0% found this document useful (0 votes)
20 views34 pages

Quantum Mechanics

Uploaded by

amit tiwari
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)
20 views34 pages

Quantum Mechanics

Uploaded by

amit tiwari
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/ 34

CHEM0019: PHYSICAL CHEMISTRY

Quantum Mechanics
Prof. Sarah (Sally) L Price

[email protected], Room G18A Christopher Ingold Lab


Academic Support hours Tue 10-11, Thur 12-1.
Section Topics covered
1 Overview
2 Postulates of quantum mechanics: probability interpretation of Ψ, operators and Hamiltonians,
eigenvalue equations, Schrödinger equation, expectation values, Variation principle
3 Exact solutions to the Schrödinger equation: harmonic oscillator, particle on a ring, particle on a sphere
(rigid rotor), hydrogen atom
4 Beyond exact solutions: helium atom
5 Molecular orbitals: Born-Oppenheimer approximation, linear combination of atomic orbitals, secular
equations, two-orbital systems
6 Hückel theory for π-electron systems; allyl radical
7 Concluding remarks
Recommended textbooks:
“Quantum Mechanics for Chemists” by D. O. Hayward, RSC.
“Molecular Quantum Mechanics” by P.W Atkins and R.S. Friedman, 5th edition, OUP, available on Explore
“Physical Chemistry” by P.W Atkins and J. de Paula, 8th edition, OUP.

2
1. Overview
The principal aim of this course is to understand wavefunctions, how to calculate them and how to use them in
chemistry. To do this we will first define some fundamental principles of quantum mechanics. To help explain
these principles we will use test problems, such as the harmonic oscillator and the rigid rotor. Understanding
these cases allows us to understand the electronic structure of the hydrogen atom. We will then move on to
consider the electronic structure of helium atom, and an introduction to molecular orbital theory. This helps
illustrates why approximate wavefunctions can be used both qualitatively and quantitatively throughout
chemistry.

“The fundamental laws necessary for the mathematical treatment of a large part of physics and the whole of
chemistry are thus completely known, and the difficulty lies only in the fact that application of these laws leads to
equations that are too complex to be solved”

Dirac 1929

3
2. Postulates of quantum mechanics
Postulate = a stipulation, an assumption, a fundamental principle

2.1 Probability interpretation of Ψ

Postulate 1: the state of a system is fully described by a mathematical function Ψ, called the wavefunction.

As an example, consider a particle which can move only in the x direction

This particle is described by a wavefunction (x).

Question: what is the probability that the particle will be found between points x and x + dx?

Answer:   x  dx  ( x) ( x)dx
2 *

This probabilistic interpretation of the wavefunction is commonly known as the Born interpretation.

Notes: 1. In this treatment we assume that the wavefunction is normalized, i.e. that the probability of finding the

particle somewhere along the x direction is 1. Mathematically we write:    x    x  dx     1 where
*
is


the “bra-ket” notation first introduced by Paul Dirac.

4
2. The wavefunction itself has no physical meaning. It may, at any given point in space, be positive or negative,
real or complex.

  x   *  x    x  is always non-negative and real, and has the probabilistic meaning described above.
2

2.2 Operators and Hamiltonians


An observable is a measurable property such as bond length, dipole moment, kinetic energy etc….

Postulate 2: every observable B is represented by an operator, and all operators can be built from the operators
for position and momentum.

The operator for position in the x direction is defined as Bˆ  xˆ

d
and for momentum in the x direction, Bˆ  pˆ x  i
dx

h
where i  1 and   1.05457160 x 1034 J s.
2

A very important operator is the total energy operator, which is called the Hamiltonian

Hˆ = Tˆ  Vˆ

where Ĥ = total energy operator, Tˆ = kinetic energy operator and Vˆ = potential energy operator.

5
1 2 (mv)2
We know that T = mv =
2 2m

For motion in the x direction px = mvx and hence

2
ˆ pˆ x2 1  d  2
d2
T   i  
2m 2m  dx  2m dx 2

The role of the operator is to operate on a wavefunction to yield information associated with the
observable that the operator represents.

2.3 Eigenvalue equations

An eigenvalue equation has the general form: B̂f  bf where the operator B̂ acts on the eigenfunction to
regenerate f multiplied by the eigenvalue b (a constant).

d ˆ  d eax  aeax  af
Example 1: If Bˆ = and f  eax , Bf
dx dx

d
and hence eax is an eigenfunction of Bˆ  with eigenvalue a .
dx

d ˆ  d sin  ax   a cos  ax   af d
Example 2: If Bˆ = and f  sin ax , Bf Hence sin  ax  is not an eigenfunction of .
dx dx dx

6
2.4 The Schrödinger equation

Let us now use the operator Ĥ to operate on Ψ(x):


2
d2
Hˆ   x   Tˆ  x    x   Vˆ  x    x      x V  x  x
2m dx 2

where Vˆ ( x) is the potential energy operator for motion in the x direction.

If Ψ(x) is an eigenfunction of Ĥ , then it is termed an exact wavefunction, and


2
d2
Hˆ   x      x   V  x    x   E  x 
2m dx 2

where the eigenvalue E is an energy (recall that Ĥ is an energy operator). This equation is the Schrödinger
equation for a particle moving in the x direction. (If Vˆ  x  is complicated, we may not be able to find an exact

  x  analytically and so may only be able to find approximate wavefunctions using a computer.)

There may very well be many more wavefunctions which are eigenfunctions of Ĥ , i.e. which are solutions to
Hˆ ( x)  E ( x) . Any two non-degenerate solutions (i.e. solutions of different energy) are orthogonal

   x    x  dx   j k  0
*
j k


7
Any two wavefunctions are orthonormal if

 j  k   jk

where  jk is the Kronecker delta.

 jk  1 if j is the same as k i.e. j  k (the wavefunction is normalised) and

 jk  0 if j is different from k i.e. j  k (the wavefunctions are orthogonal)

2.5 Expectation values

The expectation value of an operator B̂ for a wavefunction  is denoted B and is defined as

  Bˆ d
*
 Bˆ 
B  
    
*
d

where d tells us that the integration is being performed over all space. If the wavefunction  is normalised to 1,
then the expectation value simplifies to

B   Bˆ    * Bˆ d

8
If  is an exact wavefunction, i.e. it is an eigenfunction of Ĥ , and it is also an eigenfunction of B̂ such that
B̂  b , then

B   Bˆ    * Bˆ d   *bd  b  *d  b .

What would happen if we attempted to measure the value of observable B? Because  is an exact
wavefunction and an eigenfunction of B̂ , a series of identical experiments will all yield b i.e. they will give
a single value.

Let us now consider the specific and real case of an atom or molecule with more than one electron. This system
has a set of wavefunctions Ψn which satisfy the Schrödinger equation Hˆ  n  E n . However, these

wavefunctions are unlikely to be eigenfunctions of of B̂ i.e. Bˆ n  bn n . Under these circumstances a series of
identical experiments would not all yield the same answer but each individual experiment could give any one of
the eigenvalues b0, b1, b2…. and the expectation value of B̂ is the average of all the values that would be
obtained from a large number of experiments.

This is represented by our final postulate 3 and its implication 3’:

Postulate 3: when a system is described by a wavefunction , the average value of the observable B in a series

of measurements is equal to the expectation value of the corresponding operator B̂ , i.e. B̂ .

9
Postulate 3': when Ψ is an eigenfunction of B̂ , determination of B always yields one result, b. When Ψ is not an
eigenfunction of B̂ , a single measurement of B yields a single outcome which is one of the eigenvalues of B̂ ,

and a large number of measurements will yield an average of the eigenvalues of B̂ i.e. B̂ .

2.6 The Variation Principle


Postulate 3 leads to a plausibility argument for the variation principle, which is used extensively for the
computational treatment of many-electron atoms and molecules, and is the foundation for the approximate
methods for determining molecular orbitals in Section 5.

We want to find the energy of the ground state  0 , which has an associated energy E0 of a molecule which has a

set of wavefunctions  n which satisfy the Schrödinger equation Hˆ  n  En  n . As the exact wavefunctions for
many-electron systems are not known and we must employ an approximate wavefunction  trial . As  trial is not

exact, it will not be an eigenfunction of Ĥ , and Hˆ trial  Etrial .

If we attempted to measure the energy of a hypothetical system described by  trial , a series of identical
experiments would not all yield the same answer but each individual experiment could give any one of the
eigenvalues E0 , E1 , E2 …. and the expectation value of Ĥ , i.e. Etrial   trial *Hˆ trial d , is the average of all the

values that would be obtained from a large number of experiments.

10
The variation theorem states that for any trial wavefunction  trial , the expectation value of the energy can never
be less than the true ground state energy E0:

*Hˆ  trial d  trial Hˆ  trial





trial
E0
 trial * trial d
 trial  trial

(If  trial is normalised, then E0   trial Hˆ  trial ).

The expectation value of the energy is an average of the true energies of the system E0 , E1 , E2 …., and this can
never be less than E0 .

11
3. Exact solutions to the Schrödinger equation
One of the most important exact solutions is that for the particle in a box with V ( x)  0 . This was covered in
Basic Physical Chemistry (CHEM0009) and you should make sure you understand it.

3.1 Harmonic oscillator


Harmonic oscillations occur when a system experiences a restoring force proportional to the displacement from
equilibrium, e.g. pendulums, vibrating springs. Consider a one-dimensional harmonic oscillator,

F
0 x

The restoring force is F  kx , where k , the constant of proportionality, is called the force constant. This is
Hooke’s Law.

dV
Force = negative gradient of potential energy, i.e. . F   . Hence
dx

kx 2 ˆ kx 2
V    Fdx    (kx)dx  k  xdx  and V 
2 2

12
We now combine this operator for potential energy with that for kinetic energy (see section 2.2) to obtain the
2
d 2 kx 2
Hamiltonian: Hˆ  Tˆ  Vˆ   
2m dx 2 2
2
d 2  kx 2 
and hence obtain the Schrödinger equation: Hˆ      E
2m dx 2 2

Note that the we seek a wavefunction such that the total energy E is constant as it must be an eigenvalue to be
an exact solution.

We can find a solution by begining with the trial wavefunction

  Ae x
2

d 2  x 2
where A is a normalisation constant and  is another constant. Since  ( 2  4 2 2
x ) Ae
dx 2

kx 2 Ae x
2
2
we can subsitute into the Schrödinger equation to give Hˆ    (2  4 x ) Ae  x 2
  EAe x
2
2 2

2m 2

 2 1
( )(2  4 2 x 2 )  kx 2  E
2m 2

13
We know that the total energy E is constant. This can be the case only if E is independent of x, which will be true
 2 1 km
2
only if the two terms in x cancel, i.e. if ( )(4 2 x 2 )   kx 2 . This will be true if  
2m 2 2
2
 2
km k
We can now substitute for  into the equation for the energy: E   
m m2 2 m

 k
Hence   Ae x is an exact wavefunction with energy E  where  
2
.
2 m

This wavefunction   Ae x is the first in a series which satisfy the Schrödinger equation for the harmonic
2

oscillator, with general formula:  v  A H v  x  e x where Hv is a Hermite polynomial.


2

H 0  x   1, H1  x   x, H 2  x    4 x 2  1 , etc

14
3.1.1 The harmonic oscillator wavefunctions

The wavefunctions and associated energies for the harmonic oscillator are summarised in the following table:

Wavefunction Energy


 0  A0e x
2
E
2

3 
 1  A1 xe x
2
E
2

5 
 2  A2 (4ax 2  1)e x
2
E
2

 v  Av H v ( x)e x  1
2
E  v   
 2

Let us now examine the first four wavefunctions:

15
V=1/2kx2

v=3
v=2
v=1
v=0

16
Notes:

1) The eigenvalue of the ground state E0 ≠ 0. This is called the zero-point energy.

2) The energies have even spacing, E   . If the force constant k increases,  increases and hence so
does E . This will also be the case if the mass decreases.

3) In classical mechanics, the maximum extension of a vibrating string/bond, or of a pendulum, occurs at the
1 2  1 k
point where the total energy equals the potential energy i.e. when kxmax  E  V   v   . Thus the
2  2 m
edges of the potential well correspond to xmax , the classical limit.

4) The wavefunctions look rather similar to those of the particle in a box, although the wavefunctions for the
harmonic oscillator extend beyond the edges of the potential well. Thus the quantum oscillator is tunnelling to
regions of negative kinetic energy.

5) As  increases,  v2 increasingly peaks towards the edges of the well i.e. the classical limit.

17
3.1.2 Application to spectroscopy
The potential between two atoms in a diatomic molecule can be represented schematically by the plot of V(x)
against x below. The well is very steep for small x due to the large replusion between the nuclei, and the
potential tends to zero at large x as the bond weakens then breaks.

De

x=xe

The harmonic oscillator potential, shown by V(2)(x) on the plot, is a good approximation to V(x) around the
equilibrium (most stable) internuclear separation xe ,but clearly not away from the equilibrium region.

18
The simple harmonic oscillator is used as a first approximation, and is most accurate for the ground state energy
(  0 ), the zero-point energy.

One way to treat diatomic molecules more realistically than with the harmonic oscillator approach is to use the
Morse potential


V  x   hcDe 1  e  k
2
 a x  xe 
with a  , and De is the depth of the minimum of the potential curve (i.e. it is the
2hcDe

dissociation energy of the bond, ignoring the vibrational energy). The solution to the Schrödinger equation with
2
 1  1
the Morse potential gives the energy levels Ev           e , with   0,1,2,3.....
 2  2

k a2
where   , e  is the anharmonicity constant and  is the reduced mass.
 2

mAmB
For the diatomic molecule AB,   .
mA  mB

The term containing  e in the expression for E above reduces the energy from the harmonic oscillator value
and becomes increasingly important as  becomes large. Thus the vibrational energy level separation is not
constant as for the harmonic oscillator, but converges as  gets larger. The molecule dissociates as    .

19
3.2 Particle on a ring – circular motion in a fixed plane
We now consider the case of a particle of mass m moving at a constant velocity v around a circle of radius r in
the xy plane:
y

r v

x
m

The potential energy is constant (Vˆ  0 ) and the Hamiltonian for this motion is just the kinetic energy part:
2
 2 2  2
 2 1  1 2 
Hˆ         
2m  x 2 y 2  2m  r 2 r r r 2  2 

(using the tranformation from Cartesian to polar coordinates  x, y    r ,  ).

20
Because the particle is confined to move at a fixed radius, the best way to proceed is to use polar coordinates.
Since r is a constant, any term containing a derivative of r will vanish and so the Hamiltonian reduces to
2
2
Hˆ   .
2mr 2  2

The moment of inertia, I of the particle about the z axis is given by I  mr 2 and hence the Schrödinger equation
2
2 2
2
becomes: Hˆ               E  
2mr 2  2 2 I  2

2
ml2
The solutions to this equation take the form:     Neiml with E  where ml is a constant (a quantum
2I
number). We can expand this wavefunction in the form:     N  cos ml  i sin ml 

from which we can see that  ( ) consists of two sinusoidal waves, one real and one complex (imaginary).

We must now establish the boundary conditions (c.f. for the particle in a box, the wavefunction has to go to
zero at the ends of the box). In the present case, in order for the wavefunction to be single-
valued    2      .

The boundary condition is thus cos ml  cos ml   2   cos  ml  ml 2 

and sin ml  sin ml   2   sin  ml  ml 2  which requires ml to be equal to zero or an integer.

21
Acceptable wavefunctions are therefore:

 ( )  Neim  ml  0, 1, 2....


l

where the  ml represent anti-clockwise and clockwise motion of the particle around the ring. When ml  0 , the
particle is stationary.

Plots of the real (cosine) parts of  ( ) for ml  0, 1, 2 are shown; the imaginary
(sine) parts are similar but rotated through 90o.

Next we need to find the normalisation constant N . The probability of finding the
particle between  and   d around the ring is given by  ( ) * ( ) d .The
particle must be somewhere on the ring, and hence:
2 2 2

       d  1   Ne  d  N
iml  iml
d  N 2
* 2 2
Ne
0 0 0

The solution to the particle on a ring is therefore:


2
1  iml ml2
    e ml  0, 1, 2, 3,....... with E  .
2 2I

22
Interpretation Notes:

d
Recall that the operator for momentum in the x direction (linear momentum) is pˆ x = i . We can rewrite Ĥ for
dx
the particle on a ring in terms of an analogous operator for angular momentum about the z axis:

L2 d
Hˆ  z where Lz  i
2I d

What are the eigenvalues of this operator?

Lz     i
d
d
  im 

Ne l   i N  iml  e l    ml   
 im 

The z component of the angular momentum of a particle moving at constant velocity in a circle, Lz , can
have only certain values which are multiples of . The angular momentum is quantized.

L2
Classical mechanical treatment of the particle on the ring yields E  , where L is the rotational angular
2I
momentum. Equating (quantum mechanical) Lz with (classical mechanical) L gives the quantum mechanical
2
ml2
energy levels of the particle on a ring as E  ml  0, 1, 2, 3....
2I

23
3.3 Particle on a sphere – rotation in 3 dimensions
Now consider the case of a particle of mass m which is free to move anywhere on the surface of a sphere, i.e. at
a fixed distance r from an origin. As with the particle on a ring, the potential energy is constant and can be
neglected (Vˆ  0 ). This model can be applied to the rotation of rigid molecules and also provides the angular part
of the hydrogenic atomic orbitals.

The Hamiltonian for the kinetic energy in three-dimensions is

2
 2
 2
 2
 2
Hˆ =   2  2  2  2
2m  x y z  2m

where  2 is known as the Laplacian operator.

The restriction of the particle to fixed r means that it is best to tackle this problem by transforming to spherical
polar coordinates. The relationship between  x, y, z  and  r , ,  is:

z  r cos
x  r sin  cos 
y  r sin  sin 

24
and the Hamiltonian in spherical polar coordinates is:

1   2  
2
1     1 2 
Hˆ =  r   sin  
2m  r 2 r  r  r 2 sin      r 2 sin 2   2 

1  2  
This can be simplified a little, since r is a constant, the r  term vanishes, and we can write Ĥ as
r 2 r  r 

2 2
 1     1 2 
Hˆ    2
where Hˆ =  2   sin   2 2
2m 2mr  sin      sin   

2 is called the Legendrian. Remembering that I  mr 2 , the Schrödinger equation becomes


2
  2   ,   E  ,  .
2I

The solutions to this equation take the form:   ,         where Θ is a function only of the variable , and

 is a function only of . The solutions  ( , ) are better known as the spherical harmonics, given the symbol
Yl ,ml ( , ) where l and ml are the quantum numbers controlling the form of the wavefunctions. i.e.

 ( , )   ( ) ( )  Yl ,m ( , ) . We have already met the function  ( ) ; it is identical to the one for rotational
l

motion in a fixed plane:  ( )  Neiml ml  0, 1, 2....

25
The form of the solutions for  ( ) depend on the values of l and ml . The allowed values of l and ml are

ml  l , l  1, l  2,......., l  2, l  1, l . The spherical harmonics have the property  2Yl ,ml ( , )  l (l  1)Yl ,ml ( , ) .

Hence subsitituting into the Schrödinger equation for the particle on a sphere:
2 2
Hˆ   ,     Yl ,ml  ,  
2
l (l  1)Yl ,ml  ,   El ,ml Yl ,ml  ,   El ,ml   , 
2I 2I
2
and so El ,ml  l  l  1 where l  0,1,2,3....
2I

The energy is thus independent of ml and hence for every value of l there are  2l  1 states with the same
energy (degenerate). The first nine spherical harmonics, ignoring their normalisation constants:

l ml Yl ,ml ( , ) l ml Yl ,ml ( , )

0 0 real 2 0 3cos2   1 real


constant

1 0 cos real 2 +1 cos sin  ei

1 +1 sin  ei
2 -1 cos sin  ei

1 -1 sin  ei
2 +2 sin 2  e2i

2 -2 sin 2  e2i

26
An important property of degenerate wavefunctions is that any linear combination of them will have the exact
same energy, and be a solution to the corresponding Schrödinger equation. Consider the expressions for
Yl ,ml ( , ) in the table above that are in degenerate pairs with e iml . Linear combinations of the (not normalised)

spherical harmonics Y1,1 ( , ) and Y1,1 ( , ) are

Y1,1  ,   Y1,1  ,   sin   ei  ei   2sin  cos

Y1,1  ,   Y1,1  ,   sin   ei  ei   2i sin  sin 

The angular component of x in spherical polar coordinates is sin  cos and for y it is sin  sin  . Hence it is usual
to call the linear combinations Re(Y1,1  Y1,1 )  Y1, x and Im(Y1,1  Y1,1 )  Y1, y

Note that they are both all real or all imaginary in the angular dependence. Furthermore, the linear combinations
have the same energy as the original Y1,1 ( , ) and Y1,1 ( , ) .

The spherical harmonics Yl ,ml ( , ) can be visualised in different ways.

27
Polar plots, in which the value of the wavefunction is indicated by
distance from the origin, of the real wavefunctions for l = 0 and l = 1.

28
3.4 Rigid rotor – the rotation of diatomic molecules
Consider a diatomic molecule AB. The atoms rotate about the centre of mass, and the motion of the masses mA
and mB about the centre of mass is mathematically equivalent to the rotation of a single particle of reduced mass
mAmB
 about a fixed point:
mA  mB

The distance between the fixed point


and the particle is equal to the
bondlength r . We can therefore treat
the rotation of a diatomic molecule as
if it were the motion of a particle of
mass  on the surface of a sphere.

Earlier we saw that the energies of


the wavefunctions for the particle on
2
a sphere are E  l (l  1) .
2I

In rotational spectroscopy we normally use J as the symbol for the quantum number, and so the allowed
2
rotational energies for a diatomic molecule are: E  J ( J  1) where J  0,1,2,3,.... and I   r 2 .
2I

29
3.5 The hydrogen atom
The H atom consists of a nucleus with a single proton, and an orbiting electron. We can approach it as an
extension to the particle on a sphere/rigid rotor to the situation where neither r nor the potential V  r  is constant.

The nuclear charge Z is +1 for the hydrogen atom, but we will consider the general case of hydrogenic atoms,
so that the solutions can be extended to He+ and Li2+ etc. The potential energy of the electron in the electrostatic
field generated by the nucleus is given by Coulomb’s law:

Ze2
V (r )   Z r
4 0 r  
Nucleus electron

where e is the charge on the electron and  0 is the vaccuum permittivity.

To simplify matters, quantum mechanics usually work in atomic units. In atomic units, the charge on the
electron e  1 and its mass me  1, and also  1 and 4 0  1. This is consistent with the atomic unit of length, the

4 0 2
Bohr radius a0  2  1, where in SI units a0 =5.29177208x10-11 m.
e me

Thus, in atomic units the potential energy operator Vˆ  r   Z / r .

30
Now, since the potential V is independent of the angles  and  , the part of the Hamiltonian that depends on
these angles will be the same as for the rigid rotor. The kinetic energy operator only differs in that r can now
vary. Thus the full Hamiltonian for a hydrogenic atom is:

ˆ 1 1   2   1     1 2  Z
H =   2 r   sin   
2  r r  r  r 2 sin      r 2 sin 2   2  r
For the particle on a sphere, solution wavefunctions are of the form  ( , )   ( ) ( ) i.e. products of

independent functions of  and  . Solutions to the H atom can be found in an analogous manner in the form
  r , ,   R(r )( )( ) , where R  r  is the radial part of the wavefunction and  ( ) ( ) are the spherical

harmonics,  ( ) ( )  Yl ,ml ( , ) . (See document “Mathematical details for hydrogenic atom” for more information

on this relationship between the particle on a ring, on a sphere and the hydrogenic atom.)

The solutions for the hydrogenic atom are Rn,l  r   Ln,l  r  r l e Zr / n where Ln ,l are the Laguerre polynomials and r is

4 0 2
Zr

in atomic units. The exponential decay of the wavefunction is e na0
where a0  ,the Bohr radius. The
mee2

solutions for Rn,l  r  add a third quantum number n , which takes values 1, 2, 3,…. The quantum number l has a

maximum value of (n  1).

31
The final solutions to the Schrödinger equation for the H atom are thus:

mee4 Z2
  r , ,   Rn,l  r Yl ,ml  ,  E  Z 2
  2  Eh 
32n2 2 02 2
2n

where the second expression for the energy is in atomic units. The atomic unit of energy is also called a hartree,
(1 Eh =4.3597438 x 10-18 J=27.211383 eV).

Note that the energies depend only on the quantum number n . This is true only for the H atom, or any other
atomic ion with only one electron. The hydrogenic atomic orbitals are summarised up to n  3 and the radial forms
plotted below, in terms of the dimensionless parameter   Zr / ao . Note that because the orbitals are
degenerate, we can take linear combinations of orbitals to give real atomic orbitals, but this makes the 3d orbitals
look less equivalent. The angular parts of these wavefunctions are the spherical harmonics discussed in 3.3.

The radial parts are shown in the following table of hydrogenic atomic orbitals up to n  3 expressed in terms of
  Zr / ao . The normalisation constants are not included.

32
n l ml Rn,l (r ) Yl ,ml ( , ) Orbital type

1 0 0 e  Constant 1s Unnormalised hydrogenic atom

2 0 0  Constant 2s orbitals to n=3


(2   )e 2
  Zr / ao .
2 1 0 e
 cos 2p0 2pz
  r , ,   Rn,l  r Yl ,ml  , 
2


2 1 +1 e 2 sin  ei 2p+1
Z2
2px and 2py E   2  Eh 

2 1 -1 e 2 sin  ei 2p-1 2n

3 0 0  Constant 3s
(27  18  2 2 )e 3

3 1 0 (6   ) e

3 cos 3p0 3pz


3 1 +1 (6   ) e 3 sin  ei 3p+1
3px and 3py

3 1 -1 (6   ) e 3 sin  ei 3p-1


3 2 0  2e 3 3cos2   1 3d0 3dz2


3 2 +1  2e 3 cos sin  ei 3d+1


3dxz and 3dyz
3 2 -1  2e 3 cos sin  ei 3d-1


3 2 +2  2e 3 sin 2  e2i 3d+2
 3dx2-y2 and 3dxy
3 2 -2 2  3
 e sin 2  e2i 3d-2
The number of radial nodes for a given orbital is n  l  1
. Note that the plots below show just the radial
dependence of  (left) and * (right) which is
modified by the angular terms. At a given value of r the
volume of a shell of thickness dr is 4 r 2dr , which must
be included in probability and expectation value
calculations

34

You might also like