0% found this document useful (0 votes)
134 views

Quantum Mechanics

This document discusses separating the Schrodinger equation in spherical polar coordinates for a central potential that depends only on the radial distance r. It relates Cartesian and spherical polar coordinates, defines the Laplacian operator and Hamiltonian in spherical coordinates. It separates the Schrodinger equation into radial and angular parts, and solves the angular equation to obtain the spherical harmonics Ylm(θ,φ). The key steps are: 1) Separating the Schrodinger equation into radial and angular parts. 2) Solving the angular equation subject to boundary conditions to obtain the spherical harmonics and quantum numbers l and m. 3) Expressing the orthonormality of the spherical harmonics.

Uploaded by

randima fernando
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
134 views

Quantum Mechanics

This document discusses separating the Schrodinger equation in spherical polar coordinates for a central potential that depends only on the radial distance r. It relates Cartesian and spherical polar coordinates, defines the Laplacian operator and Hamiltonian in spherical coordinates. It separates the Schrodinger equation into radial and angular parts, and solves the angular equation to obtain the spherical harmonics Ylm(θ,φ). The key steps are: 1) Separating the Schrodinger equation into radial and angular parts. 2) Solving the angular equation subject to boundary conditions to obtain the spherical harmonics and quantum numbers l and m. 3) Expressing the orthonormality of the spherical harmonics.

Uploaded by

randima fernando
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 24

92

Separation of the Schrodinger Equation in Spherical Polar Co-ordinates

Consider the motion of a particle of mass m in a central (radial) potential V(r) (i.e. V(r) depends

only on r  r ). Since V(r) is then spherically symmetric, it is convenient to use spherical polar
co-ordinates in this problem.

The Cartesian co-ordinates (x, y, z) of the point P and the spherical polar co-ordinates (r, , ) are
related by

x  r sin  cos  , y  r sin  sin  , z  r cos  , where 0  r   , 0     , 0    2 .

In spherical polar co-ordinates (r ,  ,  ),


1   2  1     1 2
2   r    sin   
r 2 r  r  r 2 sin      r 2 sin 2   2


d r  d 3 r  r 2 d r sin  d d  r 2 d r d, where d = sin  d d

Hamiltonian of the particle,


2 2 2  1   2   Lˆ2 
Hˆ     Vˆ (r )     r     Vˆ (r ) ,
2m 2 m  r 2 r  r  2 r 2 
 1     1 2 
where Lˆ2   2   sin    
 sin      sin 2   2 

TISE for the particle,


  2  1   2   Lˆ2    
  2  r   2 2
 Vˆ (r )   (r )  E (r )
 2 m  r r  r   r  

Rearranging terms, assuming a solution of the form  (r )   (r ,  ,  )  R(r )Y ( ,  ) and
separating variables
93

1   2   2 m r2 1
r
R ( r ) r  r
R(r )  
  2 
E  Vˆ ( r )  2 
 Y ( ,  )
Lˆ2Y ( ,  )   (constant)

1   2   2m 
r 2 r 
r
 r  
 
R(r )    2 E  Vˆ (r )  2  R(r )  0 … (1) Radial Equation, to be
r 
solved for R(r)
ˆ2 2
L Y ( ,  )    Y ( ,  ) …………… (2) Angular Equation, to be solved for Y ( ,  )

Let Y ( ,  )  ( )  ( ) , substitute in Equation (2) with the definition of L̂2 and separate
variables.
sin      2 1  2 ( )
 sin   ( )    sin    2
 m 2 (constant)
( )      ( ) 
1      m2 
  sin   ( )       ( )  0 ……. (3)
sin       sin 2  
d 2  ( )
 2
 m2  ( )  0 ………………………………(4)
d
Equation (4) has a particular solution,  ( )  ei m 
Using the requirement that eigenfunctions be single-valued,
 ( )  (  2 ) (since  and  + 2 are the same angle)
When  = 0,   (0)  (2 )  1  cos 2 m  i sin 2 m
 m  0, 1, 2, 3, .......... or m  0,  1,  2,  3, ........
Normalization of  ( ) requires that, when  ( )  Cei m  ,
2
2 1
 ( ) d  1  C  1 2   m ( )  ei m  , 0    2
0
2
In Equation (3), substitute w  cos  with 1  w  1
d  2 d   m2 
 
dw 
1 w
dw
P ( w) 

 

 
1  w2
 P( w)  0, where ( )  P (w)
 

Since the wave function must be finite everywhere, P( w) and  ( ) must be finite everywhere.
All possible solutions those for which P( w) is finite everywhere exist, only if

(i)   l (l  1), where l  0, 1, 2, 3, ..............

(ii) m  l (i.e.  l  m  l and for a given value of l, there should be (2l + 1) values of m
such that  l ,  l  1,  l  2, ..... , 0, ...., ......, l  2, l  1, l .)

Such finite solutions of the above equation are called Associated Legendre functions Pl m ( w) of  
degree l and order m  l .

The angular part Y ( ,  ) (normalized) of the wave function  (r ,  ,  ) can be written as


Yl m ( ,  )  Cl m Pl
m
( w)  m ( ) - known as ‘Spherical Harmonics’  P(w)  P
l
m
(w) 
94

1 1
 (2l  1)(l  m)! 2 1  (2l  1)(l  m)!  2
m m
Clm  (1)     (1)  
 2(l  m)!  2  4 (l  m)! 
1
 (2l  1)(l  m)!  2
m
 Yl m ( ,  )  ( 1)  m
 Pl (cos  ) e i m , m  0
 4 (l  m)! 
 Note that, Yl m ( ,  )  Yl m ( ,  ) 
In Dirac notation, the spherical harmonics, Yl m ( ,  )  l m Yl*m ( ,  )  l m

Spherical harmonics satisfy the orthonormality relation

2 
* *
 
Yl m ( ,  ) Yl  m ( ,  ) d    Yl m ( ,  )  Yl m ( ,  ) d sin  d   l l  m m
0 0

In Dirac notation, the orthonormality condition, l m l  m   l l  m m

For spherical harmonics yielded by a given (fixed) value of l with


m   l ,  l  1,  l  2, .........,  1 , 0, 1, ........, l  2, l  1, l , the orthonormality condition
reduces to l m l m   m m .

Low-order spherical harmonics


l m Spherical harmonic Yl ml ( ,  )
0 0 Y0,0  1 4 
12

1 0 Y1,0   3 4 
12
cos 
12
+1 Y1,  1    3 8  sin  e  i
12
1 Y1,  1    3 8  sin  ei 

Some Important Properties of Yl m ’s


(1) Yl m ‘s form a complete set in the variables  and  , i. e. any angular function F ( ,  ) can
be expanded in terms of them.
 l *
F ( ,  )   
l  0 m  l
al m Yl m ( ,  )  al m   Yl m ( ,  )  F ( ,  ) d

(2) Lˆ2 Yl m ( ,  )  l (l  1)  2Yl m ( ,  ) (from equation (2))


i.e., Yl m ’s are eigenfunctions of L̂2 with the eigenvalue l (l  1)  2 .

   
(3) Behavior of Yl m ’s under the parity operation r   r  i.e. ˆ  (r )   ( r )  .

i.e. r  r,    ,    
95

z

 

  y

ˆ Yl m ( ,  )  Yl m (   ,    )  Pl m [cos(   )]  m (   )

m m l m m
Pl [cos(   )]  Pl [  cos  ]  (1) Pl (c os  )
m
 m (   )  (1)  m ( )
m
ˆ Yl m ( ,  )  (1)l 
  m
Pl (c os  )  (1)
m
 m ( )
ˆ Yl m ( ,  )  ( 1)l Yl m ( ,  )
  i.e. Yl m has the parity of l (‘even’ for even l and ‘odd’
for odd l.)
2
Physical properties of L̂

z  Consider a particle of mass m, having linear


 
t p momentum p and position vector r with respect to
pt origin O. Then, the angular momentum of the
 r   
pr particle with respect to O is defined as L  r  p .

 ( L is perpendicular to the plane containing
r m   
r and p . L  L  r p sin  , where  is the
 
angle between r and p )
y

x
  
Taking r̂ as a unit vector along r and tˆ as a unit vector in the plane of r and p and perpendicular

to r̂ , p  pr rˆ  pt tˆ  p 2  pr2  pt2

 
Since pˆ   i    pˆ 2    2  2  pˆ r2  pˆ t2
1   2   Lˆ2 2   2   L̂2
As shown before,  2   r      2 2
    r  
r 2 r  r  2 r 2 r 2 r  r  r2
2   2   Lˆ2
pˆ r2  pˆ t2   2 r   2  Lˆ2  r 2 pˆ t2
r r  r  r
96

r̂  tˆ is a unit vector along the direction of the orbital angular momentum. Therefore, we
2
identify L̂ as the operator corresponding to the ‘square of the orbital angular momentum’.

In all isolated systems (with no external torques), the total angular momentum is conserved. In
classical mechanics, angular momentum is an ordinary vector. But, in quantum mechanics, it is a
vector operator and its three components do not commute with each other.
 
L  ( Lx , Ly , Lz ) , L  L  r p sin 


The cartesian components of L are

Lx  y p z  z p y  Ly   z px  x pz  Lz   x p y  y px 
In q.m., the corresponding operators are

Lˆ x  yˆ pˆ z  zˆ pˆ y 
Lˆ y   zˆ pˆ x  xˆ pˆ z  Lz   xˆ pˆ y  yˆ pˆ x 
          
  i  y  z    i  z  x   i  x  y 
 z y   x z   y x 
ˆ  
 The above three relationships are represented by the vector operator L   i  rˆ    
ˆ
It can be verified that L is Hermitian.

Consider the commutation relation  Lˆx , Lˆ y 

 
  yˆ pˆ z  zˆ pˆ y ,  zˆ pˆ x  xˆ pˆ z     yˆ pˆ z , zˆ pˆ x    yˆ pˆ z , xˆ pˆ z    zˆ pˆ y , zˆ pˆ x    zˆ pˆ y , xˆ pˆ z 

To evaluate each commutator individually, consider  yˆ pˆ z , zˆ pˆ x 


  yˆ pˆ z , zˆ  pˆ x  zˆ  yˆ pˆ z , pˆ x     zˆ , yˆ pˆ z  pˆ x  zˆ  pˆ x , yˆ pˆ z 
   zˆ, yˆ  pˆ z pˆ x  yˆ  zˆ, pˆ z  pˆ x  zˆ  pˆ x , yˆ  pˆ z  zˆ yˆ  pˆ x , pˆ z 
=  i  yˆ pˆ x

Similarly, it can be shown that


 yˆ pˆ z , xˆ pˆ z   0 ,  zˆ pˆ y , zˆ pˆ x   0 ,  zˆ pˆ y , xˆ pˆ z   i  xˆ pˆ y

   
 Lˆx , Lˆ y   i xˆ pˆ y  yˆ pˆ x    Lˆx , Lˆ y   i  Lˆz
 

Similarly, it can be shown that  Lˆ y , Lˆz   i  Lˆ x  Lˆz , Lˆx   i  Lˆ y


   
 Operators representing any two components of the orbital angular momentum do not
commute. i.e. they are not simultaneously measurable with absolute precision.
ˆ ˆ ˆ ˆ ˆ
 The above three relationships are equivalent to L  L  i  L, Note: L  L  0 
ˆ ˆ 2
Since Lˆ2  Lˆ2x  Lˆ2y  Lˆ2z (Note: L̂ 2  L 2  L )

 Lˆ2 , Lˆx    Lˆ2x  Lˆ2y  Lˆ2z , Lˆ x    Lˆ2y , Lˆ x    Lˆ2z , Lˆ x   0 (prove this)


       
Similarly, it can be shown that  Lˆ , Lˆ y   0 ,
2  Lˆ , Lˆz   0

2

97

 Lˆ2 , Lˆ   0 
 The above three relationships are equivalent to
 
   

ˆ
Simultaneous eigenfunctions of L̂ 2 and one of L ’s components can be found. i.e. both the
magnitude and one component of angular momentum can be measured simultaneously with
absolute precision.

In spherical polar co-ordinates, it can be shown that,


        
Lˆ x   i    sin   cot  cos  , Lˆ y   i    cos   cot  sin  , Lˆ z   i 
        
 1     1 2 
 Lˆ2   2   sin    
 sin      sin 2   2 
Note that, Lˆ x , Lˆ y , Lˆ z and Lˆ2 are purely angular operators. (i.e. independent of r)
  Lˆx , fˆ (r )    Lˆ y , fˆ (r )    Lˆz , fˆ (r )    Lˆ2 , fˆ (r )   0
       
  1 i m  1 i m
Note that, Lˆ z  m ( )   i   e   m e  m   m ( )
  2 
 2 
 m  is an eigenvalue of Lz with eigenfunction  m ( ) , where m  0,  1,  2,  3, ........

Therefore, a measurement of the z-component of the orbital angular momentum can yield the
values 0,  1,  2,  3, ........ Since the z-axis can be chosen arbitrarily, the component of the
orbital angular momentum on any axis is quantized.
Lˆ z  m ( )  m   m ( )  Lˆz  m ( ) ( )  m   m ( ) ( )
 Lˆ Y ( ,  )  m  Y ( ,  )
z lm lm

The eigenvalues and eigenfunctions of L̂ 2 and Lˆz


Since  Lˆ2 , Lz   0 , we can look for simultaneous eigenfunctions of L̂ 2 and Lˆ z . We see that

Y ( ,  ) are simultaneous eigenfunctions of L̂ 2 and Lˆ , because


l ml z

Lˆ2 Yl ml ( ,  )  l (l  1)  2Yl ml ( ,  ) LˆzYl ml ( ,  )  ml  Yl ml ( ,  )

l  0, 1, 2, 3, .............. is called the orbital angular momentum quantum number.


 l  ml  l and ml  0,  1,  2,  3, ........ is called the orbital magnetic quantum number.
Note: In order to show the connection to l, we have replaced m by ml .

We designate the orbital angular momentum states corresponding to l  0, 1, 2, 3, .............. by


the symbols s, p, d, f, ………. (designation of the subshell system of atoms)

l 0 1 2 3
Designation s (sharp) p (principal) d (diffuse) f (fundamental)
98

[When there is more than one particle, we have L  0, 1, 2, 3, .............. as the total orbital
angular momentum quantum number and S, P, D, F, ……. are used to designate their states.]

The angular part of the wave function Yl m l ( ,  ) , is the same for all spherically symmetric
(central) potentials. The actual shape of the potential, V(r), affects only the radial part of the wave
function, R(r), which is determined by the radial equation

1   2   2 m l (l  1) 
r 2 r 
r
 r  

R(r )    2 E  Vˆ (r ) 
r 2 
 R (r )  0 …… (1)

By substituting u (r )  r R(r ) , Equation (1) can be simplified into the modified radial equation
 2 d 2 u (r )  l (l  1)  2 
  V ( r )   u ( r )  E u ( r ) …………………….(5)
2m d r 2  2 mr 2 
It is identical to the one-dimensional TISE for a particle of mass m and total energy E moving in a
l (l  1) 2
potential (effective), Veff (r )  V (r )  . The term added to V(r) is called the centrifugal
2mr2
term.

The Hydrogenic Atoms


We consider the hydrogenic atom to consist of a heavy motionless nucleus (kept at the origin) of
charge +Ze together with a much lighter electron of mass me and charge e that circles around it,
held in orbit by the mutual attraction of opposite charges. eg: H(Z = 1), He+(Z = 2), Li++(Z = 3),
Be3+(Z = 4), ……
( Ze)( e)  Z e2
From Coulomb’s law, the potential energy (in SI units) is V (r )   .
(4 0 )r (4 0 ) r
pˆ 2 2 2  Z e2
The Hamiltonian, Hˆ   Vˆ (r )    
2 me 2me (4 0 ) r
TISE for the electron reads,
  2  1   2   Lˆ2   Z e2   
   2  r       (r )  E (r )
 2me  r r  r  2 r 2  (4 0 ) r 

Since the potential is central, the angular part of  (r ) are spherical harmonics. Also, the radial
equation can be shown to take the form of the modified radial equation (with u (r )  r R(r ) ).

2 d 2u (r )   Z e2 l (l  1)  2 
     u (r )  E u (r ) (similar to equn (5))
2 me d r 2  (4 0 ) r 2 me r 2 

Solutions to the above equation can be obtained for

(i) continuum states (with E > 0), describing electron-proton scattering

(ii) discrete bound states (with E < 0), representing the hydrogenic atom

Here we confine our attention to bound states only. It can be shown that, the physically acceptable
solutions to this equation are obtained, when
99

2
m  Z e2  1
E   e2   2
2   4  0 n
where n  1, 2, 3, .............. and n is called the principal quantum number.

Note: For an arbitrary value of n, there are n number of possible values of l , as given by

l  0, 1, 2, .............., n  1 . (recall that, l  0, 1, 2, 3, .............. )

Energy Levels
The allowed bound state energy eigenvalues are given by

2
m  Z e2  1 E1
En   e2   2  2 , n  1, 2, 3, ..............
2   4  0  n n

This is the famous Bohr formula and Bohr obtained it in 1913 by using a mixture of classical
physics and premature quantum theory. The Schrodinger equation did not come until 1924.

The energy values can also be given in the form


e2 Z2 1 2 (Z )
2
En     me c ,
(4   0 ) a0 2 n 2 2 n2
(4   0 )  2
where the Bohr radius, a0  2
 0.529  10 10 m . ( = radius of the 1st Bohr orbit of H)
me e
 4  0 2  n2 a0 n 2
The radii of Bohr orbits are given by an    
 m e2
 e  Z Z

As n takes on values 1, 2, 3, .............,  , bound state energy spectrum of hydrogenic atoms


2
m  Z e2 
contains an infinite number of discrete levels from E1   e2   to zero. For atomic
2   4   0 
hydrogen (Z = 1), E1   13.6 eV . This negativeness in bound state energies provides stability to
the atom.

Spatial wave functions of different states in hydrogenic atoms are labeled by three quantum

numbers n, l and ml as  nl ml (r )  Rn l (r ) Yl ml ( ,  ) .

However, En depends only on n and is independent of (i.e. degenerate with respect to) l and ml .

Next, we can find the number of different states of this type possible for a bound level of
energy En , which is the degeneracy of that level.

For each value of n, l may take on values, l  0, 1, 2, .............., n  1 (i.e. n number of values)
100

For each l value, ml may take on values in the range, ml   l ,  l  1, ...  1, 0, 1, ........, l  1, l
(i.e. 2l + 1 number of values)
n 1
2n (n  1)
* The total degeneracy of the bound state energy level En =   2l  1   n  n2
l 0 2

The hydrogenic energy levels depend on n and are independent of l and ml . The degeneracy with respect
to ml is present for any central potential. But, the degeneracy with respect to l is a characteristic of the
Coulomb potential. (i.e. V ( r )  1 r ).

The n-fold degeneracy with respect to l is removed when the dependence of V(r) is modified to become a
non-Coulombic potential. Then, the energy of the electron becomes dependent on l and gives rise to n no. of
distinct energy levels as En l ( l  0, 1, 2, .............., n  1 ). Eg: Alkali atoms

The 2l + 1 - fold degeneracy w.r.t. ml is removed when an external magnetic field is applied. Then, the
energy becomes dependent on ml.

Spectroscopically, atomic states are labeled as n l . For eg: 1s, 2s, 2p, 3s, 3p, 3d, …..

Ground level (n = 1)  l=0  ml = 0  1s state only


 Total # of possible states = 1 (Degeneracy = n2 = 1)
First excited level (n = 2)  l=0  ml = 0  2s state
 l=1  ml = 1, 0, +1  2p 1 , 2p 0 , 2p 1 states
 Total # of possible states = 4 (Degeneracy = n2 = 4)

We see that the energy levels predicted by the Bohr theory and by Schrodinger theory agree with
each other. But, the Schrodinger theory is more powerful, because, the eigenfunctions it yields
enable us to calculate probability densities, expectation values of operators, transition rates, etc.
101

The radial eigenfunctions of bound states

It can be shown that, physically acceptable solutions to the radial equation are
Rn l ( r )  N e   2 . l L2n ll 1 (  ) ,
where N is a normalization constant and the associated Laguerre polynomials
n  l 1
[(n  l )!]2  k  2Z 
L2n ll 1 (  )   (1)k 1 with    r
k 0 (n  l  1  k )! (2 l  1  k )!  n a0 

2 2
Normalization of Rnl (r )   Rnl (r ) r dr  1
0
3  3 3
2 na  2  n a  2 n  ( n  l )!
2
 N  0 
 2l  L2n ll 1 (  )   2 d  1  N  0
 2Z 
e  
 2 Z  ( n  l  1)!
 1
0

Normalized radial eigenfunctions for the bound states of hydrogenic atoms are given as
12
 2 Z 3 (n  l  1)! 
  2 l 1
Rn l (r )     e  2 . l L n  l (  )
3
 n a0  2 n [(n  l )!] 

Using the above expression, the first few radial eigenfunctions can be written as follows.

n l Radial wave functions Rn l ( r )


1 0 32
R10 ( r )  2  Z a0  exp   Z r a0 
2 0 32
R20 (r )  2  Z 2 a0  1  Z r 2 a0  exp   Z r 2 a0 
1 1 32
R21 (r )  Z 2 a0  Z r a0  exp   Z r 2 a0 
3
3 0
R3 0 (r )  2  Z 3 a0 
32
1  2 Z r 3 a0  2 Z 2 r 2 
27 a02 exp   Z r 3 a0 

4 2
1 R31 (r )   Z 3 a0 3 2 1  Z r 6 a0  Z r a0  exp   Z r 3 a0 
9

4 32 2
2 R32 (r )  Z 3 a0  Z r a0  exp   Z r 3 a0  .
27 10

m 
We can use the known expressions of Rn l (r ) and Yl l ( ,  ) in  nl ml (r )  Rn l (r ) Yl ml ( ,  ) to
obtain the wave functions of the first few states of hydrogenic atoms as follows.
102

Shell Quantum Sub 


Wave functions  nl m ( r )
Numbers shell l

n l ml
 1
K 1 0 0 1s
 100 ( r )   Z a0 3 2 exp   Z r a0 

 1
L 2 0 0 2s
 200 (r )   Z a0 3 2 1  Zr 2a0  exp   Z r 2a0 
2 2
 1
2 1 0
2p0  210 (r )   Z a0 3 2  Zr a0  exp   Z r 2a0  cos 
4 2
 1
2 1 1 2p 1  21 1 (r )    Z a0 3 2  Z r a0  exp   Z r 2a0  sin  exp   i 
8 
 1
2p1
 21 1 ( r )    Z a0 3 2  Z r a0  exp   Z r 2a0  sin  exp   i 
2 1 1 8 

In the above work on hydrogenic atoms, it was assumed the mass of the nucleus to be infinitely
large when compared to the electron mass so that the nucleus remains fixed in space (at the origin).
This is a good approximation even for hydrogen, because mnuc  2000  me . However, for a better
agreement of the above theoretical results with accurate spectroscopic data, it requires the nuclear
mass to be taken into account as finite. In such a situation, the electron and the nucleus revolve
around their common centre of mass and it is possible to show that the electron moves around the
nucleus as though the nucleus is fixed and the mass of the electron is slightly reduced to a value 
called the reduced mass of the system. The equations of motion of the electron-nucleus system are
the same as those we have considered before, if we simply substitute  for me

me M
It can be shown that,   , where M is the mass of the nucleus. As
me  M
M      me .

Under this change, the more accurate expressions of energy eigenvalues and energy eigenfunctions
can be obtained by replacing me by  in the relevant expressions.

Eg:
2
  Z e2  1 e2 Z2
En     2  
2 2  4 0  n (4   0 ) a 2 n 2
(4   0 )  2
me
and the modified Bohr radius, a  a . 
 e2  0
The fraction by which the energy levels of a hydrogen atom is shifted =
 M 1,836
   0.99945 .
me me  M 1,837

The more accurate expressions of radial wave functions and energy eigenfunctions can be obtained
by replacing a0 by a in the relevant expressions.
32  1 32
Eg: 
R10 (r )  2 Z a  
exp  Z r a   100 ( r ) 

 Z a  exp   Z r a 
103

Consider an electron in a stationary state  nl ml ( r ,  ,  ) .


2 2
Position Probability Density =  nl ml (r ,  ,  )  Rn l (r ) 2 Yl ml ( ,  )

  
Probability of finding the electron in the volume element dr at the point r =  nl ml (r ,  ,  ) 2 d r

Probability of finding the electron in the shell of inner and outer radii r and r + dr
 2  2

2 2
=    nl ml (r ,  ,  ) d r =  d sin   d  nl ml ( r ,  ,  ) r 2dr
  0  0  0  0
 2
2
= r 2 Rn l (r ) d r  d sin   d Yl ml ( ,  ) 2

 0  0

2   2 
m 2
= r 2 Rn l ( r ) d r  since  d sin   d Yl l ( ,  )  1
 
  0  0 
= Dn l (r )d r ,
2
where the Radial Distribution Function, Dn l (r )  r 2 Rn2l (r ) , Rn l ( r )  Rn2 l (r )

Rn2l (r ) = Electron density along a given direction


2
Dn l (r )  r Rn2l (r ) = Probability of finding the electron within a unit length at a distance r
in any direction from the nucleus (i.e. within a shell of radius r and of
unit thickness).

Consider the ground state of hydrogen (Z = 1), n = 1, l = 0, ml = 0.


1
 100 (r )   1s (r )  1 a0 3 2 exp   r a0 

As can be seen from the plots given below,  1s ( r ) and  12s (r ) are exponentially decreasing
functions of r with their maximum occurring at the origin

To make a comparison of quantum mechanical results with those of Bohr theory, we look for the
Radial Distribution Function of the ground state.

32
D1 0 ( r )  r 2 R10
2
(r ) , where R10 (r )  2 1 a0  exp   r a0 

[Note: For states where  (r ) is independent of  and  (eg: 1s, 2s, 3s, …..)
2
Probability distribution function = 4 r 2  .]

For the first few radial functions, Rn l ( r ) vs. r and Dnl (r )  r 2 Rn2l (r ) vs. r plots are as given
below.
104

The maximum in D1 0 ( r ) curve plotted versus r indicates that the probability of detecting the
electron is maximum at some distance. This value of r is called the most probable distance and to
find it, we set
d
D1 0 (r )  0  r  a0
dr

So, Q. theory is similar to the Bohr picture of the hydrogen atom. However, there is a finite
probability of seeing the electron at other values of r between 0 and  and therefore the size of the
atom is not well-defined in Q. theory.

The value r can also be calculated for the ground state of hydrogen, as follows.
  2
 1 3
r   1*s (r ) r  1s (r )d r   exp( 2r
3
a0 ) r d r  sin  d  d  a0  m. p. d.
 a03 0 0 0
2
As we see, the mean value is found to be larger than the most probable distance. This is because of
the diffuseness incorporated by quantum theory to the electron orbit making it an ‘electron cloud’.
According to quantum theory, the electron can be found at distances from r = 0 to r   from the
nucleus with various probabilities (The maximum probability occurs at r  a ). So, their average
value is found to be equal to 1.5a . The Bohr theory predicts precisely a as the distance where
the electron can be found. However, in quantum theory, that is the distance where the electron can
be found with the maximum probability.
105

To obtain a reasonable estimate for the size of the atom, we find a value r(0.9) such that there is a
90% probability that the electron is at a distance r  r(0.9) .
r(0.9) r(0.9)
2 2 2
Prob.   r R10 (r ) d r  0.9   exp( 2r a0 ) r 2 d r  0.9  r(0.9)  2.6 a0 .
0 a03 0

For most purposes, an effective atomic radius is taken as from a0  2.5a0 ~ 10 8 cm = 1 Å.

Classically, a moving particle has a turning point when the total energy = potential energy. At this
point, the kinetic energy (hence the velocity) = 0 and the particle is expected to be reflected by the
potential barrier.

For an electron in the ground state of hydrogen atom, classical turning point occurs, when

 e2  e2
V (r )  E1    r  2 a0 .
(4 0 ) r 2(4   0 ) a0

As shown by D10 ( r ) versus r curves, there is a finite probability of finding the electron for
r  2a0 . i.e. the electron has access to a region which is forbidden by classical theory. This effect
of tunneling (penetration) through potential barrier is typical in q. theory results.
For values of r  2 a0  V  E , and in order to satisfy the condition E  T  V , the
kinetic energy T would have to be negative. This unrealistic situation can be explained by using
H. u. p. as follows.

In q.m., a particle is in a certain region means that its position measurement has been made (with
some uncertainty). According to the u. p.
uncertainty in position  uncert. in momentum  uncert. in kinetic energy
106

It can be shown that, when the electron is in the classically forbidden region, the uncertainty
introduced to its kin. energy is sufficiently large to compensate for the negative value required by
conservation of energy.

The energy eigenvalue equation (TISE) for a hydrogenic atom reads


  E
Hˆ  nl m (r )  En  nl m (r ) , where En  21 , n  1, 2, 3, ..........
l l
n
2
e Z2
Ionization potential in the nth shell, I P  En 
(4   0 ) a 2 n 2
m m
ˆ  Rn l (r ) Yl l ( ,  )   Rn l ( r )( 1)l Yl l ( ,  )
ˆ  nl m (r ,  ,  )  

l  

 ˆ  nl m (r ,  ,  )  ( 1)l  nl m ( r ,  ,  )
   nl ml ’s have the parity of l.
l l


In Dirac’s notation,  nl m (r )  n l ml
l

Eigenfunctions  nl ml ( r ) form an orthonormal basis as given by
*   
 n l m (r ) n lm (r ) d r   n n  ll   ml ml or equivalently n l ml n l  ml   n n  ll  m
l l l
ml

  
Once  nl ml ( r ) functions are known, the solutions to TDSE are given by  (r , t )   n l ml ( r ) ei En t  .

Orbital angular momentum and orbital magnetic Dipole moment of the electron
Consider a hydrogen atom and assume the electron orbit as circular. Since this acts as a current
loop, there is a magnetic field in the direction perpendicular to the plane of the orbit.


v

e
me

I Current


At large distances from the loop, the B field produced by the current loop is similar to that
produced by a magnetic dipole placed at the centre of the orbit.
 
The magnetic dipole moment due to a current loop enclosing a small area dA, M  IdA
107

M is in a direction normal to the loop.
e e ev
Current (magnitude) due to electron,  
T 2 r v 2 r

Area of the loop, dA   r 2


ev evr
Orbital magnetic dipole moment (magnitude) of the electron, M L    r2 
2 r 2
Orbital angular momentum (magnitude) of the electron, L  me v r

ML ev r 1 e
    constant
L 2 me v r 2me

ML ML e B
Usually, is written as,  gl  gl
L L 2me  
e
Orbital g-factor, gl = 1 Bohr magneton,  B   9.27  1024 J T 1
2 me

 L  B L  
ML  gl B  M L   gl ( M L is anti-parallel to L )
 
B
 M L z   gl Lz


Quantum mechanically, it can be shown that L  L  l  l  1  , where l = 0, 1, 2, ….. is the

orbital angular momentum quantum number.


Note: This L is the average value of the magnitude of the orbital angular momentum.

Depending on the value of l, we can characterize the orbital angular momentum states of electrons
in atom as s, p, d and f sub shells (states).

Quantum mechanically, it can be shown that Lz  ml  , Lx = Ly = 0.

Note: These Lx, Ly, and Lz are the average values of the components of orbital angular momentum.
ml is the orbital magnetic quantum number.

Quantum mechanically, it can be shown that, for a given value of l, ml varies in the range
 l  ml  l , in integer steps.  For a given values of l, there are (2l +1) number of ml
values.

Possible values of ml are  l,  l + 1,  l + 2, ……, 0, ..….. , l  2, l  1, l

For s- states, l=0  0  ml  0  ml = 0 one value

For p- states, l=1  1  ml  1  ml = 1, 0, 1 three values


108

For d- states, l=2  2  ml  2  ml = 2, 1, 0, 1, 2 five values

For f- states, l=3  3  ml  3  ml = 3, 2, 1, 0, 1, 2, 3 seven values

  L gl B l  l  1 
Quantum mechanically, M L  M L  gl B   g l  B l  l  1
 
 B Lz 
Quantum mechanically,  M L  z  g l  gl B  ml   gl  B ml
 

For a given value of l,  M L  z can have (2l + 1) number of discretely quantized values.

Note: These M L and  M L  z are average values of the respective quantities.

Vector model of orbital angular momentum



L  L  l  l  1  , where l = 0, 1, 2, …….

Lz  ml  , and for a given value of l,  l  ml  l .


Lx = Ly = 0

The above facts suggest that the orbital angular momentum L precess about the axis of
quantization (z-axis) so that its (2l + 1) # of projections on that axis are given by Lz  ml  (ml are

 l,  l + 1,  l + 2, ……, 0, ..….. , l  2, l  1, l). L may be viewed as lying on the surface of a

cone with altitude of ml  and z-axis as the axis of symmetry. All the orientations of L on the

surface of this cone are equally probable.


Eg: When l = 1

z Axis of quantization
 
ML L


 2
2 
 L
ML

  2
 
ML L

l = 1, L  2, Lz =   , 0, +  cos   1 2
Since q.m suggests the vector model of orbital angular momentum, it also suggests the vector
model of orbital magnetic moment.
109

 M L  z can have (2l + 1) # of discretely quantized values.


 
However, classically L (hence M L ) can have any # of orientations with respect to the z-axis. i.e.

 M L   M L z  M L

An atom possesses magnetic dipole moment due to the orbital motion of the electron. When an

atom is in an external magnetic field of strength B , the potential energy of the interaction
 
V   ML  B
       
The net force acting on the atom, F   V     M L  B 
 ML   B 
  
 B  B  B
Fx  M L  , Fy  M L  , Fz  M L  ,
x y z

If B is uniform, then Fx = Fy = Fz = 0.

Stern and Gerlach Experiment (1922)


This experiment was carried out to measure the magnetic dipole moments of atoms by detecting the
deflection of an atomic beam by an inhomogeneous magnetic field.

A collimated beam of Ag atoms entered the inhomogeneous magnetic field between the magnetic
poles and was detected after falling on a cooled collecting plate. (The whole setup was under
vacuum) Due to the geometry of the magnets, the force acting on the atoms was mainly in the z-
direction.
110

B B
Fz   M L  z  , - field gradient
z z
B
Along the x-axis, remains constant  Fz   M L  z .
z

With no With magnetic field on With magnetic field on


magnetic field (Classically expected) (Observed)

Since the field gradient is a maximum at the center, Fz is maximum at the center and the deflection
should be maximum. If the magnetic moments of the atoms are randomly oriented as expected
classically, every value of  M L  z such that  M L   M L  z  M L would occur and the deflected

beam would be spread into a continuous band (Fig. 2). Quantum mechanically only discretely
quantized values of  M L  z are possible (  M L  z  g l  B ml ) and hence the deflected beam would

B
be split into several discrete components. (According to Fz   M L  z  ). What Stern &
z
Gerlach observed was two distinct lines (Fig. 3). This is called the space quantization of the
components of the magnetic dipole moment along the direction defined by the magnetic field (i.e.
z-direction)
We know that,
l is an integer  # of possible ml values for a given l = (2l + 1 ) is an odd integer.
For any value of l, one of the ml values is zero.  There should be an undeflected
component of the Ag atomic beam (because ml = 0 result in a zero force)

Consider Ag atom
Ag (Z = 47): 1s2 2s2 2p 6 3s2 3p 6 4s2 3d10 4p6 5s1 4d 10

The net magnetic moment of an Ag atom is determined by the single 5s electron (all the other 46
electrons in completely filled sub shells contribute to nothing.)

For an s electron  l=0  ml = 0   M L z =0  Fz = 0


111

No deflection in the Ag beam is possible but, Stern & Gerlach observed two deflections!

An explanation came from Goudsmit and Uhlenbeck in 1925. i.e., in addition to the magnetic

moment produced by the orbital motion, electron may possess an intrinsic magnetic moment, M s ,

where the component of M s in a given direction (z-axis) can take only two values. It can be
 
 
postulated that M s is due to an intrinsic angular momentum S of the electron.

S - Spin angular momentum of the electron
 
Similar to the relationship of M L with L , we have

 B S
M S   gs , g s – Spin g-factor (spin gyromagnetic ratio), gs  2


Ms - Spin magnetic moment of the electron.

Unlike the orbital angular momentum the spin angular momentum does not correspond to any
   
physical motion. (i.e., it is not possible to write a relationship equivalent to L  r  p for S )

Quantum mechanically, S and S z are defined to have average values

S  S  s  s  1  , S z  ms 

For a given value of s,  s  ms  s 

(2s + 1) # of ms values (s,  s + 1, s + 2, ……. , s  2, s  1, s) are possible.

s – Spin angular momentum quantum number ms – Spin magnetic quantum number

Average values of S x  0 and S y  0 .

1 1
For electrons, s  12   12  ms  1
2
 ms   and 
2 2

3  
S   and Sz  and 
2 2 2
For other subatomic particles such as protons, neutrons, s  1. They are commonly called ‘spin-
2

1 particles’ or ‘fermions’.
2

For photons, s = 1  1  ms  1  ms = 1, 0, +1

S  2 and S z   , 0 ,  

Quantum mechanically defined values of S , S x , S y and S z suggest the vector model of spin

angular momentum.
112

1
Eg: s 
2

z Axis of quantization
 
MS S “Spin up”


2  S= 3 2

S= 3 2
 2

  “Spin down”
MS S


Since S z values are quantized   M s  z must also be quantized.
 B Sz 
 M S z   gs

  g s B  ms    g s  B ms   2   B   12 =   B

 
The force responsible for splitting the silver atomic beam in S & G experiment,

B B
Fz  M s   z  z   B
z

Magnetic moment of the Ag atom


 47  47     B s ( s  1)

M   ML   MS i  ML  i  5s   M S 5s  0  gs

i 1 i 1

s - electron  l = 0  ml = 0 s= 1  ms =  12
2

ms = 12 ms =  12

ml = 0

p - electrons  l=1  ml = 1, 0, +1 s= 1  ms =  12


2

ml = 1 ml = 0 ml = 1
113

This explains why it is possible to accommodate 2 electrons in an s-subshell and 6 electrons in a p-subshell,
etc.
Electron beams which have equal #s of ‘Spin up’ and ‘Spin down’ electrons are called ‘unpolarized
s
electron beams’ and those which have unequal # of ‘Spin up’ and ‘Spin down’ electrons are called
‘polarized electron beams’.

N  N
The ‘degree of polarization’ = P 
N  N

N  and N  are the #s of spin up and spin down electrons in the beam.
For an unpolarized electron beam, N  N  P0

For a polarized electron beam, N  N  P0

For a perfectly polarized beam, N  = 0  or N   0   P 1 0  P  1

S – G magnets are sometimes called spin filters.

Tutorial 5
TISE in Spherical Polar Co-ordinates, Orbital Angular momentum, Hydrogenic Atoms

1. (i) If  n l m l ( r ) describes the energy eigenfunction of the electron in the nth state of a hydrogen atom,
describe the quantum numbers which characterize energy eigenstates and give their limits. Sketch
energy levels and indicate possible eigenstates (using the above notation) up to the 2nd excited level.

(ii)Explain briefly why the sub states (shells) 1p, 1d, 1f, 2d, 2f, 3f, … etc, can not occur in atoms.

(iii) Explain briefly why it is not possible to accommodate more than 02 electrons in a s-sub state, 06
electrons in a p-sub state, 10 electrons in a d-sub state, 14 electrons in a f-sub state, in an atom.

2. Find the energy eigenfunctions and the corresponding energy eigenvalues of a particle of mass m, that
rotates on a circle of radius R in a potential V(r), where V(R) = V0.
 2 1     1 2 

In plane polar co-ordinates (r ,  ),   r  
 r r  r  r 2  2 

3. At a given instant of time, a quantum system is in the normalized orbital angular momentum state
 ( ,  )  N sin  sin  , where N is the normalization constant.

(i) Determine the constant N.


(ii) What possible values of Lz will be found by measurement and with what probability will these
values be occurred ?

(iii) What is L2 for this state ?

(iv) What is Lz for this state ?


3
Some useful spherical harmonics are Y1 1 ( ,  )  
sin  e  i .
8
4. Consider an electron in a hydrogen atom whose wave function at t = 0 is the following superposition of
  1   
energy eigenfunctions  n l ml ( r ) :  (r , t  0) =  2 10 0 (r )  3 2 00 ( r )   32 2 ( r ) 
14  
114

(i) Is the wave function an eigenfunction of the parity operator ?


(ii) What is the probability of finding the system in the ground state (100) ? In the state (200) ? In the
state (322) ? In another energy eigenstate ?

(iii) What is the expectation value of the energy (in terms of En ): of the operator L2 ; of the operator
Lz ?

5. Consider a hydrogen-like atom with atomic number Z, containing an electron in the ground state (1s).
 
The wave function of this electron is  1s (r )  N exp  Z r a , where N is the normalization
constant, r is the radial co-ordinate and a is the modified Bohr radius.
(i) Determine the constant N.
(ii) Determine the average value of the potential energy of the electron.
2 2
(iii) What is meant by the quantities  1s and 4 r 2  1s ?
(iv) Evaluate the average value of the distance to the electron from the origin.
(v) Evaluate the distance from the origin at which the probability of finding the electron is a
maximum.
(vi) Comment on your results for parts (iv) and (v).
(vii) Determine the classically forbidden region for the electron in the ground state of a hydrogen
atom, if its bound-state energy eigenvalue is 13.6 eV.
(viii) What is the probability of finding the ground state electron in this region ?

n a y n!
It is given that, I  y e dy  , where Real( a )  0 and n is a positive integer.
0 a n1

6. A system is found to be in an orbital angular momentum state which is a mixture of orthonormal basis
states corresponding to l = 1, as given by    1 1   1 0   1  1 , where ,  and 
are real numbers.
(i) What is the condition that needs to be satisfied by ,  and , for the normalization of  ?
ˆ
(ii) Find the expectation values of the operators Lˆ z and L2 for the state  of the system.

Electron's Spin Angular Momentum, Spin-Orbit Interaction


7. With regard to a Stern-Gerlach experiment, in each of the following cases, illustrate the recorded
configuration of the atomic beam that would be seen on the on the detecting plate giving reasons. (a) If the
magnetic field was turned off (b) If the magnetic field was homogeneous (c) If the magnetic field was
inhomogeneous and the space quantization did not exist (d) If the magnetic field was inhomogeneous and
the space quantization did exist

8. A beam of hydrogen atoms, emitted from an oven running at a temperature 400 K, is sent through a Stern-
Gerlach magnet of length 1 m. The atoms experience a magnetic field with a gradient of 10 T/m. Calculate
the transverse deflection of a typical atom in each component of the beam, due to the force exerted on its
spin magnetic dipole moment, at the point where the beam leaves the magnet.

9. Consider a beam of particles of mass M and spin ½ , propagating in the +x-direction. The beam has a
cross section d. It interacts with an Stern-Gerlach apparatus whose field is in the z-direction.
Employing relevant uncertainty relations, show that the smallest uncertainty in the normal displacement
(z) grows large with decreasing mass.
Prepared by: Revised in September 2007

Dr. WMKP Wijayaratna


Department of Physics
University of Colombo,
Colombo 03.
115

You might also like