note that the first table is for j2 = 21 and the second one for
j2 = 1. For j2 = 12 , the two columns correspond to the two values, 21 and − 12 , of
m2 . The two rows correspond to the tow possible values of total angular momentum
J = j1 + 12 and J = j1 − 21 . Similarly the second table, corresponding to j2 = 1, has
three columns for the three values m2 = 1, 0, −1 and the three rows correspond to
three allowed values J = j1 + 1, j1 , j1 − 1.
16
1 5
Solution : Since j2 = 2the first table is needed here. The allowed values of J are 2
3 5 3
and 2. For all J = states we must use the first row of the first table and for J =
2 2
states the second row should be used.
1 The state with highest values J = 52 and M = 25 can be written down directly as
there is only one possible set of values (m1 , m2 ) = (2 12 ) is allowed. Hence
| 25 52 i = |; 2, 21 i (1.64)
5 3
2 For the next state with J = 2 and M = 2 the possible values of (m1 , m2 ) are (1, 12 )
and (2, − 12 ) and hence
| 25 , 32 i = α1 |; 1, 21 i + α2 |; 2, − 12 i (1.65)
5 1
3 For the next state with J = 2 and M = 2 the possible values of (m1 , m2 ) are (0, 12 )
and (1, − 12 ) and hence
| 25 , 32 i = α3 |; 0, 21 i + α4 |; 1, − 12 i (1.68)
and the coefficients α1 , α2 are again to be read from the first row.
s s r
j1 + M + 12 2 + 12 + 12 3
α3 = = = ; (1.69)
2j1 + 1 2.2 + 1 5
s s r
1
j1 − M + 2 2 − 12 + 12 2
α4 = = = (1.70)
2j1 + 1 2.2 + 1 5
4 For the next state with J = 52 and M = − 21 the possible values of (m1 , m2 ) are
(−1, 12 ) and (0, − 21 ) and hence
| 52 , − 12 i = α5 |; −1, 21 i + α6 |; 0, − 12 i (1.71)
and the coefficients α1 , α2 are again to be read from the first row.
s s r
j1 + M + 12 2 − 12 + 12 2
α5 = = = ; (1.72)
2j1 + 1 2.2 + 1 5
s s r
j1 − M + 12 2 + 12 + 12 3
α6 = = = (1.73)
2j1 + 1 2.2 + 1 5
17
and the coefficients α1 , α2 are again to be read from the first row.
s s r
1
j1 + M + 2 2 − 32 + 12 1
α7 = = = ; (1.75)
2j1 + 1 2.2 + 1 5
s s r
j1 + M − 12 2 + 32 + 12 4
α8 = = = (1.76)
2j1 + 1 2.2 + 1 5
6 Finally the state | 52 , − 52 i has the highest value of J and the lowest value of M , and
easily written as
| 25 , − 52 i = |; −2, − 21 i (1.77)
3
7 Now we take up the construction of states with J = 2 and M = 32 , 12 , − 12 , − 23 we
write
| 32 , 3
2i = β1 |; 1, 21 i + β2 |; 2, − 21 i (1.78)
3 1
|2, 2i = β3 |; 0, 21 i + β4 |; 1, − 21 i (1.79)
| 2 , − 12 i
3
= β5 |; −1, 12 i + β6 |; 0, − 21 i (1.80)
| 32 , − 32 i = β7 |; −2, 12 i + β8 |; −1, − 21 i (1.81)
(1.82)
where β1 , β3 , β5 , β7 are to be taken from the entries in the second row, first column
and β2 , β4 , β6 , β8 are to read from the second row, second column of the first table
with appropriate values of M . Thus
s r s r
j1 − M + 12 1 j1 + M + 21 4
β1 = − =− ; β2 = = (1.83)
2j1 + 1 5 2j1 + 1 5
3 3
M= 2 M= 2
s r s r
j1 − M + 12 2 j1 + M + 21 3
β3 = − =− ; β4 = = (1.84)
2j1 + 1 5 2j1 + 1 5
1 1
M= 2 M= 2
s r s r
j1 − M + 12 3 j1 + M + 21 2
β5 = − =− ; β6 = = (1.85)
2j1 + 1 5 2j1 + 1 5
1 1
M =− 2 M =− 2
s r s r
1 1
j1 + M + 2 4 j1 − M + 2 1
β7 = − =− ; β8 = = (1.86)
2j1 + 1 5 2j1 + 1 5
3 3
M =− 2 M =− 2
18
This algebra imples that the operator S ~ 2 = S 2 + S 2 + S 2 commutes with all the
x y z
three components of spin. Since different components of spin do not commute,
a commuting set of operators consist of S ~ 2 and component of the spin along any
one direction; most popular choice being S ~ 2 and Sz . Because the algebra of spin
operators is assumed to be the same as the angular momentum operators the results
on angular momentum apply to the spin also.We therefore have the following results:
~ 2 are given by s(s + 1)~2 , where s is a positive integer or
• The eigenvalues of S
half integer.
• A particle will be said to have spin s if the the maximum allowed value of Sz
is s~, which is same as S~ 2 having value s(s + 1)~2 .
19
In all there are (2s + 1) values of m ranging from −s to s and therefore (2s + 1)
eigenvectors |smi. The vector space needed to describe spin is linear span of all the
vectors |smi and is (2s + 1) dimensional.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
14001;
The interpretation of the numbers cm is that square of its modulus, |cm |2 , gives the
probability that Sz will have the corresponding value m~. Following the convention
of arranging the basis vectors in the order of decreasing values for the spin projection
m, the |χi will be represented by a (2s + 1) component column vector of the form
α1
α2
|χi = .. (2.7)
.
α(2s+1) .
The spin wave function of a particle with spin s is a column vector with (2s + 1)
components.
20
This follows from the fact that the basis vectors |smi are eigenvectors of Sz with
eigenvalues m~. (WHY?) The matrices for Sx and Sy are found by first obtaining
the matrices for S± and using Sx = 12 (S+ + S− ) and −i2
(S+ − S− ). To construct
these matrices one needs to know the matrix elements hs, m′ |S± |s, mi which can be
computed by making use of the result
p
S± |s, mi = s(s + 1) − m(m ± 1) ~ |s, m ± 1i (2.9)
We shall give the answer for spin 12 and spin 1 matrices. The spin half matrices
are related to the Pauli matrices σx , σy , σz and are given by
~ = ~ ~σ
S (2.10)
2
This result is derived in the solved problem below. The corresponding result for the
spin one matrices
is
0 1 0 0 −i 0 1 0 0
Sx = √~2 1 0 1 ; Sy = √~2 i 0 −i ; Sz = ~ 0 0 0 .
0 1 0 0 i 0 0 0 −1
qm-
and is left as en exercise for the reader.
ymp-
20001
⊲ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : .................................. ◭ QM-Lec-
14002;
§ 2.1 Problem.
Construct the spin matrices for a spin half particle.
1
Solution : For a spinparticle,
n 2let |o21 i and | − 21 i denote the eigenvectors of Sz with
eigenvalues ± 12 .In the basis | 12 i, | − 12 i the matrix for Sz is
~ 1 0
Sz = . (2.11)
2 0 −1
21
To summarize we get the result that the spin 21 matrices are given by ~
2 times the Pauli
matrices:
S~ = ~ ~σ . (2.19)
2
§3 Pauli Matrices
⊲
Pauli-
Matrices
We summarize some important properties of Pauli Matrices.
1. The three Pauli matrices are given by
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = . (2.20)
1 0 i 0 0 −1
There does not exist a nonzero 2 × 2 matrix which anticommutes with all the
three Pauli matrices.
22
23
Frequently, the total wave function factorises and assumes the form
Ψ(~r) = ψ(~r) × χ (2.38)
where χ is a column vector with (2s + 1) components, so for an electron we will have
α
χ= . (2.39)
β
In case the wave function can be written in the product form, we shall refer to ψ(~r)
as the space part of the wave function and the column vector χ as the spin part of
the wave function. The function ψ(~r) describes the translational degrees of freedom,
as usual, and the spin degrees of freedom are described by the column vector χ.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-20-01-
L3;
24
- Precise values for observable is given up, in general only probabilities can be
computed.
We shall see that for a system of two or more identical particles, the identity of
individual particles looses its meaning in quantum mechanics; one can only talk
about the system as a whole. We also need to introduce a new hypothesis about
the states of a quantum mechanical system of several identical particles.
Consider an electron electron scattering experiment. Classically it is always
possible to distinguish two electrons because they are point particles having well
defined positions. More importantly, this distinction can always be maintained
at all times because each particle has a well defined trajectories. By successive
measurements of its position at regular intervals very accurately the path of each
electron can, in principle, be followed without disturbing their motion. In fact
classical mechanics admits the possibility of predicting the position of a particle
at all times knowing initial conditions and interactions.classical mechanics admits
the possibility of distinguishing and maintaining the distinction between the two
electrons at all times.
For a quantum mechanical system consisting of several identical particles we note
the following.
• In the first place the particles cannot have well defined positions, only some
probabilities can be assigned to different values of the position.
• At any given time, it is certainly possible to localise each particle with great
accuracy. However, in general, the wave packets spread and it will not be
possible to maintain the localisation at later times. Thus when the particles
come very close we would not be able to tell which particle was which one.
25
for some real α, as required by the first postulate. The above equation is valid for
all ξ1 , ξ2 and hence, replacing ξ1 with ξ2 and ξ2 with ξ1 we get
26
• For a system of two identical particles with integral spin bosons the total wave
function must be symmetric under simultaneous exchange of all the variables
such as the space and spin variables. For a system of two identical particles
of half integral spin fermions the full wave function must be anti-symmetric
under a simultaneous exchange of all the variables such as the space and spin
variables.
If ξ1 , ξ2 denote the set of all variables such as, space and spin, of two identical
particles. Then the symmetrisation postulate states that the total wave func-
tion ψ(ξ1 , ξ2 ) must be symmetric for bosons and antisymmetric for fermions
under an exchange of ξ1 and ξ2 .
should hold for all pairs AK. Similarly, for a system of n− identical fermions
the total wave function must be anti-symmetric under simultaneous exchange
of variables ξj , ξk for every pair j, k
For a system of several identical bosons, the total wave function Ψ(ξ1 , · · · , ξn )
remains unchanged under an arbitrary permutation of ξ1 , · · · , ξn ; where as for
fermions the wave function remains unchanged under an even permutation but
changes sign under an odd permutations.
• The postulate is a statement about the full wave function of the system of
identical particles under a simultaneous exchange of all the variables. For
example, there is no constraint on the space part (or the spin part) of the wave
function alone need.
27
§7 Illustrative Examples
qm-
xmp-
20001
⊲
We shall now take up some examples of consequence of symmetrization postulate
for a system of two identical particles. We begin with two useful statements which
will be needed repeatedly in this connection.
⋆1) When we add two, equal, angular momenta j the possible resulting values are
J = 2j, 2j − 1, · · · , 0. Of these the state with the highest value, J = 2j, is
symmetric under an exchange of the two particles, the next one, with 2j − 1 is
antisymmetric; the states being alternately symmetric and antisymmetric as
J takes on the values in descending order.
⋆2) For a two particle system, the effect of an exchange of the positions of the
two particles is same as the parity on the wave function in the centre of mass
frame. Therefore under an exchange of the space variables the space part of
the wave function is symmetric for even ℓ and antisymmetric for odd ℓ.
§ 7.1 Example
Let us now consider two particles interacting via a spherically symmetric potential
v(r) where r is the distance between the particles. We write the total wave function
as
Ψtot (~r, ms1 , ms2 ) = ψspace (~r) ψspin (2.50)
If each particle has spin s the total spin has the values S = 2s, 2s − 1, 2s − 2, · · · 0
with symmetry properties as given by rule ⋆1). We shall see that for each value of
total spin S only odd, or even, angular momentum ℓ values will be permitted when
the particles are idemtical. We discuss four cases in the following.
28
b) When the two particles are identical fermions, 2s = odd integer, the total wave
function must be antisymmetric under an exchange of space and spin variables.
Therefore, antisymmetric space wave functions ( ℓ = odd) must be chosen with
symmetric spin wave functions, (S = 2s, 2s − 2, · · · ). Also the space part of wave
function must be symmetric, (ℓ = even ) when the spin part of the wave function is
antisymmetric (S = 2s − 1, 2s − 3, · · · ).
c) When the two particles are identical bosons, s = integer, the total wave function
must be symmetric. Hence either both space and spin parts must be symmetric or
both must be antisymmetric. Thus even ℓ values will correspond to total spin S =
2s, 2s−2, · · · and odd ℓ values will correspond to the total spin S = 2s−1, 2s−3, · · · .
d) In the special case of two identical spin zero bosons, the total wave function is just
the space part alone which must be symmetric and hence ℓ must be even. This give
the statement that the spin and parity of a system of two identical spin zero bosons
must both be even.
(−1)ℓ+2S = 1 (2.51)
or ℓ + 2S must be even.
§ 7.2 Problem.
Three identical spin zero bosons exist in states described by normalized wave func-
tions χ, φ, ψ write the total wave function for the system.
√
The factor (1/ 6) has been put to ensure correct normalization.
§ 7.3 Problem.
For a system of three identical fermions in states ψ1 , ψ2 , ψ3 write the total wave
function.
29
If the electrostatic interaction, e2 /|~r1 −~r2 |, between the two electrons is neglected
as a first approximation, the hamiltonian becomes a sum of two hydrogen atom like
hamiltonians. In this approximation the electronic states are described by quan-
tum numbers (n1 , l1 , m1 ) and (n2 , l2 , m2 ) for the two electrons. Let u1 , u2 denote
corresponding H-atom wave functions. The space part of the wave function for the
two electrons will be product wave function u1 (~r1 )u2 (~r2 ), which must be properly
symmetrized or anti-symmetrized as discussed below.
In very many situations the total wave function is a product of a part describing
space properties and a spin wave function. Thus we write
where m1 , m2 refer to the spin variables for the two electrons. As each electron
carries spin 1/2, the total spin can take values 1 (triplet) and 0 (singlet). The values
of total spin determines the symmetry property of spin wave function under an
exchange of spin variables. It is known that spin wave function must be symmetric
for S = 1 and antisymmetric for S = 0 states. The requirement that total wave
function be antisymmetric (for 2 electron systems) fixes the symmetry property of
the space part of the wave function as summarized in the table given below.
Spin State Total Spin Spin wave function Space wave function
30
1
ψ± (~r1 , ~r2 ) = √ (u1 (~r1 )u2 (~r2 ) ± u1 (~r2 )u2 (~r1 ))
2
the symmetric combination ψ+ should be used for the singlet states (S = 0) and the
antisymmetric combination ψ− should be used for triplet states (S = 1). The ground
state corresponds to n1 = n2 = 1 l1 = l2 = 0 m1 = m2 = 0 and the antisymmetric
combination ψ− vanishes. Only the symmetric combination is nonzero. Thus the
ground state is a singlet state; the same is true of all other states corresponding to
electrons having identical (n, l, m) quantum numbers.
When the two electron states correspond to different (n, l, m) quantum numbers,
both symmetric and antisymmetric combinations ψ± (r̄1 , r̄2 ) are possible. However,
the antisymmetric combination ψ± (~r1 , ~r2 ) vanishes when ~r1 = ~r2 . Therefore, the
probability that the two electrons will be found close to each other will be small
for ψ− ( for triplet states) and large for ψ+ (singlet states). Since the Coulomb
interaction between two electrons is positive and is large when their separation is
small, total Coulomb energy will be higher in singlet states as compared to its value
in the triplet state. These predictions are in accordance with the results on energy
levels of He atom derived from the spectrum of He atom.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
13006;
31
Variation Method
QM-
§1 Variation Method
21-
01- ⊲
L1 Theorem 1. Let E0 , E1 , E2 , · · · be the exact energy eigenvalues of Hamiltonian of
a system corresponding to the ground state, the first excited state, the second excited
state etc., respectively,
Huα = Eα uα (3.1)
where uα are eigenfunctions corresponding to eigenvalue Eα . Let ψ be any square
integrable function then we have the following results.
R ∗
ψ Hψd2 x
Eα ≤ R ∗ 3 (3.2)
ψ ψd x
and
Z !
X X
ψ ∗ Hψd3x = Cn u n , H Cn u m
n
XX
= Cn∗ Cm (un , Hum )
n m
XX
= Cn∗ Cm Em (un , Hum)
n m
XX
= Cn∗ Cm δmn Em
n m
X
= En |Cn |2 (3.5)
n=o
32
The right hand side is positive because En − E0 > 0, for all n 6= 0. Hence
Z Z
∴ ψ Hψd x − E0 ψ ∗ ψd3 x ≥ 0 .
∗ 3
(3.8)
or Z Z
∗ 3
ψ Hψd x ≥ E0 ψ ∗ ψd3 x . (3.9)
or R ∗
ψ Hψd3 x
R ≥ E0 . (3.10)
ψ ∗ ψd3 x
Thus we have R ∗
ψ Hψd3 x
E0 ≤ R ∗ 3 (3.11)
ψ ψd x
for all square integrable ψ.
• First we choose a trial wave function ψ which satisfies all general properties
of bound states in one dimension. First requirement is that the trial function
chosen must be square integrable. In addition the trial function must satisfy
usual continuity conditions. Thus, for the case of a potential which is finite
everywhere, the trial function and its derivative should be continuous. If the
potential has a infinite jump at a point, the trial wave function must be se-
lected to have a zero at that point. In accordance with the general results the
trial wave function for the ground state must not have any node. The trial
wave function must have even parity, in case the potential is an even function
of x.
R
• Normalise the trial function to unity ψ ∗ ψd3 x = 1.
• Compute Z
Eψ ≡ ψ ∗ Hψd3 x (3.12)
The trial wave function ψ will contain some unknown parameter(s). The
unknown parameters are then varied and one we demand that the average
33
∂Eψ
=0 (3.13)
∂α
The value α0 , for which Eψ is minimum, can now be used for calculating Eψ
which gives an estimate for the ground state energy. A good choice of trial
wave function can lead to very good estimate for the ground state energy. The
variation method has been successfully applied to the energy for ground state
of He atom. The variation method is most useful for the ground state energy
although it can be modified to estimate energies of excited states also.
• It can be easily seen that the variation method will give an upper estimate of
nth excited state, if an additional requirement that of the trial wave function
being orthogonal to the wave functions for all the states of lower energy levels
is imposed. So for example, for a potential which is an even function of x, use
of an odd function of x as a trial wave function, can be used to get an estimate
of first excited state energy. Due to this requirement thus the variation method
is not suitable for higher excited states.
34
H = H0 + H ′
in such a way that the eigenvalues and eigenfunctions of H0 can be found exactly.
We further assume that the additional part λH ′ has small in matrix elements as
compared to H0 .
The approximate solution is obtained by assuming that the exact eigenvalues
and eigenfunctions have an expansion in powers of λ and that it is sufficient to keep
first few terms of the expansion in powers of λ.
We shall be interested in first and second order perturbation theory (terms up
order λ2 ) for the following two cases when the eigenvalues of H0 are non-degenerate
or may have degeneracy. Notation: We shall write
H = H0 + λH ′ (4.1)
where λ is introduced for book keeping purposes and at the end of all computations λ
will be set equal to one. Eigenfunctions of H0 will be denoted by un and eigenvalues
by En
H0 un = En un (4.2)
Eigenfunctions of the full Hamiltonian will be denoted by ψ and the corresponding
eigenvalues by W .
Hψ = W ψ (4.3)
Further, it will be assumed that, ψ and W can be expanded in powers of λ
ψ = ψ0 + λψ1 + λ2 ψ2 + · · · (4.4)
W = W0 + λW1 + λ2 W2 + · · · (4.5)
35
H0 ψ0 = W0 ψ0 (4.9)
H0 ψ1 + H ′ ψ0 = W0 ψ1 + W1 ψ0 (4.10)
H0 ψ2 + H ′ ψ1 = W0 ψ2 + W1 ψ1 + W2 ψ0 (4.11)
(4.12)
These equations will be utilized and corrections to the energy levels and wave func-
tions for different cases, degenerate and nondegenerate levels, will be obtained in
the first and second orders in perturbation.
§2 Nondegenerate Case
We shall obtain corrections, (assumed to be small), to eigenvalue and eigenfunc-
tions corresponding to a fixed level n. Let un and En be the corresponding energy
eigenfunction and eigenvalue, so that
H0 un = En un (4.13)
where n is a given, fixed number throughout this discussion and refers to a particular
energy level of H0 .
In this section we assume that the energy level En is nondegenerate.
The equation Eq.(4.9) shows that ψ0 is an eigenfunction of H0 with eigenvalue
W0 . Since we are interested in finding corrections to energy and eigenfunctions of
the n-th energy level (assumed to nondegenerate), we take
W0 = En ψ0 = un (4.14)
H0 ψ1 + H ′ un = En ψ1 + W1 un (4.15)
H0 ψ2 + H1′ ψ1 = W0 ψ2 + W1 ψ1 + W2 un (4.16)
The overall normalisation of the full eigenfunction is arbitrary and can be always be
selected to be equal to unity at the end. It will turn out to be convenient to choose
the normalisation of ψ in the intermediate steps to be
(un , ψ) = 1 (4.17)
36
The above equation holds for all k. We shall consider the two cases k = n and k 6= n
separately.
§ 2.1.1 Case k = n
For k = n Eq.(4.22) takes the form
The first terms on the two sides cancel, noting (un , un ) = kun k2 = 1, because un are
assumed to be normalized, we get from Eq.(4.23)
W1 = (un , H ′ un ) (4.24)
This equation gives then first order correction to the energy eigenvalue En . Thus
the energy eigenvalue upto first order becomes
W = En + (un , H ′ un ) (4.25)
§ 2.1.2 Case k 6= n
In this case (uk , un ) = 0 and the last term in Eq.(4.22) in the r.h.s. drops out and
we get
Ek (uk , ψ1 ) + (uk , H ′ un ) = En (uk , ψ1 ) (4.26)
which on a rearrangement gives
37
∞
X ′ < k|H ′ |n >
ψ1 = uk (4.34)
En − Ek
k=1
P′
where means sum over all values of k except k = n. Note that in the first order
correction, Eq.(31), the constant cn is not determined. Upto first order in λ the
wave function becomes
ψ ≈ ψ0 + λψ1 (4.35)
X∞
< k|H ′ |n >
≈ λ uk (4.36)
k=1
E n − Ek
k6=n
H0 ψ2 + H ′ ψ1 = W0 ψ2 + W1 ψ1 + W2 ψ0 (4.37)
38
Now we take scalar product of Eq.(4.37) with up and consider cases p = n and p 6= n
separately. Eq.(4.37) gives
(up , H0 ψ2 ) + (up , H ′ ψ1 ) = W0 (up , ψ2 ) + W1 (up , ψ1 ) + W2 (up , ψ0 ) (4.42)
Recall W0 = En , ψ0 = un and rewrite the first term as (up , H0 ψ1 ) = (H0 up , ψ1 ) =
Ep (up , ψ1 ). In the first term on the right hand side of (36) use W0 = En , ψ0 = un
and rewrite (36) as
Ep (up , ψ2 ) +(up , H ′ ψ1 ) = En (up , ψ2 ) +W1 (up , ψ1 ) + W2 (up , un ) (4.43)
| {z } | {z }
(1) (2)
where the cn terms drop out when Eq.(4.40) is used. The first term, marked (1), on
the right hand side is zero because of orthogonality (un , uk ) = 0 when k 6= n. Thus
we get
X < k|H ′|n >
W2 = < n|H ′|k > (4.47)
k6=n
E n − Ek
39
§ 2.2.2 Case p 6= n
We now determine the second order corrections to the wave function. Eq.(4.43) now
takes the form
The last term (III) is zero because (up , un ) = 0 for p 6= n. Combining the terms
marked, (I) and (II) and using (up , ψ2 ) = 0 in Eq.(4.49) gives
where
(un , H ′ un )(up , ψ1 ) (up , H ′ψ1 )
dp = − (4.54)
Ep − En Ep − En
The first order correction ψ1 , appearing in the above equations is known and is given
by Eq.(4.41).
X′ hk|H ′|ni
ψ1 = uk (4.55)
k
E n − Ek
Also, applying the operator H ′ on Eq.(4.55) and taking scalar product with up gives
∞
X ′ hk|H ′ |ni
′
(up , H ψ1 ) = (up , H ′uk ) (4.57)
k=1
E n − Ek
40
41
and the coefficients dn , n 6= k are just the constants given by dk = (uk , ψ1 ) and hence
X (uk , H ′ψ0 )
ψ1 = d(1)
n un
(1)
+ d(2)
n un
(2)
+ uk (4.81)
k6=n
Ek − En
42
ψ = ψ0 + λψ1 + λ2 ψ2 (4.82)
= α1 u(1) (2) 2
n + α2 un + λψ1 + λ ψ2 (4.83)
Without loss of generality, we may assume that ψ1 and ψ2 are orthogonal to the
(1) (2)
unperturbed eigenfunctions un , un . Demanding this amounts to changing, as yet
unknown, parameters α1 , α2 . We start with the equation, Eq.(4.65) ,
H0 ψ2 + H ′ ψ1 = W0 ψ2 + W1 ψ1 + W2 ψ0 (4.84)
(1)
Taking scalar product with un , we get
The first two terms on the right hand side are zero due the orthogonality of ψ1 , ψ2
(1)
to un . The first term on the left hand side also vanishes on using the hermiticity
of H0
(u(1) (1)
n , H0 ψ2 ) = (H0 un , ψ2 ) (4.86)
= En (u(1)
n , ψ2 ) (4.87)
= 0
(u(1) ′ (1)
n , H ψ1 ) = W2 (un , ψ0 ) (4.88)
= α1 W2 (4.89)
X hk|H ′|ψ0 i
= W2 α1 (4.90)
k6=n
Ek − En
43
The two eigenvalues, for W2 , give the second order corrections and the eigenvectors
give the values of α1 , α2 which determines the lowest order wave function ψ0 . It
must be remembered that the above formula apply only in case the degeneracy is
not removed to the first order in λ.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-23-02-
L3;
α2 mc2
En = − , (4.94)
2n2
e 2
where α = 4πǫ 0
in si units, and is known as the fine structure constant. It is
dimensionless and has a value ≈ 1/137.07. The Schrödinger theory predicts that
for a fixed value of the principal quantum number n, the angular momentum can
take values ℓ = 0, 1, · · · , (n − 1) and for each ℓ the z component of the angular
momentum has (2ℓ + 1) values mℓ = −ℓ, −ℓ + 1, · · · , ℓ. For each level corresponding
to these combinations of nℓ, mℓ quantum numbers, the component of the spin of the
electron along a specified directions can have two possible values to ± ~2 . The energy
depends only on n and hence there is a degeneracy of 2n2 . For a few energy levels
we schemetically show the set of quantum numbers in figure below.
44
Each line shown here in the diagram must be regarded as a bunch of 2n2 lines
nℓ and ms quantum numbers explicitly indicated in the Fig.1.(b) of the diagram. In
Fig.1.(c) we show the labels written in terms of n, ℓ, j, mj quantum numbers, where
j refers to the total angular momentum. For each ℓ there are two values j = ℓ ± 12 ,
and again mj takes (2j + 1) values from −j, −j + 1, · · · , j.
45
3s 1 , 3p 1 , 3p 3 , 3d 3 , 3d 5
2 2 2 2 2
are all degenerate. The relativistic effects split the level into three groups as
given below.
The results described above are shown in figure, not to the scale, below. As already
noted all the above splittings are small and are of the order of α4 mc2 .
Lamb shift
The Lamb shift refers to a very tiny splitting of energy levels with same nj and
different ℓ values. The Lamb shift is of the order of α5 mc2 . For example, experi-
mentally the levels 2s 1 and 2p 1 are not degenerate and the energy difference is given
2 2
by
Hyperfine structure
The energy levels of Hydrogen atom after taking into account of of the fine structure
and the Lamb shift appear as shown in figure below. Each of these levels have a
hyperfine structure; each level splits further into two levels with energy differences
m 4
∼M α mc2 where M is the proton mass.
Order of magnitudes of various observed energy level splittings can be sum-
marised as follows.
me
α2 mc2 α4 mc2 α5 mc2 mp
α4 mc2
∼ a few eV ∼ 10−4 eV ∼ 10−6 eV ∼ 10−7 eV
46
3. Darwin term
π~2
H3 = (Ze2 )δ(~r).
2m2 c2
47
48
Computing the averages of 1/r and 1/r 2 gives the first order correction (∆E)1 .
~2
hnℓjm|H3 |nℓjmi = (j(j + 1) − ℓ(ℓ + 1) − 1/4)hnℓjm|ξ(r)|nℓjmi, (4.106)
2
where
1 Ze2
ξ(r) = , (4.107)
2m2 c2 r 3
and V (r) = Ze/r for the Coulomb potential due to the nucleus has been substituted.
For spin half case in hand, j can take two values ℓ ± 21 and we get
(
~2 ℓ j = ℓ + 21
hnℓjm|H3 |nℓjmi = hξ(r)i (4.108)
2 −ℓ − 1 j = ℓ − 12 .
49
This completes calculation of the shift in energy due to spin ornit term, (∆E)2 .
Remarks:
The fine structure was explained within the Schrödinger non-relativistic quantum
mechanics before Dirac equation was proposed. The presence of relativistic variation
of mass and the spin orbit interaction, but not the Darwin term, can be understood
without recourse to the Dirac equation.
The relativistic energy of a particle of mass m is given by
p p
E = p2 c2 + m2 c4 = mc2 1 + (p/mc)2 (4.112)
p2 p4
E = mc2 + − +··· (4.113)
2m 8m3 c2
The first term is the rest energy, the second term is just the kinetic energy of a point
mass in non-relativistic mechanics. The third term is lowest order correction due to
relativistic variation of mass.
The spin orbit coupling term represents the interaction of spin magnetic moment,
ge ~
~µ = − 2mc S, of the electron with the magnetic field due to its orbital motion and
where g is the gyromagnetic ratio for the electron and has the g ≈ 2. In the rest
frame of the nucleus there is Coulomb potential V (r) gives an electric field. This
electric field
~ = −∇V (r) = − dV (r) ~r
E
dr r
50
B ~ = 1E
~ = − 1 ~v × E ~ × ~v = − 1 dV (r) (~r × ~v ). (4.114)
c c cr dr
The interaction of the magnetic dipole moment of the electron due to its spin,
e ~ ~ which has the form
~µ = − mc S leads to an interaction −~µ · B
~ = e (B
−~µ · B ~ · S)
~ (4.115)
mc
e dV (r)
= − 2 ~r × ~v · S~ (4.116)
mc r dr
e dV
= − 2 2 ~r × p~ · S ~ (4.117)
m c r dr
e 1 dV ~ ~
= − 2 2 L·S . (4.118)
m c r dr
where L~ is the orbital angular momentum of the electron. a naive application of
Lorentz transformation on the electric field to obtain the magnetic field in the rest
frame is not justified as the electron’s velocity keeps changing and hence the rest
frame of the electron is not an inertial frame. Thus the spin orbit term obtained
above requires a correction, known as Thomas precision correction, which is of the
same form as the electron spin magnetic moment interaction with the magnetic field
but requires a numerical factor of 1/2 in the final expression. This gives H2′ as the
final answer for the spin orbit term. There is no way of seeing the appearance of
Darwin term, which was first derived from the Dirac equation.
Hyperfine structure
The hyperfine structure comes from the interaction of electron’s dipole moment with
the magnetic field due to the spin dipole moment of the nucleus. The proton, like
ge ~
every charged spin half particle, carries a magnetic moment given by µ~p = 2M pc
S,
g, called the gyromagnetic ratio, is found to be 5.59. The magnetic moment for the
proton is very small compared to the magnetic moment of the electron because the
mass Mp appears in the denominator. The hyperfine structure splitting works out
to be proportional to the average value of S~e · S~p . This operator has diagonal matrix
elements between states with total spin S ~ = S~e + S~p . The total spin for the electron
proton system can be 0 (singlet) or 1 (triplet) and we write
1 2
S~e · S~p = S − Se2 − Sp2 (4.119)
2
(
1
4
for triplet, S = 1
= 3
(4.120)
− 4 for singlet, S = 0
Thus the effect of hyperfine interactions is to split each level into two levels corre-
sponding to total spin S = 0, 1. Assembling all the factors, this splitting works out
to be very tiny, of the order of 10−6 eV. For the ground state of hydrogen atom, the
photon emitted in transition from triplet to singlet has the wave length of 21cm,
and is the famous 21 cm line is present everywhere in radiation in the universe.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
16006;
51
52
[2] Use the first order non-degenerate perturbation theory to compute the correc-
tion to the ℓ = 1 level of ( use units so that ~ = 1 )
H = L2 + αLz + βLx
(a) Find the corrections to the ℓ = 1 energy level using perturbation theory
to the lowest non-vanishing order in H ′ . Compare your answer with the
exact answers.
(b) Obtain the eigenvectors of H upto lowest order in β.
Tutorial-II
[1] A particle moves in two dimensional circular oscillator potential with
p2x p2y 1
H0 = + + mω 2 (x2 + y 2)
2m 2m 2
(a) What are the quantum numbers of the first excited state? Is it degenerate
or not?
(b) If a small perturbation H ′ = λxy is applied compute the lowest order
correction to the energy of the first excited state.
[2] Use the first order degenerate perturbation theory to compute the correction
to the l = 1 level of ( use units such that ~ = 1)
H = L2 + αLx + βLy
53
[3] Following the steps given in the class, find eigenvectors and eigenvalues of the
matrix
1 ǫ 0
ǫ 1 −ǫ
0 −ǫ 2
upto the lowest non-vanishing order in ǫ.
Questions
[1] Show that the second order perturbation correction to the ground state energy
is always negative.
[2] Write the steps you would follow to find the first order correction to the energy
levels in perturbation theory when the unperturbed energy eigenvalues are
degenerate.
b commutes with X,
[3] If a hermitian operator H b and |ni is an eigenvector of H
b
with eigenvalue En . Prove that
b
hn|X|mi = 0.
if the eigenvalues Em and En are different. What is the use of this relation in
perturbation theory ? Is it necessary that Hb be a hermitian operator ?
b is replaced with the parity operator in the previous question, what does
[4] If H
one get? Interpret your answer.
[5] Give two examples of operators X, Y , such that their commutator is propor-
tional to one of them. Where does this property appear in quantum mechanics
and how is it used ?
b Yb be two operators such that X is hermitian, and their commutator is
[6] Let X,
b Yb ] = αYb
[X,
b with eigenvalue x0
If |x0 i is an eignevector of X
b 0 i = x0 |x0 i,
X|x
[7] Why is that the first order perturbation theory, for the ground state energy,
will never give a better estimate as compared to the value that can be obtained
by using the variation method ?
54
[9] When is the effect of finite size of nucleus to the electronic energy levels most
important for alkali atoms ? Assume hydrogen like levels and answer in terms
of the quantum numbers n, ℓ etc .
[10] A singlet excited state of He atom always has a higher energy as compared to
the corresponding triplet state. Explain why ?
55
§1 Introduction
QM-
Lecs-
17001
⊲
The Wentzel Kramers Brillouin or the WKB approximation scheme is a method of
solving the Schrodinger equation by means of an expansion in powers of ~. The
leading term, in the limit ~ → 0 makes contact with the classical mechanics, there-
fore this method is also known as semiclassical approximation. This scheme was
developed for solving differential equations long before it was applied to quantum
mechanical problmes. For bound state problems it connects the Bohr Sommerfeld
quantisation rule with the Schrodinger theory. The Bohr Sommerfeld scheme of old
quatum theory worked for bound state problems but failed for alpha decay. The
WKB method is historically important because it was successfully used by Gamow
(1928) and by Gurney and Condon (1929) to give first explanation of alpha decay
of nuclei.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lecs-
17001;
§2 Semiclassical expansion
QM-
Lec-
17002
⊲
We are interested in solving the Schrödinger equation in one dimension
~2 d2 ψ(x)
+ V (x)ψ(x) = Eψ(x)
− (5.1)
2m dx2
by writing an ~ expansion for S related to the wave function by
ψ(x) = exp(iS/~). (5.2)
The equation satisfied by the function S is seen to be
2
1 dS i~ d2 S
− = E − V (x) (5.3)
2m dx 2m dx2
or 2 2
dS dS
− i~ = 2m(E − V (x)). (5.4)
dx dx2
We expand S in powers of ~
S = S0 + (~/i)S1 + (~/i)2 S2 + · · · (5.5)
56
S = S0 − i~S1 + · · · , (5.15)
Therefore, the most general solution for the wave function in this approximation is
a linear combination
c1 i Z c2 iZ
ψ(x) = p exp p(x)dx + p exp − p(x)dx , (5.18)
p(x) ~ p(x) ~
57
58
dp dV (x) 1
= −m p (5.24)
dx dx 2m(E − V (x))
mF
= (5.25)
p
where F = −dV (x)/dx is the force. Therefore, the condition Eq.(5.21) for validity
of the WKB approximation can also be cast in the form
m~|F |
≪ 1. (5.26)
p3
The above condition implies that the WKB approximation is not valid at those
points where the force is large or the momentum is small. Since near a truning
point p ≈ 0, Eq.(5.26) leads us back to the conclusion that the WKB approximation
breaks down near turning points.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
17003;
59
60
with a similar form of solutions for the WKB wave functions in region R4 near the
turning point x2 . One then demands that the Airy function solution in region R2
, ψ2 (x), matches with WKB solution ψ1 (x) for x < x1 and with ψ3 (x) for x > x1 .
This helps us in fixing the constants a3 , b3 in terms of a1 , an exercise that will be
done in the next section. We call the resulting solution ψ3 (x), for region R3 , as χ(x)
and write it as
Z
A 1 x π
χ(x) ≡ p cos p(x) dx − , x ∈ R3 . (5.34)
p(x) ~ x1 4
Thus we have arrived at two answers, χ(x) and φ(x), for the wave function in region
R3 , and both represent the same bound state wave function in this region. Obviously,
these two solutions themselves must agree for all x in region R3 implying that
Z Z
1 x π 1 x2 π
cos p(x) dx − = ± cos p(x) dx − (5.36)
~ x1 4 ~ x 4
for all x. This equation can be satisfied for all x only if the sum, or the difference,
of arguments of the two cosine functions is independent of x and equal to a multiple
of π. Note that the argument of the cosine on the left is an increasing function and
61
giving Z x2
1
p(x) dx = (n + 1/2)π, (5.38)
~ x1
which we rewrite as Z x2
2 p(x) dx = (n + 1/2)h. (5.39)
x1
Note carefully that the right hand side of the above equation has Planck’s constant h
and not ~ = h/2π. Remembering that p(x) is the H classical momentum, the integral
in Eq.(5.39) is just the action integral J(E) = p(x) dx taken over one period, i.e.
around the classical orbit of the particle from x1 to x2 and then back to the point
x1 and the quantisation condition takes the form
This condition, known as WKB quantisation rule for bound states, determines the
bound state energies without solving the Schrodinger equation and differs from the
Bohr Sommerfeld quantisation rule by presence of extra term h/2 in the right hand
side.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
17004;
~2 d 2 ψ
+ − V (x0 ) − F (x − x0 ) − E ψ=0 (5.42)
2m dx2
dV
can be solved exactly. If x0 is a turning point, F = − dx is the force at the turning
x0
point x0 and V (x0 ) = E, Eq.(5.42) takes the form
~2 d 2 ψ
+ F (x − x0 )ψ = 0. (5.43)
2m dx2
62
63
The answer Eq.(5.55) obtained for a linear potential, expressed in terms of the
original variable x, denoted as ψ2 takes the form
h √ i
√ 2α 2 2mF 3/2
4
exp − 3 ~
(x0 − x) , x < x0
2mF (x0 −x) h i
ψ2 (x) ≈ √ . (5.56)
√4 2α cos − 2 2mF
(x − x ) 3/2
− π
x > x
2mF (x−x ) 3 ~ 0 4 0
0
a1 h 2 √2mF i
3/2
ψ1 (x) = p exp − (x0 − x) , x < x0 (5.61)
4
2mF (x0 − x) 3 ~
64
Connection rules The results derived above can be summarised as follows. For
the case of an infinite barrier on the left of the turning at x0 , the decaying solution
on the left of turning point is connected to an oscillatory solution on the right by
the rule
1 Z x0 Z x
1 2 π
√ exp − |p(x)|dx → √ cos p(x)dx − . (5.67)
|p(x)| ~ x p x0 4
In the case of an impenetrable barrier to the right of the turning point, the connec-
tion rule becomes
Z x0 1Z x
2 π 1
√ cos p(x)dx − ←√ exp − |p(x)|dx . (5.68)
p x 4 |p(x)| ~ x0
These rules can be used only in the indicated direction. In order that the use of
connection rules may be justified it is important that the two regions, in which WKB
solutions are valid and in which linear approximation for potential holds, must be
overlapping regions. Therefore, rules will fail for a potential which varies fast.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lec-
17005;
65
Fig. 5.3.
qm-cdr-wkb03;Tunneling Through a Barrier
In Fig. 5.3 the regions R1 and R5 denote the classical regions and R3 is nonclas-
sical region, all away from the turning points a, b, where the WKB approximation
holds. These are separated by regions R2 , R4 around the turning points where the
WKB approximation breaks down and linear approximation to the potential has
been used to derive the connection rules.
We assume the barrier to be very wide and almost impenetrable and take the
wave functions in regions R1 and R3 to be given by the WKB result:
R
√2A 1 x
cos ~ a
p(x) dx , x ∈ R1
p(x) h i
A
R x
ψ(x) = √|p(x)| exp − ~1 a |p(x)| dx , x ∈ R3 (5.69)
h R i
√C exp ~i bx p(x) dx ,
x ∈ R5 .
p(x)
For x to the right of the barrier in R5 , only transmitted wave must be present and
the form of the WKB solution above in region R3 has been chosen so as to represent
waves travelling to the right in R5 .
The above solution (5.69) in region R3 will be rewritten in the form
A h 1Z x i
ψ(x) = p exp − |p(x)|dx
|p(x)| ~ a
h Z Z i
A 1 b 1 b
= p exp − |p(x)| dx + |p(x)| dx x
|p(x)| ~ a ~ x
D h1 Z b i
= p exp |p(x)| dx ∈ R3
|p(x)| ~ x
. (5.70)
66
It must be noticed that the connection rules, derived previously, are not applicable
to relate the coefficients C and D because the form of the above solution has a form
in region R3 and R5 different from that appearing in the connection formulae. A
connection between the above two forms for the wave function in regions R3 and R5
can be established by making use of the result that the Wronskian of two solutions
of the Schrodinger equation is independent of x.
dφ2 (x) dφ1 (x)
W (x) = φ1 (x) − φ2 (x) = constant. (5.73)
dx dx
To use this we take the wave function in Eq.(5.72) as φ1 . The WKB solution
appearing in the connection rule Eq.(5.67) , with classical region to the right of the
turning point, is taken as φ2 (x):
√ 1 exp − 1 R b |p(x)|dx x ∈ R3
|p(x)| ~ x
φ2 (x) = R
2 cos x p(x)dx − π x∈R
b 4 5
We compute the Wronskian W (x) to the left and to the right of x = x1 and equate
the two values at x = x2 . Using the above results the transmission coefficient can
now be worked out and is found to be
2 2Z b
C
T = = exp − |p(x)|dx . (5.74)
A ~ a
where Yℓm (θφ) are spherical harmonics and R(r) is the radial wave function obeying
the radial equation
~2 1 d 2 d ℓ(ℓ + 1)~2
− r R(r) + V (r) + R(r) = ER(r) (5.76)
2m r 2 dr dr 2mr 2
67
r = exp(s) (5.79)
φ(s) = 0 (5.82)
as s → ± ∞. Thus we can take over the WKB result in one dimension and the
quantization condition can now be written as
Z s2 1/2
2s 2 1 2
2 2m(E − V (s))e − ~ (ℓ + 2 ) ds = (n + 12 )h (5.83)
s1
It should be noted that the centrifugal term appears with coefficient (ℓ + 12 )2 instead
of the usual ℓ(ℓ + 1). The same result can also be derived by starting from Eq.(5.77)
and applying the boundary condition Eq.(5.78).
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-22-03-
L1;
68
[2] Show that the WKB approximation gives exact answer for energy eigenvalues
of a harmonic oscillator.
[3] Using WKB approximation find the energy levels of a particle in a gravitational
potential of the earth. You may assume the acceleration due to gravity to
remain constant and that the surface of the earth is a rigid boundary.
[2] Obtain the energy quantization condition using WKB method for a particle
in one dimensional well with potential given by
V0 0 < x < a
V (x) = 0 a < x < a + b
∞ x > a or x > a + b
69
[Mavromatis]
[5] Find the energy levels of a particle in a potential well
V (x) = A|x|1/2
using the WKB approxmation.
15~A2 25
En = n2/5 √
8 2m
[6] Find the energy levels of a particle moving in a one dimensional potential
1
V (x) = −A|x|− 2
using WKB approximation and argue that an infinte number of levels are
possible
2 4~|A|−2 − 3
2
−
En = −n 3 √
2m
[7] Use WKB quantization condition to argue that the number of levels for a
potential well
V0
V (x) = − 4
(x + a4 )
cannot be infinte.
70
71
72
Scattering Theory
§1 Introduction
In this unit description of scattering and scheme of computation of cross section in
quantum mechanics is introduced. This is achieved by imposing a suitable bound-
ary condition on the solution of the time independent Schrödinger equation and
converting Schrödinger equation into an integral equation using the Green function
for the free particle Schrödinger equation. A perturbative solution of the integral
equation leads to the Born approximation for the scattering amplitude.
73
The most general solution will be a superpostion of the above special solutions. The
free particle behaviour will hold for the scattering solutions for a potential which
goes to zero for large distances, giving infinite number of solutions. Thus specifying
the energy alone is not sufficient to pick a unique solution, it is necessary to specify
a boundary condition suited to the scattering problem.
74
No of particle detected per sec = σ(θ) × ∆Ω × Flux of the incident beam. (6.6)
knowing the wave function, the number of particles detected per second can be
computed using the probability current density ~j. The opening of the detector is
kept perpendicular to the radius vector and usually covers only a small solid angle,
the probability of a particle entering the detector per sec is given by the surface
integral x
jr dS ≈ jr ∆S (6.7)
S
over the surface of the detector, where jr is the radial component of the probability
current. The number of particles detected per sec will be N times the expression
(6.7). Thus Eq.(6.6) becomes
N × jr ∆S = σ(θ) × ∆Ω × Flux of the incident beam = σ(θ)∆Ω × Njz . (6.8)
where jz is the z component of the probability current. Using
~j = i~ ψ ∗ ∇ψ − ψ∇∗ ψ . (6.9)
2µ
75
~k
N(|f (θ)|2 /r 2 )∆S = σ(θ) × ∆Ω × N ,. (6.12)
µ
We have ignored the terms in the current coming from the interference of the in-
cident and scattered waves. These are of the order of 1/r 2, and are proportional
to exp(ikr(1 − cos θ)). Due to the presence of large r in the exponential, this term
oscillates rapidly with θ and sums out to vanishingly small value over range of θ
values.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lecs-
18001;
§3 Integral Equation
QM- A simple example
18-
01- ⊲
L1 Very often the problem of solving a linear differential equation can be replaced with
solution of an integral equation. An important feature of the integral equation
approach is that the initial conditions to be satisfied by the solution is built into the
integral equation. The integral equation can be solved many a times by an iterative
procedure which we shall illustrate by a simple example.
Suppose we are interested in solving the differential equation
dy
= λy (6.14)
dx
subject to the boundary condition
y(x)|x=x0 = N (6.15)
76
The constant in the above equation is fixed by making use of the initial condition
Eq.(6.15), and we get const= N and
Z x
y(x) = N + λ y(t)dt (6.18)
0
It is to noted that the unknown function y appears inside the integral sign, hence
an equation of this type is called integral equation. This equation can be solved
iteratively giving a solution as a series in powers of λ. This method gives the exact
answer in the case of this simple example under consideration. As a first step, we
set y equal to y0 where
y0 = N (6.19)
is the solution in the zeroth order in λ and is simply taken to be equal to the first
term Eq.(6.18). Next the zeroth order ’solution’ ,y0 , is substituted in the right hand
side of Eq.(6.18) to get the solution in the first order in λ. Thus we have
Z x
y1 (x) = N + λ y0 dx (6.20)
Z0 x
= N +λ Ndx (6.21)
0
77
y(x) = α0 + λα + λ2 α2 + · · · + αn λn + · · · (6.33)
we substitute Eq.(6.33) in Eq.(6.32) and compare powers on both the sides. There-
fore we get
Z x
2 n
α0 (x)+λα1 (x)+λ α2 (x)+. . . λ αn (x)+· · · = N+λ α0 (t) + λα1 (t) + λ2 (t)α2 (t) + · · · dt
0
(6.34)
on comparing coefficients of different powers of λ we successively get
α0 (x) = N
Z x
α1 (x) = α0 (t)dt = Nx
0
Z x
x2
α2 (x) = α1 (t)dt = N
2
Z0 x
x3
α1 (x) = α2 (t)dt = N
3!
Z0 x
xk
αk (x) = αk−1 (t)dt = N (6.35)
0 k!
Therefore we get
λ2 2 λn n
y(x) = N 1 + λx + x + · · · + x + · · · (6.36)
2! n!
78
∇2 φ0 (~r) = 0, (6.39)
then Z
ρ(~r′ ) 3 ′
Φ(~r) = φ0 (~r) + G(~r − ~r′ ) dr (6.40)
ε0
is a solution of the Poisson equation which can be easily verified by applying ∇2 on
both sides of Eq.(6.40).
Z
2 2 2 ρ(~r′ ) 3 ′
∇ Φ(~r) = ∇ φ0 (~r) + ∇ G(~r − ~r′ ) dr (6.41)
ε0
Z
ρ(~r′ ) 3 ′
= ∇2 G(~r − ~r′ ) dr (6.42)
ε0
Z
1
= − δ 3 (~r − ~r′ )ρ(~r′ )d3~r′ (6.43)
ε0
ρ(~r)
= − (6.44)
ε0
It can be shown that one solution of Eq.(6.38) is
1
G(~r) = . (6.45)
4πr
This Green function gives the potential due to a charge distribution subject to
the condition that the potential vanishes at infinity. The exact form of the Green
function and the solution φ0 of the Laplace equation is determined by the boundary
conditions of the problem.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-18-01-
L2;
QM-
§4 Integral Equation for Scattering
18-
02- ⊲
L1 In order to convert the Schrodinger equation
2
−~ 2
∇ + V (r) ψ = Eψ (6.46)
2µ
into an integral equation we first rewrite it as
∇2 + k 2 ψ = U(~r)ψ(~r) (6.47)
2µE 2µ
where k 2 = 2 , U(r) = 2 V (r) and we have defined Green function G(~r) as a
~ ~
solution of
∇2 + k 2 G(~r) = −δ(~r) (6.48)
79
To verify that ψ(~r) given by Eq.(6.55) does indeed have correct asymptotic property
~r
we expand |~r − r~′ | in powers of we shall assume that the potential is short range
~
r′
potential so that the contribution to integral over r~′ comes from small value of r ′ .
Expand |~r − r~′ | in powers of r ′
√
|~r − r~′| = r 2 + r ′2 − 2~r · ~r′ (6.56)
1/2
~r · ~r′ r ′2
= r 1−2 2 + 2 (6.57)
r r
80
We shall next discuss how to use Eq.(6.55) and Eq.(6.66) to obtain scattering
amplitude in the Born approximation.
QM-
§5 Born Approximation
18-
02- ⊲
L2 We have seen that the energy eigenfunctions, with a correct asymptotic behaviour
corresponding to the scattering solutions satisfy the following integral equation
Z
~ 1 exp{ik|~r − ~r ′ |}
ψ(~r) = exp(iki .~r) − U(~r ′ )ψ(~r ′ )d3 r′ (6.67)
4π |~r − ~r ′ |
where we use the notation
81
This process can be continued indefinitely and it becomes very cumbersome to com-
pute the wave function beyond first few orders.
The first order Born approximation consists in using the first, the plane wave,
term in the above series as approximate wave function in the expression, Eq.(6.66) ,
for the scattering amplitude
Z
1
f (θ, φ) = − exp(−i~kf · ~r′ )U(r ′ )ψ(r ′ )d3 r ′ (6.73)
4π
µ Z
= − exp(−i~kf · ~r′ )V (r ′ )ψ(r ′ )d3 r ′ (6.74)
2π~2
giving Z
µ ~ ~
f (θ, φ) ≈ − ei(ki −kf )·~r V (r)d3r (6.75)
2π~2
or Z
µ
f (θ, φ) ≈ − e−i~q·~r V (r)d3 r (6.76)
2π~2
where ~q = ~ki − ~kf is the momentum transfer. The result Eq.(6.76) is the well known,
first order, Born approximation result for the scattering amplitude.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-18-02-
L2;
82
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-18-02-
L3;
In the next subsection we shall derive, the condition for validity of the Born
approximation for the square well potential.
83
Low energy scattering At low energy the de Broglie wave length is much larger
than the range of the potential i.e. 2kR0 << 1. We then have
ρ2 − 2ρ sin ρ + 2 cos ρ
2 ρ3 ρ2 ρ4
≈ ρ − 2ρ ρ − + ···) − 2 1 − + +··· +2
6 2 24
ρ4
= (6.89)
4
Hence at low energies the Born approximation is applicable if
High energy limit In the high energy limit ρ >> 1 and we get
12
ρ2 − 2ρ sin ρ − 2 cos ρ + 2 ≈ρ (6.91)
84
V (~r) = gδ 3 (~r)
[2] A nucleon is scattered from a heavy nucleus. The effect of the nucleus can be
represented by a square well potential
−V0 r < R
V (r) =
0 r>R
Compute the differential cross section in the Born approximation.
[3] Classically the total cross section for scattering from a hard sphere of radius
R is πR2 . What happens if you try to compute the quantum mechanical cross
section using Born approximation? Discuss and give your comments.
[4] Compute the differential cross section in first Born approximation for the gaus-
sian potential
V (r) = −V0 exp(−r 2 /a2 )
and show that the total cross section is given by
! !
σt = V02 π 2 µ2 a4 /2~4 k 2 1 − exp(−2a2 k 2 )
Verify that the total cross section has correct dimension of area.
[5] For 2 MeV neutrons hitting protons at rest, find out if the Born approximation
is a good approximation. Consider nuclear forces to be described by a square
well of depth V0 = −35MeV and range R = 2.5 fm.
Cross Links
[1] Verify that the Green function
m
GE (x) = −i exp(ik|x|)
~2 k
p
where k = 2mE/~2 satisfies the one dimensional equation
2 2
~ d
+ E GE (x) = δ(x) (6.94)
2m dx2
(a) Set up the integral equation for reflection and transmission coefficients
for the one dimensional problems.
85
[3] (a) Use the Green function as in Q[1] and find an approximate expression for
reflection coefficient for an arbitrary potential V (x).
(b) Show that for a square well potential
−V0 for |x| < a2
V (x) =
0 otherwise
m2 V02
R≈ 4 4
sin2 ka
~k
86
§1 Introduction
We shall now take up the metho of partial waves for calculating cross section. This
method, applicable to the spherically symmetric potentials, tyrns out to be partic-
ularly useful for short range potentials at low energies. We shall, therefore, restrict
our attention to the spherically symmetric potentials V (r) which for large r go to
zero faster than 1/r 2 , i.e., r 2 V (r) → 0.
QM-
§2 Partial Wave Expansion of Plane Waves
19-
01- ⊲
L1 The free particle Schrodinger equation in three dimensions
~2 2
− ∇ ψ = Eψ (7.1)
2µ
can be solved by separating the variables in the cartesian as well as the spherical
polar coordinates. Writing the above equations as
∇2 ψ + k 2 ψ = 0, (7.2)
where
2µE
k2 = , (7.3)
~2
solution of the free particle Schrodinger equation in the cartesian coordinates is
given by
ψ~k (~r) = exp(i~k · ~r), ~k = (k1 , k2 , k3) (7.4)
Using ~n to denote the direction of ~k to write ~k = kn̂, we see that for each energy
value E > 0 the degeneracy is infinite; there being one solution for each n̂, corre-
sponding to the direction of propagation of the particle. Note that these solutions
are2 eigenvectors of a complete commuting set of operators consisiting of the three
momentum operators ~pˆ and the free particle Hamiltonian.
The free particle Schrodinger equation Eq.(7.2) can also be solved by separating
variables in spherical polar coordinates and we get the solution
87
The two sets of eigenfunctions in Eq.(7.6) and Eq.(7.7) form two different bases for
in the subspace of solutions of given energy E; every function in Eq.(7.6) can be
expanded in terms of the functions in the set Eq.(7.7) and also, every function in
Eq.(7.7) can be expanded in terms of the functions in the set Eq.(7.6). Therefore, we
can write the plane wave solutions, Eq.(7.6), as linear combinations of the spherical
wave solutions Eq.(7.7).
X
exp(i~k · ~r) = Cℓm jℓm (kr)Yℓm (θ, φ). (7.8)
ℓm
We shall now consider a special case when ~k is along the z− axis ~k = (0, 0, k) and
~k · ~r = kr cos θ. The left hand side depends only on cos θ and is independent of φ.
Hence only m = 0 terms will be present in the right hand side giving Cℓm = 0 if
m 6= 0. Also noting that
r
2ℓ + 1
Yℓ0 (θ, φ) = Pℓ (cos θ), (7.9)
4π
the Eq.(7.8) takes the form
∞
X
exp(ikr cos θ) = Aℓ jℓ (kr)Pℓ (cos θ). (7.10)
ℓ=0
X∞
2
= Aℓ jℓ (kr) δℓm , (7.12)
ℓ=0
2ℓ + 1
88
Thus as r → ∞ we get
Z 1
1
exp(ikrt)Pn (t)dt −→ eikr − (−1)n e−ikr + O(1/r 2) (7.18)
−1 ikr
einπ/2 ikr−inπ/2
= e − e−ikr+inπ/2 (7.19)
ikr
2einπ/2
= sin(kr − nπ/2). (7.20)
kr
Using Eq.(7.15) and Eq.(7.20) in Eq.(7.14) we get
Thus the expansion of the plane waves, Eq.(7.10) takes the form
∞
X
ikz ikr cos θ
e =e = (2ℓ + 1)iℓ jℓ (kr)Pℓ (cos θ) (7.22)
ℓ=0
A formula similar to Eq.(7.22) can be written down when the plane wave propagates
in a direction other than the z− axis. Let α, β be the polar angles of the direction
of propagation n̂ so that
89
in Eq.(7.24) we get
∞ X
X ℓ
i~k·~
r
e = 4π (2ℓ + 1)iℓ jℓ (kr)Yℓm
∗
(α, β)Yℓm(θ, φ) (7.26)
ℓ=0 m=−ℓ
For a sperically symmetric potential, the operators L ~ 2 and Lz commute with H and
are constants of motion. Thus if incident particle has a definite value of L~ 2 and Lz ,
it contunues to have the same values at all times and the problem of finding the
scattering amplitude is solved by finding the amplitude for different values of L ~2
and Lz separately. More over, it is sufficient to find the solutions for Lz = 0. This
is because if we select the z − axis parallel to the incident momentum ~ki , the initial
value of Lz will be equal to zero and will remain zero at all times. Recall that
~ˆ z = −i~ ∂
L
∂φ
90
3 Relate the large r behaviour of the radial wave function to the scattering
amplitude and hence to the cross section.
QM-
Lecs-
19003
⊲ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-Lecs-
19002;
We assume that the potential has a finite range and that the potential V (r) → 0,
as r → ∞, faster than r12 , i.e., r 2 V (r) → 0 as r → ∞. Thus for large r, V (r) can
be neglected as compared to ℓ(ℓ + 1)~2 /2 µr 2 term in the radial equation
~2 1 d 2 dRℓ (r) ℓ(ℓ + 1)~2
− r + E − V (r) − Rℓ (r) = 0 (7.33)
2µ r 2 dr dr 2µr 2
which for large r assumes the form
~2 1 d 2 dRℓ (r) ℓ(ℓ + 1)~2
r + E− Rℓ (r) ≈ 0 (7.34)
2µ r 2 dr dr 2µr 2
or
1 d 2 dRℓ (r) 2 ℓ(ℓ + 1)
r + k − Rℓ (r) ≈ 0 (7.35)
r 2 dr dr r2
The most general solution of this equation is a linear combination of the spherical
Bessel functions jℓ (kr) and nℓ (kr)
r→∞
Rℓ (r) ≈ Aℓ jℓ (kr) + Bℓ nℓ (kr) (7.36)
Note even though the radial equation looks like a free radial particle equation, the
combination Eq.(7.36) is approximate solution for large r only, where as for the free
particle solution is jℓ (kr) for all r, a term nℓ (kr) is absent in the free particle solution.
Since we interested only in the large r behaviour of the radial wave function, we use
the following asymptotic expansions for spherical the Bessel functions:
1
jℓ (kr) ≈sin(kr − ℓπ/2) (7.37)
kr
1
nℓ (kr) ≈ − cos(kr − ℓπ/2) (7.38)
kr
91
Comparing the scattering solution for the potential Eq.(7.44) with the free
particle solution Eq.(7.40), we find that the Eq.(7.44) is shifted by a phase
δℓ given by Eq.(7.44). This quantity δℓ is called the phase shift for the ℓth
partial wave. The phase shift is a function of energy and of course the angular
momentum ℓ and carries all the information about the scattering for angular
momentum ℓ.
92
A′ℓ
Rℓ (r) ≈ sin(kr − ℓπ/2 + δℓ ) (7.47)
kr
in the boundary condition to be required of the wave function for large r
eikr
ψ(~r) −→ exp(ikz) + f (θ, φ) (7.48)
r
and collect the coefficients of exp(±ikr) in the the sides of Eq.(7.48). This gives us
the following expression for the left and the right hand sides of Eq.(7.48).
Left hand side of Eq.(7.48) The wave function is a superposition of the radial
solution for different partial waves and we have
∞
X
ψ(~r) = Cℓ Rℓ (r)Pℓ (cos θ) (7.49)
ℓ=0
X∞
1
≈ A′ℓ Pℓ (cos θ) sin(kr − ℓπ/2 + δ)ℓ ) (7.50)
ℓ=0
kr
X∞
1 i(kr−ℓπ/2+δℓ
= A′ℓ Pℓ (cos θ) e − e−i(kr−ℓπ/2+δℓ (7.51)
ℓ=0
2ikr
∞
1 X ′
= Aℓ Pℓ (cos θ)e−iℓπ/2 ei(kr+δℓ − (−1)ℓ e−ikr−δℓ (7.52)
2ikr ℓ=0
eikr X ′
∞
−iℓπ/2 iδℓ
= A Pℓ (cos θ)e e (7.53)
2ikr ℓ=0 ℓ
e−ikr X ′
∞
− Aℓ Pℓ (cos θ)(−1)(ℓ+1) e−iδℓ (7.54)
2ikr
ℓ=0
Right hand side of Eq.(7.48) We make use of the plane wave expansion in the
right hand side to get
93
eikr
X∞
= 2ikf (θ) + (2ℓ + 1)iℓ Pℓ (cos θ)e−iℓπ/2 (7.58)
2ikr ℓ=0
−ikr X
∞
e
+ (2ℓ + 1)iℓ Pℓ (cos θ)(−1)eiℓπ/2 (7.59)
2ikr ℓ=0
Since the exponentials eikr and e−ikr are linearly independent functions, their coef-
ficients in Eq.(7.48) must be equal giving
∞
X ∞
X
ℓ −iℓπ/2
2ikf (θ) + (2ℓ + 1)i Pℓ (cos θ)e = A′ℓ Pℓ (cos θ)e−iℓπ/2 eiδℓ (7.60)
ℓ=0 ℓ=0
and
∞
X ∞
X
ℓ iℓπ/2
(2ℓ + 1)i Pℓ (cos θ)(−1)e = A′ℓ Pℓ (cos θ)(−1)(ℓ+1) e−iδℓ (7.61)
ℓ=0 ℓ=0
Because the Legendre polynomials Pℓ (cos θ) are linearly independent their coeffi-
cients in the two sides of Eq.(7.61) must be equal. This gives
A′ℓ e−iℓπ/2 (−1)ℓ+1 e−iδℓ = (2ℓ + 1)iℓ eiℓπ/2 (7.62)
Hence we get the result
A′ℓ = (2ℓ + 1)iℓ δℓ (7.63)
ℓ iℓπ/2
Substituting Eq.(7.63) in Eq.(7.60) and noting i = e , we get
∞
X ∞
X
2ikf (θ) = − (2ℓ + 1)Pℓ (cos θ) + (2ℓ + 1)e2iδℓ Pℓ (cos θ) (7.64)
ℓ=0 ℓ=0
∞
X
= (2ℓ + 1)Pℓ (cos θ) e2iδℓ − 1 (7.65)
ℓ=0
94
QM-
§5 Phase Shifts for Square Well
19-
03- ⊲
L1 As an example of calculation of phase shifts, we will compute the phase shifts for a
square well potential. (
−V0 if r < R0
V (r) = (7.70)
0 if r > R0
The radial equation for a spherically symmetric potential V (r) is
~2 1 d 2 dRℓ (r) ℓ(ℓ + 1)~2
− r + E − V (r) − Rℓ (r) = 0 (7.71)
2µ r 2 dr dr 2µr 2
The potential for r < R0 is −V0 and the most general solution of the radial equation
is
(I)
Rℓ (r) = αjℓ (qr) + βnℓ (qr) r < R0 (7.72)
where q 2 = 2µ(E + V0 )/~2 and for r > R0 the potential is zero and the most general
solution is
(II)
Rℓ (r) = α′ jℓ (kr) + β ′ nℓ (kr), r > R0 , (7.73)
where k 2 = 2µE/~2. Next we must impose regularity conditions on the solutions.
Since nℓ (ρ) blows up as ρ → 0, we demand that β in Eq.(7.72) must be zero.
Next the radial solution and its first derivative must be continuous at all points, in
particular at r = R0 . These two conditions give the restrictions
95
For ℓ = 0, using the expressions for the spherical Bessel functions in terms of sine
and cosine functions, for s− wave one can easily obtain
k
tan (δ0 (k) + kR0 ) = tan(qR0 ) (7.81)
q
Threshold behaviour of the phase shifts We can now derive the behaviour of
the phase shifts, for a square well potential, for small k. When kR0 ≪ 1 we have
1
jℓ (kR0 ) ∼ (kR0 )ℓ nℓ ∼ (7.82)
(kR0 )lℓ+1
1
jℓ′ (kR0 ) ∼ (kR0 )ℓ−1 n′ℓ ∼ (7.83)
(kR0 )lℓ+2
This gives
δℓ (k) ≃ (kR0 )2ℓ+1 (7.84)
Therefore, as momentum k goes to zero, the ℓth partial wave phase shift goes to
zero as (2ℓ + 1) power of k. This result is generally true for finite range potentials.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-19-03-
L1;
96
QM-
§1 Introduction
24-
01- ⊲
L1 Not all problems can be solved exactly. Therefore, we need approximation methods.
We have so far discussed approximation methods for energy eigen values and eigen
functions. These methods give approximate solution of the eigenvalue problem
Hψn = En ψn . (8.1)
The next class of problems, we are interested, require us to solve the time dependent
Schrödinger equation
∂ψ
i~ = Ĥψ. (8.2)
∂t
The exact solution of Eq.(8.2) can be written as
−iĤ(t − t )
0
ψ = exp ψ(x, t0 ) (8.3)
~
where ψn are eigenfunctions of the full Hamiltonian and Cn are the expansion coef-
ficients which are computed from the knowledge of the wave function at an initial
time t = t0 . Thus
X
ψ(x, t0 ) = Cn ψn (x) (8.5)
Zn
Cn = ψn∗ (x)ψ(x, t0 )dx. (8.6)
97
H = H0 + H ′ (8.7)
• If the time dependence of the Hamiltonian is small, it varies slowly with time,
the approximation scheme is known as adiabatic approximation. It is assumed
that the instantaneous eigenvalues and eigenvectors of the Hamiltonian are
known at ecah instant of time.
QM-
§2 Time Dependent Perturbation Theory
24-
02- ⊲
L1 In the perturbation theory, one assumes that the total Hamiltonian can be spilt as
H = H0 + H ′ (8.9)
where H0 is independent of time and is such that its eigenvalues and eigen functions,
i.e. the solutions of the equation
H0 un = En un (8.10)
are known exactly. H ′ is a perturbation and its matrix elements are assumed to be
small compared to the matrix elements of H0 . H ′ may or may not, depend on time,
but H0 must be independent of time. We shall write
H = H0 + λH ′ (8.11)
98
QM-
§3 Transition Amplitude
24-
03- ⊲
L1 We shall at first derive an exact equation for amplitude for transition from an initial
state ui to a final state uf .
We shall start with the expression
H = H0 + λH ′ (8.12)
where λ to be set equal to 1 in final answer. The Schrodinger equation is
∂ψ
= Hψ i~ (8.13)
∂t
We assume that at time t = t0 the wave function of the system is given to be φ0 (~r)
and we want to compute (approx) wave function at time t.
It is assumed that the eigenfunction and corresponding eigenvalues of H0 are
known exactly and will be denoted as un and En :
H0 u0 = En un . (8.14)
To solve Eq.(8.13) we expand ψ(x, t) as a sum in terms of the stationary states of
H0 X En (t−t0 )
ψ(x, t) = Cn (t)e−i ~ un (x), (8.15)
where Cn , in general, depends on time, because the total Hamiltonian H is different
from H0 . At time t = t0 the wave function is given to be φ0 (~r). Therefore, setting
t = t0 , in Eq.(8.15) we get
X
φ0 (~r) = Cn (t0 )un (~r). (8.16)
n
99
Take the scalar product with um (x) and use the orthogonality property of u’s to get,
dC X
m
i~ e−iEm (t−t0 )/~ = e−iEn (t−t0 )/~ (um , H ′ un ) Cn (t) (8.22)
dt n
or
dCm X i(Em −En )(t−t0 )/~
i~ = e (um , H ′un )Cn (t). (8.23)
dt n
(Em −En )
Define ωmn = ~
and write Eq.(8.23) in the form
dC X
m
i~ = eiωmn (t−t0 ) (um , H ′un )Cn (8.24)
dt n
QM-
§4 Perturbation Theory
24-
03- ⊲
L2 We recall some of the equations from the previous section.
H = H0 + H ′ (8.25)
H0 un = En un (8.26)
ψ(x, t) = φ0 (x) (8.27)
t=t0
X
ψ(x, t) = Cn (t) exp(−iEn (t − t0 )/~)un (x) (8.28)
n
where the coefficients Cn are functions of time and satisfy an exact equation given
by
d X
i~ Cn (t) = exp(iωmn (t − t0 ))hm|H ′ |niCn (t) (8.29)
dt n
100
101
∆E ≈ 2π(~/t) (8.43)
Fig. 8.1.
QM-24-F02;Golden-Rule
Note that the area under the peak increases as t. Thus if we compute the transition
probability at time t, given by
Z E+∆E
(1)
|Cf (t)|2 |dE, (8.44)
E−∆E
to a set of states in the energy range E and E ± ∆E, the answer will be proportional
to time. In the case of transitions to a state in continuum, the quantity of interest
is the rate of transitions to a group of final states having the energy in the range
102
2π
wi→f = |hf |H ′|ii|2 ρ(Ef ) (8.52)
~
The analysis proceeds by keeping only one of the two terms in Eq.(8.40) and
showing that the other term is not important. The final energy Ef can have only
one of the two values Ei + ~ω or Ei − ~ω only.
Density of states
If we write the sum over final states as an integral over energy and a sum over
remaining variables, such as angles,
X X′ Z
(.) = dEf (.). (8.53)
final states final
103
which states that the density of states equals the number of states with energy in
the range E, E + dE is ρ(E)dE.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-24-03-
L3;
The allowed values of ~k in a box are kx = 2πnx /L, ky = 2πny /L, kz = 2πnz /L etc.
where nx , ny , nz are positive integers. There will be (L/2π)3 dkx dky dkz states for the
propagation vector in the range ~k and ~k + d~k. The range dk is related to the range
dE of energy given by
~2 k 2 dE ~2 k
E= ⇒ = . (8.57)
2µ dk µ
and the number of states is
L3
ρ(E)dE = dkx dky dkz (8.58)
(2π)3
where ρ(E) is the density of states. For differential cross section we need to compute
the transtion probability to states with final propagation vector in a small range of
angles θ, φ. The number of states with the direction ~k in the solid angle dΩ =
sin θ dθ dφ and magnitude in a small range dk is given by k 2 dΩ. Using Eq.(8.57), we
get
L3 2 L3 µ k
ρ(E)dE = k dk dΩ ⇒ ρ(k) = dΩ. (8.59)
(2π)3 8π 3 ~2
104
QM-
§6 Probability for Resonance Transitions
24-
03- ⊲
L4 The case of periodic perturbation with a single frequency is an important one for
many physical situations including the interaction of radiation with matter. Assum-
ing that H ′ varies periodically with time with a single frequency ω we write
105
In the perturbation approximation after integration, the large coefficients came from
those terms which were multiplied with an exponential with a small argument. For a
given ω when there are two energy levels i and f such that |Ef −E −i| matches with
~ω, we need to retain all the terms involving the two coefficients Ci and Cf in the
summation in the right hand side of Eq..(8.69) . Thus the resulting approximate
equations to be solved assume the form
dCf (t)
i~ = exp i(ωf i + ω)t hf |F |iiCi(t)
dt
+ exp i(ωf i − ω)t hf |F †|iiCi (t). (8.70)
and
dCi (t) X
i~ = exp i(ωif + ω)t hi|F |f iCf (t)
dt n
+ exp i(ωif − ω)t hi|F † |f iCf (t). (8.71)
. In these equations we retain only those exponentials which have small arguments.
Taking ~ω ≈ (Ef − Ei ), using the notation ν ≡ ωf i − ω, and therefore writing
ωif + ω = −ν, we get
dCf (t)
i~ = hf |F † |iieiνt Ci (t) (8.72)
dt
dCi(t)
i~ = hi|F |f ie−iνt Cf (t) (8.73)
dt
Next we solve these equations exactly with the initial conditions Ci (0) = 1, Cf (0) =
0. The probability of transition from the initial level Ei to the final level Ef at time
t is then given by
2|hf |F |ii|2
Pi→f (t) = | {1 − cos Ωt} , (8.74)
~2 Ω2
where
~2 ν 2 + 4|hf |F |ii|2
Ω2 = . (8.75)
~2
106
QM-
Details of Solution
24-
03- ⊲
L5
For the resonance transitions the equations satisfied by the coefficients Ci and Cf ,
Eq..(8.72) and Eq..(8.73) , are
dCf (t)
i~ = hf |F † |iieiνt Ci (t) (8.78)
dt
dCi(t)
i~ = hi|F |f ie−iνt Cf (t) (8.79)
dt
In this section we solve these equations exactly and obtain expressions for Ci (t) and
Cf (t). To solve we define
bf = Cf exp(−iǫt) (8.80)
so that
d d iǫt
Cf (t) = bf e (8.81)
dt dt
d
= bf + iǫbf eiǫt (8.82)
dt
Eliminating Cf Eq..(8.78) and Eq..(8.79) , using Eq..(8.80) and Eq..(8.81) , we
get
1
Ċi = hi|F |f ibf (8.83)
i~
and
1
ḃf + iǫbf = hf |F †|iiCi
i~
1
= hf |F |ii∗Ci (8.84)
i~
Eliminating Ci from Eq..(8.82) and Eq..(8.84) we get
1
b̈f + iǫḃf = hf |F †|iiĊi
i~
|Ff i |2
= − 2 bf (8.85)
~
107
and hence h i
Cf (t) = A exp(iα+ t) + B exp(iα− t) exp(iǫt) (8.92)
Fif∗
A(ǫ/2 + ∆) + B(ǫ/2 − ∆) = − (8.96)
~
or
Fif
2A∆ = − (8.97)
~
Fif
A = − (8.98)
2∆~
Fif∗
B = (8.99)
2~∆
108
QM-
I-Transition to a discrete level
24-
02- ⊲
L2
We assume that under the action of perturbation the transition takes place to an-
other state of system which corresponds to a discrete level. In this case Eq.(8.38)
is the basic formula, and |Cf i (t)|2 gives the probability of a transition at time t from
the initial state i to the final state f . We list a few cases of interest.
109
110
QM-
II-Transitions to a set of final states in continuum
24-
02- ⊲
L3
The cases of interest are further grouped according to the time dependence of the
perturbation Hamiltonian H ′ (t)
Interaction H ′ (t) switched on adiabatically
In the previous case we have considered an example in which a charged har-
monic oscillator is subjected to perturbation 21 mω 2 x2
H ′ = V0 exp(−t/t0 ) (8.114)
Suppose we place a hydrogen atom between the plates of charged capacitor
which is subsequently discharged through a resistance. We may then again ask
a question: what is the probability that if at time t = 0 H-atom is in an initial
state n, after some time t, it will be found in some final, excited, state m. The
question is similar to the above example of charged harmonic oscillator and
falls under the class of problems mentioned in I.
However, the H- atom offers one more possibility. Under the action of external
perturbation, the atom may get ionized and the final energy then does not
correspond to a discrete energy level. The transition in this case takes place
to a level in the continuum.
Perturbation H ′ (t) is independent of time
This is, for example, the case for scattering; we are interested in knowing the
probability per unit time that a particle gets scattered into solid and dΩ. The
differential scattering cross section, σ(θ, φ) is related to such a probability.
prob per unit time of particle getting scattered in solid angle d Ω
σ(θ, φ) =
prob per unit time of a particle in incident beam crossing a unit area
(8.115)
In case of scattering
P2
H = H0 + V (r) H0 =
(8.116)
2m
and V (r) is to be treated as perturbation (H′ ≡ V (r)). Here the initial state
corresponds to plane waves with momentum p~i and final states corresponds to
momentum ~pf . The initial and final states are eigenstates of the unperturbed
Hamiltonian H0 .
In the presence of the potential, the momentum is not conserved and hence
the possibility of momentum changing to some final value p~f after some time.
Note that here (H′ = V (r)) is not dependent on time, but we are still using
time dependent method because the question concerns, time evolution of eigen
states of H0 under the action of full Hamiltonian H.
H ′ (t) varies harmonically with time
An example of interest here is ionization of an atom in presence of an external
periodic electric field which varies with a single frequency.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : .................................. ◭ QM-24-02-
L3;
111
112
|h0|H0|1i|2
(~/T )2 + (∆E)2
where ∆E is the energy difference betwen the states |0i and |1i.
[2] A particle of charge e is confined to a cubical box of side 2b. An electric field
~ given below is applied to the system.
E
~ = (0 t<0
E
E~0 exp(−αt) t > 0)
where α > 0, The vector E0 is perpendicular to one of the surfaces of the box.
To the lowest order in E0 calculate the probability that the charged particle,
in the ground state at time t = 0, is excited to the first state at time t = ∞.
(a) Assume that the continuous energy wave functions may be approximated
by the free particle wave functions, find the density of states as ρ(E) as
function of energy E
(b) Assuming that the electric field may be treated as a perturbation find the
probability that the particle will be found in a continuous energy state
with energy in the range E and E + dE if at time t = 0 it was known to
be in the bound state.Sss
You may assume that the normalized bound state wave function for the delta
function potential V (x) = −γδ(x) is given by,
r
mγ
u(x) = 2
exp(− mγ
~2
|x|)
~
113
QM-
§1 Electromagnetic Waves in Vacuum
25-
01- ⊲
L1 We shall first discuss the interaction of a charged particle with electromagentic waves
in the classical Maxwell’s theory. Classically the electromagnetic fields are described
by the four Maxwell’s equations which written in cgs units take the form
~ = ρ
∇·E (9.1)
ǫ0
~
~ = − dB
∇×E (9.2)
dt
~
∇·B = 0 (9.3)
~
~ = µ0 J~ + µ0 ǫ0 dE
∇×B (9.4)
dt
Here ρ is the charged density and J~ is the current density which satisfy the continuity
equation
dρ
+ ∇ · J~ = 0 (9.5)
dt
~ = 0
∇·E (9.6)
~
dB
~ = −
∇×E (9.7)
dt
~ = 0
∇·B (9.8)
~
~ = µ0 ǫ0 dE
∇×B (9.9)
dt
114
1 dE~
~ =
∇2 E (9.10)
c2 dt
~
~ = 1 dB
∇2 B (9.11)
c2 dt
~
∇× E~ + dA = 0 (9.13)
dt
This shows that there exists a scalar function φ such that
~
~ + dA = −∇φ
E (9.14)
dt
Thus the electric field can be written as
~
~ = −∇φ − dA
E (9.15)
dt
~
dΛ
φ′ = φ − (9.16)
dt
~ ′ ~
A = A + ∇Λ (9.17)
give the same answers for the electric and the magnetic fields where Λ is an arbitrary
function of space time. Since all observable quantities are functions of the electric
and magnetic fields, they are also the same for the two sets of potentials. Thus for
example the Lorentz force
F~ = q E~ + ~v × B
~ (9.18)
experienced by a charged particle in electric and magnetic field does not depend
on the choice of the function Λ in Eq..(9.16) and (9.17). The transformation
~ φ) → (A
(A, ~ ′ , φ′), given by (9.16) is called a gauge transformation and we say
that the observable quantities are invariant under gauge transformation.
115
φ = 0, ~=0
∇·A (9.19)
For such a choice of gauge condition the two wave equations Eq..(9.10) -(9.11)
become equivalent to the wave equation
1 dA~
~=
∇2 A (9.20)
c2 dt
where ~k is the propagation vector, ω = ck is the frequency and the amplitude vector
~ 0 ( satisfies the transversality condition
A
~k · A
~ 0 = 0. (9.22)
116
p~ −→ b
p~ = −i~∇ (9.28)
QM-
§3 Semiclassical Treatment of Radiation
25-
02- ⊲
L1 We will describe the interaction of an electron in an atom with e.m. waves in
a semiclassical fashion. The electron will will be teated quantum mechanically,
whereas the e.m. waves will be treated classically. We shall ignore the effect of
motion of the electron on the e.m. waves. We will interested in finding out what
are allowed transitions of an electron that can take place in presence of e.m. waves
and we want to compute the corresponding transition probabilities.The effect of
117
φ=0 ~ = 0.
∇·A (9.36)
H = H0 + H ′ (9.37)
~p2
H0 = + V (~r) (9.38)
2m
q ~ q2 ~ 2
H′ = A · p~ + A (9.39)
m 2m
where H0 describe the electron in the atom, and H ′ gives the effect of e.m. waves
QM-
on the electron and will be treated as a perturbation.
25-
02- ⊲ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Status : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ◭ QM-25-02-
L1;
L2
118
Perturbation theory
The amplitude of transition from an initial state |ii to a final state |f i is given by
2 1 2 1
2 ′ 2 sin 2 (ωf i − ω)t ′ 2 sin 2
(ωf i + ω)t
|Cf i (t)| = 4|hf |H |ii| + 4|hf |H |ii| + Cross terms
(9.48)
~2 (ωf i − ω)2 ~ (ωf i + ω)2
2
where ωf i = (Ef − Ei )/~. The case Ef > Ei , with ωf i − ω ≈ 0 the first term is most
important and this corresponds to absorption. The second case when, Ef < Ei and
ωf i + ω ≈ 0, hold correspond to emission of radiation. Both emission and absorption
can take place only in presence of radiation. The emission process in presence of
external radiation is called induced emission.
In cases when the incident radiation is an incoherent superposition of radiation
of several frequencies, we can add probabilities for different frequencies, and the
probabilty of transition after time t will be given by
Z ∞
Pf i (t) = |Cf (t)|2 dω. (9.49)
0
We now repeat the same steps as in the derivation of Fermi Golden rule. To compute
the transition probability per unit time, we take the time derivative of (9.49), and
the limit t → ∞ gives Dirac δ function which can be used to do the ω integral to
give the transition rate
πq 2
wf →i = I(ω)|hf |ê · ~r|ii|2, (9.50)
3~2 cǫ0
119
QM-
§4 Einstein A and B Coefficients
25-
03- ⊲
L1 The semiclassical theory gives a reasonably good description of atomic transitions
induced by a light incident on an atom. It fails to give an explanation for ob-
served spontaneous emission of radiation when an atom is in an excited state. Ein-
stein showed how to compute the transition probability for induced emission using
Planck’s law of black body radiation.
We assume that atoms are placed inside a cavity. When the atoms are in an
equilibrium with radiation at temperature T , the number of photons of a particular
frequency emitted per sec must be the same as the number of photons absorbed
per sec by the atoms. Note that the probability of absorption per unit time as
well as the probability of induced emission per unit time is given by the same
expression, Eq.(9.50) , let’s call it A. Let N1 and N2 denote the number of atoms
in levels E1 and E2 with E2 > E1 . When in equilibrium at temperature T , the
number of atoms in the level with energy Ek , will be proportional to the Boltzmann
factor exp(−Ek /kB t). Let B denote the transition probability per unit time for the
induced emission to take place in the state E2 . We therefore have
The total number of transitions from E1 to E2 (absorption)
= N1 × Probability per unit time for induced absorption
120
121