Sei sulla pagina 1di 70



PETER G. NELSON

AN INTRODUCTION
TO THE QUANTUM
THEORY FOR CHEMISTS
AN EXPERIMENTAL APPROACH

2
An Introduction to the Quantum Theory for Chemists: An Experimental Approach
1st edition
© 2019 Peter G. Nelson & bookboon.com
ISBN 978-87-403-2732-8

3
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Contents

CONTENTS
Acknowledgements 6

1 Introduction 7

2 Background 8

3 Emission spectra 9

4 Energies of atoms 11

5 Wave-like behaviour 16

6 Wave mechanics 19

7 Schrödinger’s theory of hydrogen atom 26

8 Dirac’s theory of hydrogen atom 30

9 Modified Schrödinger theory of hydrogen atom 34

10 Modified Schrödinger theory of other atoms 36

Free eBook on
Learning & Development
By the Chief Learning Officer of McKinsey

Download Now

4
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Contents

11 Modified Schrödinger theory of molecules 46

12 Valence-bond (VB) theory 56

13 Ligand-field theory 60

14 Density functional theory (DFT) 62

15 Perspective on the quantum theory 64

16 Further reading 65

Questions 66

Answers 68

Appendix 69

5
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Acknowledgements

ACKNOWLEDGEMENTS
I am very grateful to Dr. David Johnson for reading through the text and commenting
extensively on it. His suggestions have led me to make considerable improvements.

I have taken illustrations from the internet with my thanks. As far as I know, they are all
free to copy, but if any are not, I apologize.

6
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Introduction

1 INTRODUCTION
This little book is a brief introduction to the quantum theory for chemists. In it I keep as
close to experiment as possible, and to ideas that are familiar to chemists. This approach
takes me away from one based on mathematics. There is some mathematics, but I have
kept this as simple as possible.

The quantum theory is used in chemistry in two main ways. One is to make accurate
calculations of the energies of chemical processes. These calculations are very complicated
but are made possible by modern computers. Algorithms are now available for chemists to
use. An example of such a calculation is that carried out on the sodium‒helium system at
various pressures leading to the prediction that Na2He would be stable above 1.6 × 106 bar,
a prediction verified by experiment.

Chemists also use the results of approximate calculations to understand chemical phenomena
and make tentative predictions. It is possible, for example, to explain the Periodic Table on
the quantum theory, as we shall see in a later section.

In this book, we shall keep the two uses of the quantum theory in chemistry in mind. We
shall gradually build up the principles behind accurate calculations while also discussing
some of the results from more approximate treatments.

7
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Background

2 BACKGROUND
Physicists developed the quantum theory in the early decades of the 20th century to
explain phenomena that could not be explained by classical physics. These included the
discovery that atoms comprise a positively charged nucleus surrounded by electrons (Fig.
2.1). According to classical physics, an electron is attracted to a nucleus by a force that is
inversely proportional to the square of the distance (Coulomb’s law) and should progressively
lose energy and collapse on to the nucleus. That this does not happen means that, on an
atomic scale, either Coulomb’s law breaks down (favoured by G.N. Lewis in his famous
paper on atoms and molecules), or classical laws of motion break down, or both.

Figure 2.1 Simple model of an atom


[David Johnson]

In Figure 2.1, the number of electrons (charge െ݁) is equal to the atomic number (ܼ), and
the nuclear charge to ൅ܼ݁ . The simplest atom is that of hydrogen (ܼ ൌ ͳ)

8
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Emission spectra

3 EMISSION SPECTRA
A phenomenon that helps to answer the question of what breaks down is the emission of
light from a gas or vapour through which an electric discharge is passing (Figs. 3.1 and
3.2). This light can be analysed by passing it through a prism or grating in a spectrometer.
A grating consists of a transparent material into which a large number of uniformly spaced
wires have been embedded (about 300 per millimetre). A prism or grating bends light to
different degrees according to its wavelength and splits it into its component colours. What
is found is that the light is made up of narrow ranges of colour (“lines”) with dark spaces
between them. This is shown in Figure 3.3 for a selection of elements. The spectra extend
to the left into the ultraviolet and to the right into the infrared.

Figure 3.1 Apparatus for studying the passage of


electricity through a gas [Shiksha Services]

Figure 3.2 Discharge of electricity through hydrogen


[Wikimedia Commons]

9
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Emission spectra

Figure 3.3 Emission spectra of some elements compared with the spectrum
of white light (1) [cnx.org]


The scale in Figure 3.3 is of wavelength ( � )=in ångström units (Å = 10‒10 m). The

corresponding frequency is given by

� = � (3.1)

where � is the speed of light (2.998 × 108 m s‒1). Wavelength is determined by passing
�= �
light through a grating of known spacing (݀) and measuring the angle (ߠ) at which it is
diffracted. Diffracted rays from adjacent slits reinforce each other when the extra distance

one has to travel (݀•‹ߠ) is equal to � =(Fig. 3.4). The emission spectra of metals are

obtained at elevated temperatures.


d θ

� sin �

Figure 3.4 Diffraction through slits

Now what in the gases or vapours of these elements are giving rise to the lines in their
spectra? In the case of calcium and mercury, the answer must be their atoms because these
elements are monatomic in the vapour. Sodium and hydrogen, on the other hand, are
diatomic in the gas or vapour. Under the conditions of a discharge, however, the diatomic
molecules will be to some extent dissociated, and from the general similarity in the spectra
of these elements to those of calcium and mercury, we may conclude that these too are
produced by atoms.

10
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Energies of atoms

4 ENERGIES OF ATOMS
If atoms are responsible for the light that is emitted in a discharge tube, they presumably
take electrical energy from the discharge and convert this into light. Since the light is
restricted to certain frequencies, this means that constitutional energies of atoms (i.e. energies
arising from within an atom as opposed to thermal motion of an atom as a whole) are
restricted to certain values, light being emitted when the constitutional energy of an atom
drops from one level to another (Fig. 4.1). The steps in energy are called “quanta” (plural
of “quantum”, from Latin quantus, “how much”). This phenomenon brings us up against
the problem we are addressing: according to classical physics, an atom can have any energy,
and will progressively lose energy until its electrons collapse on to the nucleus.

Energy light

Figure 4.1 Energy levels of an atom showing an


energy drop and light produced

www.sylvania.com

We do not reinvent
the wheel we reinvent
light.
Fascinating lighting offers an infinite spectrum of
possibilities: Innovative technologies and new
markets provide both opportunities and challenges.
An environment in which your expertise is in high
demand. Enjoy the supportive working atmosphere
within our global group and benefit from international
career paths. Implement sustainable ideas in close
cooperation with other specialists and contribute to
influencing our future. Come and join us in reinventing
light every day.

Light is OSRAM

11
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Energies of atoms

To establish the energies an atom can have, we need to know the relationship between the
drop in energy (‫ )’‘”†ܧ‬and the frequency of the light. This can be determined by means of
the apparatus shown in Figure 4.2 (Franck and Hertz 1914).

Figure 4.2 Franck-Hertz apparatus


[ND-RC]

In this, electrons are produced by passing a current through a filament (left). These are
accelerated through the gas or vapour by a positively charged grid, the potential of which
can be varied. Electrons passing through the grid are collected, and the current they produce
measured at G.

As the potential is increased, the current from the collector increases until, at a certain point,
it drops. At this point the gas starts to emit one of the lines in its emission spectrum. Some
examples are shown in Table 4.1, where ‫ ‡ܧ‬is the energy required to produce emission and
ߥ is the frequency of the light emitted. The energy is given by ‫ Ž‡ܸݍ‬where ‫ ݍ‬is the charge
( ݁)) and ܸ‡Ž the electric potential.

What is happening in this experiment is that electrons collide with atoms in the tube,
and impart energy to them. When the potential of the grid is low, the energy imparted by
the electrons is low, and simply raises the thermal energy of the atoms. As the potential is
increased, however, the energy imparted by the electrons goes up. When this energy is high
enough to lift the atoms to a higher constitutional level, this excitation takes place. The atoms
then emit the energy they have absorbed in this way as light and return to the lower level.

12
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Energies of atoms

�em
Gas �em 

He 31.7 × 10–19 J 4.79 × 1015 s–1 6.6 × 10–34 J s

Hg 7.8 × 10 –19
J 1.18 × 1015 s–1 6.6 × 10–34 J s

Mg 4.3 × 10–19 J 6.56× 1014 s–1 6.6 × 10–34 J s

Table 4.1 Energy required to produce emission and the frequency of the light emitted*
*After Bolton, Patterns in Physics (1974)

The ratio in the last column is constant, and allows us to write the relation between energy
drop and frequency of light emitted as

�drop = ℎ� (4.1)

�drop
where= ℎ�is a constant (6.626 × 10‒34 J s). This constant was introduced by Planck in another
context (1900) and is called “Planck’s constant”.

Energy levels of hydrogen atom


The simplest emission spectrum is that of hydrogen (Fig. 3.3). The wavelengths of the lines
are assembled in Table 4.2, along with the corresponding frequencies. As the table shows,
the frequencies are given by the simple equation

� �
� = � ��� − ��
� (� = 3, 4, 5 … )(4.2)

where ‫ ܤ‬ൌ= ܴܿ
3.2879 × 1015 s‒1. This relation

was discovered by Balmer (1885). It is usually
� =
expressed in terms of wavenumber (1/ �) in which case the constant is called the “Rydberg
constant” (ܴ). The two constants are related by ‫ ܤ‬ൌ ܴܿ.

13
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Energies of atoms

Wavelength/Å Frequency/1014 s–1 Eq. (4.2)/1014 s–1

6564.7 4.5665 4.5665 (n = 3)

4862.6 6.1649 6.1648 (n = 4)

4341.6 6.9044 6.9046 (n = 5)

4102.9 7.3064 7.3064 (n = 6)

3971.2 7.5487 7.5488 (n = 7)

3890.1 7.7060 7.7060 (n = 8)

Table 4.2 Lines in the emission spectrum of hydrogen

Similar relations were discovered for light emitted in the ultraviolet (Lyman) and infrared
(Paschen). These are respectively
360°
� = � ��� −



��


� = � ��� − �� � (� = 4, 5, 6 … )(4.4)
thinking
� (� = 2, 3, 4 … )(4.3)
.

360°
thinking . 360°
thinking .
Discover the truth at www.deloitte.ca/careers Dis

© Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers © Deloitte & Touche LLP and affiliated entities.

Deloitte & Touche LLP and affiliated entities.

Discover the truth at www.deloitte.ca/careers


14
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Energies of atoms

These formulae can be combined as follows

� �
� = � ���� − ��
� (� > �� )(4.5)

Thus, from equation (4.1), the changes in energy giving rise to the lines in the spectrum are

� �
�drop = ℎ� ���� − ��
�(4.6)

The energy levels of hydrogen atoms are therefore given by

��
� = − �� (� = 1, 2, 3 … ) (4.7)

� �� � �
� =drop
light being emitted when atoms � ���= − level
�� � ���
from � (� > ) > �� ):
� �level
��� − ��to (�

�� �� � �
�drop = � − �′ = − �� − �− ��� � = ℎ� ���� − ��
�(4.8)

� � �� �� � �
� = � �The
���
− � (� is
number
��
�� )�drop
> called = � − �′ number”.
a “quantum = − � −The
�−value
�� � = ℎ� � �� −
of is 2.1786� �× 10‒18 J (13.598 eV).
� � � �

On the model of Figure 2.1, the constitutional energy of an atom resides in the motion
of the negatively charged electrons round the positively charged nucleus. On this model,
therefore, equation (4.7) gives the different energies the one electron in a hydrogen atom
can have. From this, the energy required to remove the electron completely from the atom
(the “ionization potential”, I) is given by � = −� = ��� . For the atom in its ground state
�� �
� = − �� (�
( = 1,),2,this
3 … is) 13.598 eV.

15
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave-like behaviour

5 WAVE-LIKE BEHAVIOUR
If atoms are restricted in the energy they can have, they differ from, for example, a
pendulum, which can have any amplitude. They are, however, similar to a stretched string,
the vibrations of which are restricted to certain modes (Fig. 5.1). This raises the possibility
that the motions of electrons are wave-like in character.

Figure 5.1 Modes of vibration of a stretched string



(� =
= wavelength, ‫ = ܮ‬length of string)

[Chad Orzel, Forbes]

Now this possibility can be tested for a beam of electrons by passing it through a suitable
grating. The grating has to be fine enough for diffraction to be possible. A suitable grating
is provided by the atoms in a metal foil or crystal. Such a grating gives a diffraction pattern
for X-rays (light of very low wavelength).

When this experiment is carried out, it is found to give a diffraction pattern of the kind
expected for waves (Fig. 5.2). This confirms our speculation that the motion of electrons
is somehow wave-like.

Figure 5.2 Electron diffraction pattern


[Free Patterns]

16
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave-like behaviour

Measurement of the pattern allows the wavelength of the beam to be determined and
compared with the momentum (‫ ݌‬ൌ ݉‫ )ݒ‬of the electrons (݉ = mass, ‫ = ݒ‬speed). The latter

can be determined from the kinetic energy of the electrons ( � = �� � ) as given by the

accelerating potential (� = ��el ). The relation between � = ��el‫ ݌‬is
and

��
� = �� (5.1)

Typical results are set out in Table 5.1. These show that the wavelength of the electron
beam decreases as the momentum of the electrons is raised, and that ߣ‫ ݌‬is approximately
constant, the same constant as in Table 4.1. We can therefore write

�= �
(5.2)

This is called the “de Broglie relation” after the physicist who first suggested it (1924). From
equation (5.1), equation (5.2) can also be written

� �
�= = (5.3)
√��� ���(���)

where ‫ ܧ‬is the total energy and ܸ the potential energy (e.g. of an electron in the field of
other electrical charges).

We will turn your CV into


an opportunity of a lifetime

Do you like cars? Would you like to be a part of a successful brand? Send us your CV on
We will appreciate and reward both your enthusiasm and talent. www.employerforlife.com
Send us your CV. You will be surprised where it can take you.

17
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave-like behaviour


 �= 

3.97 × 10–24 kg m s–1 1.65 × 10–10 m 6.6 × 10–34 J s

4.36 × 10–24 kg m s–1 1.49 × 10–10 m 6.5 × 10–34 J s

6.36 × 10–24 kg m s–1 1.06 × 10–10 m 6.7 × 10–34 J s

Table 5.1 Wavelengths of electron beams diffracted by a crystal of nickel*


*Davisson and Germer (1927)

Other small particles can be diffracted by gratings in a similar way, including atoms.

What are the waves associated with particles? Physicists and philosophers still discuss this
problem. One possibility is that they are in a sub-electronic medium that buffets small
particles like waves on the sea buffeting a buoy, or water molecules buffeting a pollen grain
in Brownian motion.

18
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

6 WAVE MECHANICS
If small particles behave like waves, can we adapt the theory of waves to describe their motion?

Consider the simple case of a particle confined to a one-dimensional box of length L, with
ܸ = 0 and ‫= � = ܧ‬ ��el the box and ܸ= λ outside it. Let us�adapt the theory of waves in a
inside
�=
stretched string to this problem, and restrict the values of � as in Figure 5.1:

��
�= �
(� = 1, 2, 3 …)(6.1)

From equation (5.3), this gives

�� �� (6.2)
�= ����

The energy of the particle is therefore restricted to certain values (Fig. 6.1). This successfully
reproduces what we have found for atoms (Sect. 4). The inverse dependence of ‫ ܧ‬on ݉
and ‫ ܮ‬means that the permitted energies are very much closer together for a macroscopic
system than for one on an atomic scale.

Figure 6.1 Permitted energies of a particle in a box up to n = 4


and the waves associated with them (compare these with
the waves in a stretched string in Figure 5.1)
[Plus Magazine]

The corresponding equation for a particle in a three-dimensional box is


�� �� ��� ��
� = � ��� + + � �� (6.3)
� ��� ���

19
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

This equation correctly predicts the properties of a monatomic gas under ordinary conditions
when substituted into Boltzmann’s distribution law:

�� e��� /��
= ∑� e��� /��
(6.4)

Here i numbers the different motions a particle can have, ܰ௜ is the number executing
motion i, ‫ܧ‬௜ is their energy, ܰ the total number of particles, � =
the��absolute
el temperature,
and ݇ Boltzmann’s constant.

General equation
We can derive a general equation for the motion of a particle in one dimension as follows.

Consider again a wave in a stretched string. The amplitude of this is given by

�(�, �) = �(�) sin 2π�� (6.5)

�e Graduate Programme
I joined MITAS because for Engineers and Geoscientists
I wanted real responsibili� www.discovermitas.com
Maersk.com/Mitas �e G
I joined MITAS because for Engine
I wanted real responsibili� Ma

Month 16
I was a construction Mo
supervisor ina const
I was
the North Sea super
advising and the No
Real work he
helping foremen advis
International
al opportunities
Internationa
�ree wo
work
or placements ssolve problems
Real work he
helping fo
International
Internationaal opportunities
�ree wo
work
or placements ssolve pr

20
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

where

�π�
�(�) = � sin (6.6)

In equation (6.5), the sine function determines the oscillation of the wave up and down,
�(�,its
= �(�) sin 2π�� being �) = �(�) sinand
frequency 2π�� time (sine functions oscillate in value between +1 and ‒1).
�π�
The sine function in equation (6.6) determines the shape of the wave, and the�(�) = �,sin
factor its
�π� �
magnitude. I shall call �(�)
the=“profile”
� sin of the wave. It is the amplitude when (sin 2π��) = −1

and is the solid line in Figure 5.1; the broken line is when (sin 2π��) = −1.

Continuing the analogy with waves in a stretched string, we can write the following equation
for the profile of the wave for a particle in a box

�π�
�(�) = � sin �
(6.7)

�π�
 and= � sin
where  corresponds to �(�) to .�. Introducing equation (5.3) into this gives

��π� �(���)
�(�) = � sin � (6.8)

Double differentiation of this gives

d� � �π� �
+ (� − �)� = 0 (6.9)
d� � ��

The corresponding equation for a particle moving in three dimensions is

�π� �
∇� � + ��
(� − �)� = 0 (6.10)

where ∇� (“del squared”) is the operator

∂� ∂� ∂�
∇� = ∂� � + ∂� � + ∂� � (6.11)

This equation can also be written

��
∇� � + (� − �)� = 0(6.12)
�π� �

The corresponding equation for two or more particles is

��
∑� ∇�� � + (� − �)� = 0 (6.13)
�π� ��

where i numbers the particles (1, 2, 3 etc.) and  is a function of all their positions (x1,
y1, z1; x2, y2, z2; etc.). This is a general equation for systems whose energy does not change

21
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

with time, and is called the “time-independent Schrödinger equation” after the physicist
who devised it (Schrödinger 1926). It underlies much of quantum mechanics as applied
to chemistry. The equation successfully reproduces the energy levels of the hydrogen atom
[equation (4.7)] as we shall see in the next section.

The function  is called the “wave function”. To describe a wave, it has to be everywhere
finite, single-valued, and continuous. These conditions limit the solutions to the Schrödinger
equation, and bring about quantization.

Figure 6.2 Erwin Schrödinger

Physical significance off 


From equation (6.5), the intensity of a wave in a stretched string is given by

�(�, �) = �� (�, �) = �� (�) sin� 2π��(6.14)

�( �( � )
�)(�,
This is positive for all�(�, �and=��). �=
(�,��) =�, ��)�=
(�)�sin sin� 2π��
� 2π��

Now in wave mechanics, time variation is introduced into the Schrödinger equation in such
�� (�,
a way that, for systems with constant energy, the quantity corresponding to �(�, �),=�(�, � ��
�) =
�),= � (�)
(�) sin
does not vary with time (Box A):

�(�, �) = � � (�)(6.15)


From this, Born (1926) suggested that∫� �(�)� d� =1
represents the probability of finding a particle

between and  andd � + d�.This means that, for a particle in a box, ∫� �(�)� d� =1
corresponds
to the density of the image on a long-exposure photograph of the particle moving in the

∫� �(�)� d�
box. Plots of for =n1= 1, 2 and 3 are displayed in Figure 6.3. The condition that the
particle should be found somewhere in the box

22
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics


∫� �(�)� d� = 1(6.16)

fixes the value of � = �2/�. (6.7). This process is called “normalization”. For a particle
in equation
in a box, � = �2/�.

Figure 6.3 Wave functions and probability distributions for a particle in a box
[Brane Space]

23
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

The probabilistic description of the motion of the particle means that there is uncertainty
about the precise motion of the particle, i.e. where it is at a given time and how fast it is
travelling. In other words, “it is impossible to specify simultaneously both the momentum
and position of a particle”. This is the celebrated “Heisenberg uncertainty principle” (1927).

The interpretation of  I have given here is the one usually adopted by chemists. Other
interpretations have been proposed, which I will briefly discuss at the end (Sect. 15).

Wave functions can be “complex”, i.e. contain the square root of minus one (i). In these
cases,  is given by �� ∗ , where � ∗ is the “complex conjugate” of , i.e. , with -i in place
of i;  is then real.

A useful property of wave functions is that they are “orthogonal” to each other, i.e. they

satisfy ∫ �� ��� d� = 0 or ∫ �� ��� d� = 0 (� ≠ �� ).

Box A

Time dependence

For a particle moving in one dimension, the time-independent Schrödinger equation [equation
(6.9)] may be written:

�� d� �
− ��� � d� � + �� = �� (A1)

Now let the corresponding time-dependent equation be:

�� d� � � ��
− ��� � d� � + �Ψ = ±i �π ��
(A2)

where Ψ is a function of  and .. For a system whose energy does not change with time, this
is satisfied by

�π��
�(�, �) = �(�)e‒i �  (A3)

or

�π��
� ∗ (�, �) = �(�)e+i �  (A4)

These equations can be derived from equations (A1) and (A2) by substituting �(�, �) =
�(�)�(�). They give

�(�, �) = �(�, �)� ∗ (�, �) = �2 (�) (A5)

independent of t as required.

24
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Wave mechanics

Solving the Schrödinger equation


In a small number of cases, the Schrödinger equation can be solved exactly. We have
effectively seen this for a particle in a box. In most cases, however, the equation cannot be
solved exactly, and approximate methods have to be used.

One method is to omit the smallest term or terms in the potential energy and to solve the
equation without them. The solutions can then be used to calculate the energy from the full
equation. This is called “perturbation theory”, the small terms being treated as a perturbation.

Another method is to choose a possible solution to the equation with variable parameters
in it and to vary these parameters until they give the lowest energy (e.g. one might try the
function � = �e��� and vary the parameter a). The lowest energy will be closest to the true
energy. This is called the “variation principle”, and the method, the “variation method”.
With skilful choice of functions and the use of modern computers, solutions very close to
the true solution can be obtained in this way.

We will give examples of the use of these methods later.

Simplified form of Schrödinger equation


Equation (6.13) can be rewritten as follows:

��
− ∑� �π� � ∇�� � + �� = �� (6.17)

This can be simplified to

Ĥ� = ��(6.18)

where

��
Ĥ = − ∑� �π� � ∇�� + � (6.19)

�� �
Here Ĥ =is −a ∑mathematical
� �π� � ∇� + � “operator” (it includes differentials) called the “Hamiltonian

operator”. This equation only holds for the correct wave function, but energies can be
obtained from it with approximate functions by multiplying both sides by  (or � ∗ if is 
complex) and integrating over all the coordinates to give what is effectively a mean value:

� = ∫ �Ĥ� dτ ( normalized) (6.20)

�Ĥ� dτ stands for dx1dy1dz1dx2 … We shall use this equation later.


� = ∫Here

25
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Schrödinger’s theory of hydrogen atom

7 SCHRÖDINGER’S THEORY
OF HYDROGEN ATOM
As noted in Section 2, the hydrogen atom is the simplest atom, with one electron (e) moving
round a singly-charged nucleus (n). The potential energy in this case is the Coulomb energy
of attraction between the electron and the nucleus:
(�)(��)
�= (7.1)
�π�� �

Here  is the distance between the electron and the nucleus, and  is the permittivity of
free space.

The appropriate Schrödinger equation is equation (6.13) with � = 1 for the electron and
� = 2 for the nucleus. This equation describes essentially two motions, the motion of the
electron round the nucleus and the motion of the atom as a whole through space. These
motions can be separated by introducing the reduced mass, ,, defined by
� � �

= �e
+ �n
(7.2)

26
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Schrödinger’s theory of hydrogen atom

The motion of the electron round the nucleus is then given by equation (6.12) with � = �:

�� ��
∇� � + �� + �π� �� � = 0(7.3)
�π� � �

Here  is a function of the position of the electron relative to the nucleus. Because �n ≫ �e ,
� ≈ �e .

Equation (7.3) can be solved exactly. The result is that the energy is restricted to the
following values:


�=− ��
(� = 1, 2, 3 … ) (7.4)

where

�� �
� = ��� �� = 2.1787 × 10‒18 J (13.598 eV) (7.5)

This almost exactly reproduces equation (4.7).

Orbitals
The solutions  of equation (7.3) are called “orbitals”. They give the probability of finding
the electron at different points around the nucleus, and broadly reflect their motion (Sect.
6). The quantum number n determines the size of an orbital: the bigger n, the bigger the
orbital, and the lower the binding energy.

For � > 1, there are several motions having the same energy. This is called “degeneracy”.
These motions are characterized by two subsidiary quantum numbers,  and  . The first of
these can take the values � = 0, 1, 2 … (� − 1) and determines the shape of an orbital (more
precisely, the angular momentum of the electron). Orbitals having � = 0 are spherically
symmetrical and are called s orbitals. Orbitals having � = 1 are called p, those having � = 2
are called d, and those having � = 3, f. The labels derive from the names for series in alkali
metal spectra.

The second subsidiary quantum number takes the values �� = 0, ±1, ±2 … ± � and determines
the directionality of an orbital (more precisely, the component of the angular momentum
in a given direction). The number of orbitals for each value of  is therefore (2l + 1) .

The various orbitals for n = 1, 2, and 3 are set out in Table 7.1 below. A selection are
shown in Figure 7.1.

27
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Schrödinger’s theory of hydrogen atom

� > 1,   No. of orbitals Designation

1 0 0 1 1s

0 0 1 2s
2
1 0, ±11
0, 3 2p

0 0 1 3s

3 1 0, ±11
0, 3 3p

2 0,0, ±11,0, ±12 5 3d

Figure 7.1 Schrödinger orbitals for a hydrogen atom (sections in x–z plane,
probability densities represented by densities of dots)
[Forgotten Planet]

Orbitals having different  values are not uniquely determined by equation (7.3) and can
be combined. For example, p orbitals can be combined to give three of the kind shown in
Figure 7.1 pointing in the x, y, and z directions:

p� = (−p�� + p�� )(7.6)
√�


p� = (p�� + p�� )(7.7)
√�

p� = p�(7.8)


The factor comes from normalization (Sect. 6).
√�

28
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Schrödinger’s theory of hydrogen atom

Exercise

Show that the orbitals p� and p� have the same energy (E) as p�� and p�� . Use the simplified
equation Ĥ� = �� [equation (6.18)].

��
� = − �for
Figure 7.2 shows the energies of orbitals � (� = 1, 2, 3 … )
1, 2, and 3.

3s 3p 3d
(1) (3) (5)
2s 2p
(1) (3)

1s
(1)

Figure 7.2 Energies of orbitals for a hydrogen atom on


Schrödinger’s theory with numbers of orbitals in brackets

Hydrogen-like ions
The above treatment can readily be extended to the hydrogen-like ions He+, Li2+, etc. For
these, equation (7.1) becomes

ሺ௓௘ሻሺି௘ሻ
ܸ ൌ (7.9)
ସɎఌబ ௥

where ܼ is atomic number and ܼ݁ nuclear charge. Equation (7.4) is then

௓ మௐ
‫ ܧ‬ൌ െ (7.10)
௡మ

from which

ଵ ଵ
‫ ’‘”†ܧ‬ൌ ݄ߥ ൌ ܼ ଶ ܹ ቀ௡ᇲమ െ  ௡మቁ(7.11)

Lines corresponding to this equation have been observed for He+ (ܼ ൌ ʹ) , Li2+ (ܼ ൌ ͵),
and Be3+ (ܼ ൌ Ͷ).

29
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Dirac’s theory of hydrogen atom

8 DIRAC’S THEORY OF
HYDROGEN ATOM
There are several indications that Schrödinger’s theory is not completely accurate. One is
that, under high resolution, lines in the emission spectrum of hydrogen exhibit fine structure.
For example, the line at about 6564.7 Å (Table 4.2) is a doublet, with a splitting of 0.14 Å.

A second indication is that, when a beam of silver atoms is passed through a non-uniform
magnetic field, it splits into two (Stern and Gerlach 1922). The same happens with hydrogen
atoms (Phipps and Taylor 1927). This is not expected on Schrödinger’s theory.

A third indication is the occurrence of surfaces in probability distributions for which ߰ ଶ ൌ Ͳ


These surfaces are called “nodes”. Nodes occur, for example, in the distribution for a particle
in a box when ݊ ൐ ͳ (Fig. 6.3). They also occur in orbitals for a hydrogen atom, e.g. the
x‒y plane of 2p0 (Fig. 7.1). Now the question is, if the probability of a particle being at a
node is zero, how does the particle get from one side of the node to the other? The answer
to this question has to be that the particle travels infinitely quickly. This goes against the
theory of relativity, which limits the speed of a particle to the speed of light (Box B).

Excellent Economics and Business programmes at:

“The perfect start


of a successful,
international career.”

CLICK HERE
to discover why both socially
and academically the University
of Groningen is one of the best
places for a student to be
www.rug.nl/feb/education

30
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Dirac’s theory of hydrogen atom

Box B

Relativity

Einstein (1905) arrived at his theory of relativity by working out what effect the constancy of
the speed of light in a medium has on how different observers see a moving object. The same
results can be derived experimentally by employing the apparatus used to determine the ratio
of mass to charge of an electron (see my e-book Introduction to Chemistry). This enables the
mass (݉) of an electron to be determined at different speeds (ߥ). The results are shown in the
Figure. The mass increases with speed, and tends towards infinity as the speed approaches the
speed of light (c = 2.998 × 108 m s−1).

3.5

2.5
m/10-27 g

1.5

0.5
0 0.5 1 1.5 2 2.5 3

v/108 m s-1

Figure B1 Plot of electron mass against speed

When these results are fitted to simple equations, the best fit is obtained with the following
equation, the same as that obtained by Einstein:

‫ ݕ‬ൌ ξଵି௫ మ (B1)

where ‫ ݔ‬ൌ ‫ݒ‬Ȁܿ , ‫ ݕ‬ൌ ݉Ȁ݉଴ , and ݉଴ ൌ ݉ when ‫ ݒ‬ൌ Ͳ (the “rest mass”, 0.9109 × 10–27 g).
From this equation, other equations can be derived, including equation (8.1) in the main text
(see Appendix).

Dirac (1928) adapted Schrödinger’s theory to make it consistent with the theory of relativity.
Schrödinger’s theory starts from equation (5.1). Dirac starts from the corresponding
relativistic equation

31
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Dirac’s theory of hydrogen atom

ܶ ൌ ඥ‫݌‬ଶ ܿ ଶ ൅ ݉଴ଶ ܿ ସ (8.1)

From this he derives four simultaneous equations, the solutions of which give four functions
( ߰ଵ , ߰ଶ , ߰ଷ , and ߰ସ ), from which the probability distribution is calculated as

ߩ ൌ ߰ଵ ߰ଵ‫ כ‬+ ߰ଶ ߰ଶ‫ כ‬+ ߰ଷ ߰ଷ‫ כ‬+߰ସ ߰ସ‫(כ‬8.2)

States are characterized by four quantum numbers: ݊ , ݈ , ݆ , and ݉௝ . The first two are the
same as on Schrödinger’s theory, and take the values ݊ ൌ ͳǡ ʹǡ ͵ ǥ and ݈ ൌ Ͳǡ ͳǡ ʹ ǥ ሺ݊ െ ͳሻ.

The new quantum number ݆ takes the values ݆ ൌ ݈ േ ଶ , and ݉௝ takes the values ݉௝ ൌ
ଵ ଷ
േ ǡ േ ǥേ ݆ , totalling ሺʹ݆ ൅ ͳሻ values. The states for ݊ ൌ 1 and 2 are set out in Table 8.1.
ଶ ଶ

݊ ݈ ݆ ݉௝ No. of orbitals

ͳ ଵ ଵ
1 0 ǡെଶ 2
ʹ ଶ

ͳ ଵ ଵ 2
0 ǡെଶ
ʹ ଶ
͵ ͵ ͳ ͳ ͵
2 1 ǡ ǡെ ǡെ
ʹ ʹ ʹ ʹ ʹ
6
ͳ ଵ ଵ
ǡെଶ
ʹ ଶ

Table 8.1 Dirac orbitals for ݊ ൌ 1 and 2

Dirac’s theory resolves the problems with Schrödinger’s theory discussed earlier. None of the
probability distributions has nodes. The nearest they get are surfaces of very low probability,
which the particle can cross with the speed of light.

Also, the energy is given by equation (7.4) with a small addition that gives rise to the fine
structure:

ௐ ௐఈ మ ଵ ଷ
‫ ܧ‬ൎ െ ௡మ െ ቆ భ െ ସ௡ቇ(8.3)
௡య ௝ା

Here ߙ is the “fine-structure constant”:

௘మ
ߙ ൌ  ଶఢ (8.4)
బ ௛௖

Figure 8.1 shows the energies of orbitals on Dirac’s theory for ݊ ൌ 1 , 2, and 3 (compare
Fig. 7.2).

32
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Dirac’s theory of hydrogen atom

3s 3p 3d
(2) (6) (10)
2s 2p
(2) (6)

‫ܧ‬

1s
(2)

Figure 8.1 Energies of orbitals for a hydrogen atom on


Dirac’s theory with numbers of orbitals in brackets.
Levels are split, but by amounts too small to be shown.

Figure 8.2 Paul Dirac

33
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of hydrogen atom

9 MODIFIED SCHRÖDINGER
THEORY OF HYDROGEN ATOM
Dirac’s theory of the hydrogen atom is mathematically complicated, and is not easily
extended to other atoms. It does, however, suggest a simple modification of Schrödinger’s
theory. Comparison of Figures 7.2 and 8.1 shows that there are twice as many orbitals on
Dirac’s theory as on Schrödinger’s. To bring the two sets of numbers into line, we can add
to Schrödinger’s quantum numbers ݊ , ݈ , and ݉௟ two further quantum numbers, s and ݉௦
ଵ ଵ
taking the values ‫ ݏ‬ൌ ଶ and ݉௦ ൌ േ‫ ݏ‬ൌ േ ଶ . This is shown in Table 9.1. The additional
quantum numbers are called “spin” quantum numbers, the spin being determined by s and
its orientation by ݉௦. Analysis of Dirac’s orbitals, however, suggests that the numbers are
associated more with a corkscrew motion, right-handed or left-handed. Orbitals having
different values of ݉௦ are called “spin-orbitals”.

American online
LIGS University
is currently enrolling in the
Interactive Online BBA, MBA, MSc,
DBA and PhD programs:

▶▶ enroll by September 30th, 2014 and


▶▶ save up to 16% on the tuition!
▶▶ pay in 10 installments / 2 years
▶▶ Interactive Online education
▶▶ visit www.ligsuniversity.com to
find out more!

Note: LIGS University is not accredited by any


nationally recognized accrediting agency listed
by the US Secretary of Education.
More info here.

34
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of hydrogen atom

No. of
݊ ݈ ݉௟ ݉௦ ݉௟ ൅ ݉௦
orbitals

ଵ ଵ
1 0 0 ± ǡെଶ 2

ଵ ଵ
0 0 ± ǡെଶ 2

ଵ ଵ
2 1 0 ± ǡെଶ

6
͵ ͳ ͳ ͵
0, ±11 ± ǡ ǡെ ǡെ
ʹ ʹ ʹ ʹ

Table 9.1 Spin-orbitals for ݊ ൌ 1 and 2

Spin-orbitals (χ) are derived from Schrödinger orbitals by multiplying ߰ by ߙ for ݉௦ ൌ



± or ߚ for ݉௦ ൌ െ ଶ , where ߙ and ߚ are functions of a spin coordinate (ߦ) satisfying the
normalization equations ‫ ߙ ׬‬ଶ †ߦ ൌ ‫ߚ ׬‬ଶ †ߦ ൌ ͳ and the orthogonality equation ‫ ߦ†ߚߙ ׬‬ൌ Ͳ.

35
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

10 MODIFIED SCHRÖDINGER
THEORY OF OTHER ATOMS
For atoms and ions containing two or more electrons, the appropriate Schrödinger equation
for the motion of the electrons round the nucleus is equation (6.13) with ݉௜ ൌ ߤ :

௛మ
σ௜ ‫׏‬ଶ௜ ߰ + ሺ‫ ܧ‬െ ܸ ሻ߰ = 0(10.1)
଼Ɏమ ఓ

The potential energy is given by

௓௘ మ ௘మ
ܸ ൌ  െ σ௜ ସɎఌ ൅ σ௜ σ௝வ௜ ସɎఌ (10.2)
బ ௥೔ బ ௥೔ೕ

where ‫ݎ‬௜௝ is the distance between electron i and electron j. Equation (10.1) can be solved
by various approximate methods.

The simplest is to neglect the interelectronic repulsion terms in the equation. These are
smaller than the nuclear attraction terms. The solutions then have the form

߰ ൎ ߮ƒ ሺͳሻ߮„ሺʹሻ ǥ(10.3)

where the electrons (numbered 1, 2 …) are in orbitals (lettered a, b …) satisfying the


Schrödinger equation for

௛మ ௓௘ మ
‫׏‬ଶ ߮ + ቀ‫ ܧ‬൅ ସɎఌ ௥ቁ ߮ = 0(10.4)
଼Ɏమ ఓ బ

Each electron is then in its own orbital and its energy is given by equation (7.7):

௓ మௐ
‫ ܧ‬ൌ െ ௡మ
(10.5)

This is called “the orbital approximation”.

This version of it is very crude. Electrons do repel each other. A refinement is to treat the
repulsion between each electron as having the effect of reducing the positive charge on the
nucleus:

௛మ ሺ௓ିௌ ሻ௘ మ
଼Ɏమ ఓ
‫׏‬ଶ ߮ + ሾ‫ ܧ‬൅ ସɎఌబ ௥
ሿ߮ = 0(10.6)

36
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

S is called the “screening constant” and ሺܼ െ ܵሻ the “effective nuclear charge” (ܼ‡ˆˆ ). The
energy is now

௓‡ˆˆ ௐ
‫ ܧ‬ൌ െ ௡మ
(10.7)

Values of ܼ‡ˆˆ can be inferred from ionization energies derived from emission spectra (Sect. 4).

Thus, for example, on this approximation, a helium atom (ܼ ൌ ʹ) has in its ground state
two electrons in a 1s orbital. This is called its “electronic configuration” and written 1s2.
The energy required to remove an electron (the first ionization energy) is ‫ܫ‬ଵ ൌ ͳǤͺͳܹǤ This
is less than the energy required to remove the second electron (‫ܫ‬ଶ ൌ ܼ ଶ ܹ ൌ Ͷܹ) reflecting
the repulsion between them. The lower value corresponds to ܼ‡ˆˆ ൌ ͳǤ͵ͷǤ

When we come to the next atom (lithium, ܼ ൌ ͵), we have to introduce a new principle.
This is because, in its ground state, this does not have three electrons in a 1s orbital. The
first ionization energy is ‫ܫ‬ଵ ൌ ͲǤ͵ͻܹǤ If ݊ ൌ 1 , this corresponds to ܵ ൌ ʹǤͶ, which exceeds
the number of other electrons to do the screening (two). The third electron thus has to go
into an orbital with ݊ ൌ ʹ, in which case ܵ ൌ ͳǤͺ.

37
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

The new principle is that electrons cannot have the same motion. In other words, two
or more electrons cannot occupy the same spin-orbital. This is called “the Pauli exclusion
principle”. Thus, in a helium atom, if one electron occupies the 1s+ spin-orbital (the 1s

orbital with ݉௦ ൌ ൅ ଶ), the other has to occupy the 1s–. Likewise, in lithium, the third
electron has to go into a higher spin-orbital.

For a given value of n, screening constants increase in the order s > p > d > f. Thus for
lithium, the 2s spin-orbitals have a lower energy than the 2p. This is shown in Figure 10.1.
The ground-state configuration of a lithium atom is thus 1s22s1.

3p 3d
3s
(6) (6)
(2)
2p
2s
(6)
(2)

‫ܧ‬

1s
(2)

Figure 10.1 Energies of spin-orbitals for a lithium atom


with numbers of spin-orbitals in brackets

We are now in a position to work out the electronic configurations of other atoms. This can
be done by using the “building-up principle”. This gives the order of filling of spin-orbitals
in going from one atom to another. This is

1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s …

Note that the 3d spin-orbitals are out of line. Some configurations are irregular because
the balance between nuclear attraction and interelectronic repulsion is very close for some
atoms. Configurations for atoms up to ܼ ൌ ͵͸ in their ground states are given in Table
10.1, with irregular ones in red.

38
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

Atom Z Configuration ܰ‘—–‡”

H 1 1s1 1
He 2 1s2 2/0*
Li 3 1s22s1 1
Be 4 1s22s2 2
B 5 1s22s22p1 3
C 6 1s22s22p2 4
N 7 1s22s22p3 5
O 8 1s22s22p4 6
F 9 1s22s22p5 7
Ne 10 1s22s22p6 8/0*
Na 11 1s22s22p63s1 1
Mg 12 1s22s22p63s2 2
Al 13 1s22s22p63s23p1 3
Si 14 1s22s22p63s23p2 4
P 15 1s22s22p63s23p3 5
S 16 1s22s22p63s23p4 6
Cl 17 1s22s22p63s23p5 7
Ar 18 1s22s22p63s23p6 8/0*
K 19 1s22s22p63s23p64s1 1
Ca 20 1s22s22p63s23p64s2 2
Sc 21 1s22s22p63s23p64s23d1 3
Ti 22 1s22s22p63s23p64s23d2 4
V 23 1s22s22p63s23p64s23d3 5
Cr 24 1s22s22p63s23p64s13d5 6
Mn 25 1s22s22p63s23p64s23d5 7
Fe 26 1s22s22p63s23p64s23d6 8
Co 27 1s22s22p63s23p64s23d7 9
Ni 28 1s22s22p63s23p64s23d8 10
Cu 29 1s22s22p63s23p64s13d10 11
Zn 30 1s22s22p63s23p64s23d10 12/2*
Ga 31 1s22s22p63s23p64s23d104p1 3
Ge 32 1s22s22p63s23p64s23d104p2 4
As 33 1s22s22p63s23p64s23d104p3 5
Se 34 1s22s22p63s23p64s23d104p4 6
Br 35 1s22s22p63s23p64s23d104p5 7
Kr 36 1s22s22p63s23p64s23d104p6 8/0*

Table 10.1 Electronic configurations of atoms up to ܼ ൌ ͵͸ and number of outer electrons


*Number if outer shell is taken to be that of the preceding atom/next atom.

Configurations can be simplified by representing the configuration of the preceding inactive


gas by its symbol in bold, and by enclosing the inner electrons in square brackets. This
gives for bromine, for example, [Ar3d10]4s24p5. Note how the 3d electrons have become
part of the core.

39
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

Table 10.1 can be extended to heavier atoms. Silver atoms used in the Stern-Gerlach
experiment have a similar configuration to copper, viz. [Kr4d10]5s1. With their one outer
electron, they behave in this experiment like hydrogen atoms.

The configurations in Table 10.1 correlate with the Periodic Table and Lewis’s model of the
atom (Lewis 1916). Thus, for example, the inactive gases (high-lighted) all have complete
shells. The alkali metals have one more electron than this, which they can lose to form M+
ions. The halogens have one electron less, and can gain an electron to form X‒ ions. For
further details, see my e-book Introduction to Chemistry.

Configurations of ions do not follow the same order of filling as atoms. For cations the
order becomes

1s, 2s, 2p, 3s, 3p, 3d, 4s, 4p, 5s …

with the 3d spin-orbitals in line. Thus, for example, while the configuration of an iron
atom is [Ar]4s23d6, the configuration of a Fe2+ ion is [Ar]3d6 and of a Fe3+ ion is [Ar]3d5.

40
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

Indistinguishability of electrons
For many atoms and ions, equation (10.3) needs to be modified to take account of the
indistinguishability of electrons. Consider a helium atom with the configuration 1s12s1. For
this, equation (10.3) gives

߰ሺͳǡʹሻ ൎ ߮ͳ• ሺͳሻ߮ʹ• ሺʹሻ or ߮ʹ• ሺͳሻ߮ͳ•ሺʹሻ(10.8)

from which

ଶ ሺ ሻ ଶ ሺ ሻ ଶ ଶ
ߩሺͳǡʹሻ ൎ ߮ͳ• ͳ ߮ʹ• ʹ RU߮ʹ• ሺͳሻ߮ͳ• ሺʹሻ(10.9)

These functions distinguish the electrons: the first places electron 1 in the 1s orbital and
electron 2 in the 2s; the second does the reverse.

This problem can be overcome by taking linear combinations


߰ଵ ሺͳǡʹሻ ൎ ሾ߮ͳ• ሺͳሻ߮ʹ•ሺʹሻ + ߮ʹ• ሺͳሻ߮ͳ• ሺʹሻሿ(10.10)
ξଶ

or


߰ଶ ሺͳǡʹሻ ൎ ሾ߮ͳ• ሺͳሻ߮ʹ•ሺʹሻ ‒ ߮ʹ• ሺͳሻ߮ͳ• ሺʹሻሿ(10.11)
ξଶ

from which


ߩଵ ሺͳǡʹሻ ൎ ሾ߮ͳ• ሺͳሻ߮ʹ• ሺʹሻ + ߮ʹ• ሺͳሻ߮ͳ• ሺʹሻሿ2(10.12)

or


ߩଶ ሺͳǡʹሻ ൎ ଶ ሾ߮ͳ• ሺͳሻ߮ʹ• ሺʹሻ ‒ ߮ʹ• ሺͳሻ߮ͳ• ሺʹሻሿ2(10.13)

In equations (10.12) and (10.13), electrons 1 and 2 can now be interchanged without
and ͳ come from the normalization (Sect. 6).

changing the function. The factors
ξଶ ʹ

The two states described by equations (10.10)‒(10.13) I consider further below.

41
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

Spin-orbitals
To obtain spin-orbitals, equation (10.3) or its combinations have to be multiplied by spin
functions. For two electrons, these are

ߪଵ ሺͳǡʹሻ ൌ ߙሺͳሻߙሺʹሻ(10.14)

ߪଶ ሺͳǡʹሻ ൌ ߙሺͳሻߚሺʹሻ(10.15)

ߪଷ ሺͳǡʹሻ ൌ ߚሺͳሻߙሺʹሻ(10.16)

ߪସ ሺͳǡʹሻ ൌ ߚሺͳሻߚሺʹሻ(10.17)

Of these, ߪଵ and ߪସ do not distinguish the electrons, but ߪଶ and ߪଷ do, and must be
replaced by linear combinations:


ߪା ሺͳǡʹሻ ൌ ሾߙሺͳሻߚሺʹሻ ൅ ߚሺͳሻߙሺʹሻሿ(10.18)
ξଶ


ߪି ሺͳǡʹሻ ൌ ሾߙሺͳሻߚሺʹሻ െ ߚሺͳሻߙሺʹሻሿ(10.19)
ξଶ

As shown in Box C, functions ߪଵ, ߪା, and ߪସ correspond to a total spin of one, function
ߪି to a total spin of zero. In the latter, the electrons are paired, in the former, unpaired.

Box C

Simplified treatment of spin

To calculate the total spin for a system of two electrons, we define a mathematical operator å
by the equation

åఈ åఉ

ൌ ఉ
ൌ ‫ݏ‬ (C1)

where ‫ݏ‬ ൌ ଶ . Thus for ߪଵ ሺͳǡʹሻ, the total spin is given by
ሺåభ ାåమሻఙభ ሺଵǡଶሻ ሺåభ ାåమ ሻఈሺଵሻఈሺଶሻ ሾåభ ఈሺଵሻሿఈሺଶሻାఈሺଵሻሾåమ ఈሺଶሻሿ
total spin ൌ
ఙభ ሺଵǡଶሻ
ൌ ఈሺଵሻఈሺଶሻ
ൌ ఈሺଵሻఈሺଶሻ
ൌ ʹ‫ ݏ‬ൌ ͳ (C2)

Similarly for ߪା, ߪି, and ߪସ .

General Pauli principle


We have seen that the Pauli exclusion principle requires the ground state of the helium
atom to be 1s2 with the two electrons paired. The total wave-function for this is obtained
by combining equation (10.3)

42
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

߰ሺͳǡʹሻ ൎ ߮ͳ• ሺͳሻ߮ͳ• ሺʹሻ(10.20)

with ߪି [equation (10.19)] to give


߯ሺͳǡʹሻ ൎ ߮ͳ• ሺͳሻ߮ͳ•ሺʹሻሾߙሺͳሻߚሺʹሻ െ ߚሺͳሻߙሺʹሻሿ(10.21)
ξଶ

This equation differs from the equations that would be obtained by combining equation
(10.20) with the other spin functions (ߪଵ, ߪା, and ߪସ) in that it changes sign when electrons
1 and 2 are interchanged. It is said to be “asymmetric” with respect to this interchange.
This leads to the general Pauli principle: wave functions of electron systems are asymmetric
with respect to interchange of electrons.

We can now return to the 1s12s1 configuration of the helium atom. To satisfy the general
Pauli principle, equation (10.10) must be combined with ߪି to produce a single state
(“singlet”), whereas equation (10.11) must be combined with ߪଵ, ߪା, and ߪସ to produce a
“triplet”. Thus from equations (10.10) and (10.19),


߯•‹‰Ž‡– ሺͳǡʹሻ ൎ ଶ ሾ߮ͳ•ሺͳሻ߮ʹ• ሺʹሻ + ߮ʹ• ሺͳሻ߮ͳ• ሺʹሻሿሾߙሺͳሻߚሺʹሻ െ ߚሺͳሻߙሺʹሻሿ(10.22)

while from equations (10.11), (10.14), (10.17), and (10.18),

Join the best at Top master’s programmes


• 3
 3rd place Financial Times worldwide ranking: MSc
the Maastricht University International Business
• 1st place: MSc International Business
School of Business and • 1st place: MSc Financial Economics
• 2nd place: MSc Management of Learning

Economics! • 2nd place: MSc Economics


• 2nd place: MSc Econometrics and Operations Research
• 2nd place: MSc Global Supply Chain Management and
Change
Sources: Keuzegids Master ranking 2013; Elsevier ‘Beste Studies’ ranking 2012;
Financial Times Global Masters in Management ranking 2012

Maastricht
University is
the best specialist
university in the
Visit us and find out why we are the best! Netherlands
(Elsevier)
Master’s Open Day: 22 February 2014

www.mastersopenday.nl

43
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

- ߙሺͳሻߙሺʹሻ (10.23)
ଵ ଵ
߯–”‹’Ž‡– ሺͳǡʹሻ ൎ ሾ߮ͳ• ሺͳሻ߮ʹ•ሺʹሻ ‒ ߮ʹ•ሺͳሻ߮ͳ• ሺʹሻሿ- ሾߙሺͳሻߚሺʹሻ ൅ ߚ ሺͳሻߙሺʹሻሿ (10.24)
ξଶ ξଶ

- ߚሺͳሻߚሺʹሻ (10.25)

Energies
For the 1s12s1 configuration of the helium atom, which is lower in energy, the singlet or the
triplet? Spectroscopy indicates the triplet, but this can be derived by substituting equations
(10.22) and (10.23), (10.24), or (10.25) into the modified form of equation (6.20):

‫ ܧ‬ൌ ‫† ߯>߯ ׬‬ɒ (10.26)

Here > for a helium atom is, from equations (10.1) and (10.2),

௛మ ௛మ ʹ݁ʹ ʹ݁ʹ ݁ʹ
> ൌ െ ଼Ɏమఓ ‫׏‬ଵଶ െ ଼Ɏమ ఓ ‫׏‬ଶଶ െ ͶɎߝ െ ͶɎߝ ൅ ͶɎߝ (10.27)
Ͳ ‫ͳݎ‬ Ͳ ‫ʹݎ‬ Ͳ ‫ʹͳݎ‬

This can be simplified to

݁ʹ
> ൌ >ଵ ൅ >ଶ ൅ ͶɎߝ (10.28)
Ͳ ‫ʹͳݎ‬


where > is the Hamiltonian operator for an electron moving in the field of the nucleus.

The results are:

 
‫ –‡Ž‰‹•ܧ‬ൎ ‫ܧ‬ଵ• ൅ ‫ܧ‬ଶ• ൅ ‫ ܬ‬൅ ‫(ܭ‬10.29)

 
‫ –‡Ž’‹”–ܧ‬ൎ ‫ܧ‬ଵ• ൅ ‫ܧ‬ଶ• ൅ ‫ ܬ‬െ ‫(ܭ‬10.30)

where

 
‫•ͳܧ‬ ൌ ‫(߬† •ͳ߮ > •ͳ߮ ׬‬10.31)

 
‫•ʹܧ‬ ൌ ‫(߬† •ʹ߮ > •ʹ߮ ׬‬10.32)

௘మ మ ሺଵሻఝమ ሺଶሻ
ఝͳ•
‫ܬ‬ൌቀ ቁ‫׭‬ ʹ•
†߬ଵ †߬ଶ(10.33)
ସɎఌబ ௥భమ

௘మ ఝͳ•ሺଵሻఝʹ• ሺଵሻఝͳ•ሺଶሻఝʹ•ሺଶሻ
‫ ܭ‬ൌ ቀସɎఌ ቁ ‫׭‬ ௥భమ
†߬ଵ †߬ଶ(10.34)

 
‫•ͳܧ‬ and ‫•ʹܧ‬ are negative quantities, ‫ ܬ‬and ‫ ܭ‬are positive. Thus, if the orbitals are the same
in the two states, ‫ –‡Ž’‹”–ܧ‬is more negative than ‫–‡Ž‰‹•ܧ‬. This result remains true if the orbitals

44
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of other atoms

are optimized for each state. ‫ ܬ‬represents the classical energy of repulsion between the
ଶ ଶ
charge distributions ߮ͳ• ݁ and ߮ʹ• ݁ and is called the “Coulomb energy”. ‫ ܭ‬is a non-classical
quantity, and is called the “exchange energy”. It is this that gives, when the orbitals are the
same, the high-spin state a lower energy than the low-spin.

Self-consistent field (SCF) calculations


In the previous section, I did not specify the orbitals. For simplicity, they could have
been the orbitals for the He+ ion. To improve on these, we can use the variation method
(Sect. 6). Suitable functions are chosen for ߮ͳ• and ߮ʹ• containing variable parameters. The
parameters in ߮ʹ• are first varied to minimize the energy, then the parameters in ߮ͳ•, then
again the parameters in ߮ʹ•, and so on, until the results do not change. At this point, the
fields created by ߮ͳ• and ߮ʹ• are self-consistent.

The best possible energy obtained in this way is less negative than the true value because
of the limitations of the orbital approximation. Electrons do not move in orbitals as if the
other electrons are not there. Rather, they tend to keep away from each other. Their motions
are accordingly correlated. The energy arising from this is called the “correlation energy”.
This can be estimated by comparison with similar systems, or computed by a calculation
beyond the orbital approximation.

45
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

11 MODIFIED SCHRÖDINGER
THEORY OF MOLECULES
Since, in general, nuclei move more slowly than electrons, we can treat the electrons in a
molecule as moving round fixed nuclei. This is called the “Born-Oppenheimer approximation”.
The appropriate Schrödinger equation in this case is again equation (10.1), i.e.

௛మ
σ௜ ‫׏‬ଶ௜ ߰ + ሺ‫ ܧ‬െ ܸ ሻ߰ = 0(11.1)
଼Ɏమ ఓ

Hydrogen molecule-cation
The simplest molecule is the hydrogen molecule-cation, ଶା , with one electron moving
round two nuclei. For this, equation (11.1) simplifies to

௛మ
‫׏‬ଶ ߰ + ሺ‫ ܧ‬െ ܸ ሻ߰ = 0(11.2)
଼Ɏమ ఓ

AXA Global
Graduate Program
Find out more and apply

46
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

with

௘మ ௘మ ௘మ
ܸ ൌ  െ ସɎఌ െ ସɎఌ ൅ ସɎఌ (11.3)
బ ௥ బ ௥ బ ௥

where A and B label the nuclei, ‫ ݎ‬is the distance of the electron from nucleus A, ‫ݎ‬
from nucleus B, and ‫ ݎ‬the distance between the two nuclei. The last term represents the
internuclear repulsion and is a constant in the calculation.

Equation (11.2) can be solved exactly, and gives a series of energy levels and orbitals
(“molecular orbitals”, MO) in the same way as the Schrödinger equation for the hydrogen
atom. The equation can also be solved approximately by a method that can be used for other
molecules where an exact solution is not possible. This is based on the idea that, when the
electron is near one nucleus, its motion will be similar to that in an atom centred on that
nucleus. This suggests the approximation

߰ ൎ ܿ ߮ ൅ ܿ ߮(11.4)

where ߮ is the wave-function for a hydrogen atom centred on nucleus A and ߮ on B


(these may be 1s, 2s, 2p etc.); ܿ and ܿ are numerical coefficients whose values have to be
determined. This approximation involves a linear combination (LC) of atomic orbitals (AO)
and is called the “LCAO approximation”. Equation (11.4) gives the probability distribution as

ߩ ൎ ܿଶ ߮ଶ ൅ ʹܿ ܿ ߮ ߮ ൅ ܿଶ ߮ଶ(11.5)

In the case of the ଶା ion, symmetry requires ܿଶ in this equation to be equal to ܿଶ or ܿ ൌ
േܿ . The value of ܿ comes from the normalization condition

‫ ߬†ߩ ׬‬ൌ ͳ(11.6)

Substituting equation (11.5) into this gives


ܿ ൌ (11.7)
ඥሺଶേଶௌሻ

where ܵ ൌ ‫ ߬† ߮ ߮ ׬‬. This is called the “overlap integral” as it measures the degree of overlap
of ߮ and ߮ . It varies in value between zero (no overlap) and one (complete overlap). For
most molecules at their equilibrium internuclear distance, S is small (0.2‒0.3), but for the
ଶା ion and H2 molecule it is higher (0.59 and 0.73 respectively).

The atomic orbitals ߮ and ߮ therefore give two molecular orbitals

ఝ ାఝ
߰ା ൎ (11.8)
ඥሺଶାଶௌሻ

47
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

and
ఝିఝ
߰ି ൎ (11.8)
ඥሺଶିଶௌሻ

From equation (6.20), these give

ఈାఉ
‫ܧ‬ା ൎ (11.9)
ଵାௌ

and

ఈିఉ
‫ ିܧ‬ൎ (11.10)
ଵିௌ

where ߙ ൌ ‫† ߮> ߮ ׬‬ɒ ൌ ‫† ߮> ߮ ׬‬ɒ and ߚ ൌ ‫† ߮> ߮ ׬‬ɒ ൌ ‫† ߮> ߮ ׬‬ɒ . . The integral ߙ is
called a “Coulomb integral” and is negative; it is approximately equal to the energy of an
electron in orbital ߮ of a hydrogen atom. The integral ߚ is called a “resonance integral”
and is also generally negative.

For the ଶା ion at its equilibrium internuclear distance, calculations give ߚ ൌ ͲǤͺͺߙ . With S
= 0.59, the values of ‫ܧ‬ା and ‫ ିܧ‬are accordingly 1.18ߙ and 0.29ߙ respectively. ‫ܧ‬ା is therefore
more negative than the energy of an electron in orbital ߮ of a hydrogen atom, while ‫ ିܧ‬is
less negative. Consequently, ߰ାଶ is called a “bonding orbital” and ߰ିଶ an “antibonding orbital”.

All the orbitals of a hydrogen atom give rise to molecular orbitals. The lowest energy is
obtained with 1s orbitals. The molecular orbitals obtained from 1s orbitals are shown in
Figure 11.1 and their energies in Figure 11.2. The bonding orbital is designated ɐଵ• and the
antibonding ɐଵ• ‫ כ‬, where σ signifies that the orbital is symmetrical around the internuclear

axis, as would be expected from combining s orbitals. The ground state of an ଶା ion is
therefore ɐଵଵ• .

Figure 11.1 Molecular orbitals obtained by combining 1s orbitals


[Interactive Student Tutorial]

48
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

‫כ‬
ɐଵ•
Energy 1sA 1sB
ɐଵ•

Figure 11.2 Energies of molecular orbitals obtained


by combining 1s orbitals

Hydrogen molecule and other Period 1 homonuclear diatomic molecules and molecular ions
The hydrogen molecule, like the helium atom, can be treated in the orbital approximation
[equation (10.3)]. The ground state is then ɐଵ• ଶ . The molecule thus has two bonding

electrons, twice as many as ଶା . Their respective bond numbers are therefore 1 and ½.
This is reflected in their experimental dissociation energies, 4.48 and 2.65 eV respectively.

There is a similar correlation for other Period 1 homonuclear diatomic molecules and their
ions (Table 11.1).

Need help with your


dissertation?
Get in-depth feedback & advice from experts in your
topic area. Find out what you can do to improve
the quality of your dissertation!

Get Help Now

Go to www.helpmyassignment.co.uk for more info

49
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

Molecule Configuration Number D/eV

ଶା ɐଵଵ• ½ 2.65

ଶ ଶ
ɐଵ• 1 4.48

ଶି ଶ ‫כ‬ଵ
ɐଵ• ɐଵ• ½

‡ା

ଶ ‫כ‬ଵ
ɐଵ• ɐଵ• ½ 2.47

‡ଶ ଶ ‫כ‬ଶ
ɐଵ• ɐଵ• 0 0.00

Table 11.1 Configurations and bond numbers of Period 1 homonuclear diatomic


molecules and their ions compared with their dissociation energies (D)

Period 2 homonuclear diatomic molecules and molecular ions


Other homonuclear diatomic molecules and molecular ions can be treated by combining
atomic orbitals. For Period 2 molecules, the 1s orbitals are inner orbitals, so overlap between
them is small. As we have seen, they combine to give ɐଵ• and ɐଵ• ‫ כ‬molecular orbitals, but

these will differ little in energy. For Period 2 molecules, they will be full, and contribute
no net bonding.

The next atomic orbitals are 2s and 2p. To a first approximation, 2s orbitals combine to
give ɐଶ• and ɐ‫כ‬ଶ• orbitals, 2pz orbitals (z being along the internuclear axis) combine to give
‫ כ‬orbitals, and 2p and 2p orbitals combine to give Ɏ ‫כ‬ ‫כ‬
ɐଶ’ and ɐଶ’ x y ଶ’ೣ, Ɏଶ’೤, Ɏଶ’ೣ, Ɏଶ’೤ and
orbitals. The Ɏ orbitals are formed by p orbitals overlapping sideways and are symmetrical
above and below a plane containing the nuclei. Because the p orbitals overlap sideways, Ɏ
bonds are weaker than σ bonds. The resulting energy-level diagram is shown in Figure 11.3.

Figure 11.3 Molecular orbitals formed from 2s and 2p atomic


orbitals (first approximation, holding for elements from O to Ne)
[Bobcat Chemistry]

50
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

A complication is that both 2s and 2pz orbitals give σ orbitals, and, for the early elements
in Period 2, their energies are close. Consequently, lower energies are obtained by taking
linear combinations of these orbitals, and using the variation principle to determine the
coefficients. The results are shown in Figure 11.4.

Figure 11.4 Molecular orbitals for Li2 to Ne2


[Open Text (modified)]

From Figure 11.4, we can assign ground-state configurations to the molecules from Li2 to Ne2
as shown in Table 11.2. Also in the table are their bond numbers and dissociation energies,
showing a high degree of correlation between them. The last column gives the number of
unpaired electrons. Species with unpaired electrons are paramagnetic. The theory correctly
predicts that B2 and O2 have this property. The theory is therefore an improvement on that
behind Figure 11.3, which wrongly predicts that B2 is diamagnetic and C2 paramagnetic.

51
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

Molecule Configuration Number D/eV Unpaired

‹ଶ ɐଶଶ• 1 1.06 0

‡ଶ ɐଶଶ• ɐ‫כ‬ଶ


ଶ• 0 0.61 0

ଵ ଵ
ଶ ɐଶଶ• ɐ‫כ‬ଶ
ଶ• Ɏʹ’ೣ Ɏʹ’೤ 1 ~3.0 2

ଶ ଶ
ଶ ɐଶଶ• ɐ‫כ‬ଶ
ଶ• Ɏʹ’ೣ Ɏʹ’೤ 2 6.32 0

ଶ ଶ
ଶ ɐଶଶ• ɐ‫כ‬ଶ ଶ
ଶ• Ɏʹ’ೣ Ɏʹ’೤ ɐଶ’ 3 9.76 0

ଶ ଶ ଶ
ଶ ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଵ ‫כ‬ଵ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ Ɏଶ’೤ 2 5.11 2

ଶ ଶ ଶ
ଶ ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଶ ‫כ‬ଶ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ Ɏଶ’೤ 1 1.60 0

ଶ ଶ ଶ
‡ଶ ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଶ ‫כ‬ଶ ‫כ‬ଶ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ Ɏଶ’೤ ɐଶ’ 0 0.00 0

Table 11.2 Configurations and bond numbers of Period 2 homonuclear


diatomic molecules compared with their dissociation energies (D)

Brain power By 2020, wind could provide one-tenth of our planet’s


electricity needs. Already today, SKF’s innovative know-
how is crucial to running a large proportion of the
world’s wind turbines.
Up to 25 % of the generating costs relate to mainte-
nance. These can be reduced dramatically thanks to our
systems for on-line condition monitoring and automatic
lubrication. We help make it more economical to create
cleaner, cheaper energy out of thin air.
By sharing our experience, expertise, and creativity,
industries can boost performance beyond expectations.
Therefore we need the best employees who can
meet this challenge!

The Power of Knowledge Engineering

Plug into The Power of Knowledge Engineering.


Visit us at www.skf.com/knowledge

52
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

A similar table can be drawn up for the corresponding ions. Of interest are those occurring
in compounds. These are listed in Table 11.3.

Ion Configuration Number D/eV Unpaired

ଶ ଶ
ଶଶି ɐଶଶ• ɐ‫כ‬ଶ ଶ
ଶ• Ɏʹ’ೣ Ɏʹ’೤ ɐଶ’ 3 (8.69)a 0

ଶ ଶ ଶ
ଶା ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଵ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ 2½ 6.48 1

ଶ ଶ ଶ
ଶି ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଶ ‫כ‬ଵ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ Ɏଶ’೤ 1½ 4.09 1

ଶ ଶ ଶ
ଶି
ଶ ɐଶଶ• ɐ‫כ‬ଶ ‫כ‬ଶ ‫כ‬ଶ
ଶ• ɐଶ’ Ɏʹ’ೣ Ɏʹ’೤ Ɏଶ’ೣ Ɏଶ’೤ 1 (1.49)b 0

Table 11.3 Configurations and bond numbers of some Period 2 homonuclear diatomic molecular ions
compared with their dissociation energies
a
Bond energy in H2C2. b Bond energy in H2O2.

Hydrogen-helium cation
The hydrogen-helium cation, HHe+, is isoelectronic with the hydrogen molecule. It is
formed in discharge tubes containing hydrogen and helium. It differs from the hydrogen
molecule in having nuclei with different charges. This means that we cannot put ܿଶ equal
to ܿଶ in equation (11.5). Instead we have to express the energy in terms of ܿ and ܿ and
use the variation method to get the best value. A simple alternative is to substitute equation
(11.4) into equation (6.18), multiply the result by ߮ or ߮ , and integrate over all the
coordinates. Multiplication by ߮ gives

ܿ ߙ ൅ ܿ ߚ ൌ ‫ ܧ‬ሺܿ ൅ ܵܿ ሻ(11.11)

Multiplication by ߮ gives

ܿ ߚ ൅ ܿ ߙ ൌ ‫ܧ‬ሺܵܿ ൅ ܿ ሻ(11.12)

These give

௖ ఈ ିா
 ఉିாௌ
ൌ െ ఉିாௌ ൌ െఈ (11.13)
௖  ିா

from which

ሺߙ െ ‫ ܧ‬ሻሺߙ െ ‫ ܧ‬ሻ ൌ ሺߚ െ ‫ܵܧ‬ሻଶ(11.14)

53
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

Equation (11.14) can be readily solved in the special case that ߙ ൌ ߙ. This then gives

ሺߙ െ ‫ܧ‬ሻଶ ൌ ሺߚ െ ‫ܵܧ‬ሻଶ(11.15)

from which equations (11.9) and (11.10) follow.

To see what happens when ߙ ് ߙ, we let B be more electronegative than A, and set (by
way of illustration and to simplify the arithmetic) ߙ ൌ  െ͵ units, ߙ ൌ െ͸ units, ߚ ൌ െʹ
units, and ܵ ൎ Ͳ. Equation (11.14) is then

ሺെ͵ െ ‫ ܧ‬ሻሺെ͸ െ ‫ ܧ‬ሻ ൌ ሺʹሻଶ(11.16)

the solutions of which are ‫ ܧ‬ൌ െ͹ units and െʹ—‹–•


units. Substituting these values into equation
௖ ଵ
(11.13) gives ൌ ൅ʹ and െ ଶ respectively. The normalized wave-functions are therefore
௖
ఝ ାଶఝ
߮ା ൎ (11.17)
ξହ

and
ଶఝ ିఝ
߮ି ൎ (11.18)
ξହ

respectively. The first solution thus corresponds to the bonding orbital and is concentrated
on B; the second solution corresponds to the antibonding orbital, and is concentrated on
A. The bond is accordingly polar and the molecule has a dipole moment. The energy levels
are as shown in Figure 11.5.

ɐ‫כ‬
߮A
Energy
߮B

ɐ
Figure 11.5 Energy levels arising by combining s
orbitals having different energies

An approximate value for the dipole moment in our illustrative example can be obtained
as follows. From equation (11.17) and ߮ ߮ ൎ Ͳ,

ఝమ ାସఝమ(11.19)
ߩା ൎ ହ
ଵ ସ
The fraction of time an electron spends on atom A is therefore and on B, . The difference
ହ ହ
for two electrons gives the dipole moment of the molecule:

ߤ ൎ ʹ ቀହቁ ݁‫(ݎ‬11.20)

54
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Modified Schrödinger theory of molecules

Other heteronuclear diatomic molecules


The treatment of HHe+ can be extended to other heteronuclear diatomic molecules. Of
particular interest to chemists are CN‒ and NO+ ions. These are isoelectronic with N2 and
contain polar triple bonds.

Polyatomic molecules
Polyatomic molecules can be treated in the same way as diatomic molecules, by fixing the
positions of the nuclei, and combining atomic orbitals centred on these nuclei. The result
is an energy-level diagram like that in Figure 11.6.

======
(antibonding)
======
C 2p
‒‒‒‒‒‒‒
(antibonding)
‒‒‒‒‒‒‒
C 2s
Energy
======
4H 1s
======
(bonding)

‒‒‒‒‒‒‒
(bonding)

Figure 11.6 Molecular-orbital energy-level diagram for methane. The eight valence electrons
occupy the bonding orbitals.

55
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Valence-bond (VB) theory

12 VALENCE-BOND (VB) THEORY


This is an earlier theory than molecular orbital theory described above, and is used particularly
in describing the properties and reactions of organic compounds.

Hydrogen molecule
VB theory supposes that the bond in the H2 molecule is formed by an electron in the 1s
orbital of hydrogen atom A pairing up with an electron in the 1s orbital of B. Since the
electrons are indistinguishable, this gives

߰ ൎ ܰሾ߮ ሺͳሻ߮ ሺʹሻ ൅ ߮ ሺʹሻ߮ ሺͳሻሿ(12.1)

where
߰ ൎ ܰሾ߮ is ሺthe
ͳሻ߮normalization
 ሺʹሻ ൅ ߮ ሺʹሻ߮ ሺͳሻሿ
constant. To satisfy the general Pauli principle, this wave
function is coupled with spin function ߪି [equation (10.19)].

56
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Valence-bond (VB) theory

Equation (12.1) may be compared with the corresponding MO expression [from equation
(11.4)]:

߰ ൎ ܰሾ߮ ሺͳሻ ൅ ߮ ሺͳሻሿሾ߮ ሺʹሻ ൅ ߮ ሺʹሻሿ(12.2)

This differs from equation (12.1) in having terms in ߮ ሺͳሻ߮ ሺʹሻ and ߮ ሺͳሻ߮ ሺʹሻ , i.e. it
gives more weight to both electrons being on the same atom.

Water molecule
In terms of the orbitals ʹ’௫, ʹ’௬, and ʹ’௭ (Sect. 7), the outer configuration of an oxygen
atom is ʹ• ଶʹ’ଵ௫ ʹ’ଵ௬ ʹ’ଶ௭. The atom can therefore form electron-pair bonds with two hydrogen
atoms, ʹ’ଵ௫ coupling with the 1s electron of one and ʹ’ଵ௬ coupling with the 1s electron of
the other. Since ʹ’௫ and ʹ’௬ are 90° apart, this predicts an H—O—H angle of 90°. The
observed angle is 104.5°, showing that the treatment is only approximate.

Methane molecule
A carbon atom has the outer configuration ʹ• ଶ ʹ’ଵ௫ ʹ’ଵ௬ . By the reasoning of the last section,
we might expect it to form CH2. To describe CH4, it is necessary, first, to promote a 2s
electron to the 2p shell to give the atom the configuration ʹ•ଵʹ’ଵ௫ ʹ’ଵ௬ ʹ’ଵ௭ . The orbitals have
then to be combined in such a way as to give four identical orbitals (h) directed towards
the corners of a tetrahedron. These combinations are:


Šଵ ൌ ሺ•൅’௫ ൅ ’௬ ൅ ’௭ ሻ(12.3)


Šଶ ൌ ሺ•ԟ’௫ ԟ ’௬ ൅ ’௭ ሻ(12.4)


Šଷ ൌ ଶ ሺ•ԟ’௫ ൅ ’௬ ԟ’௭ ሻ(12.5)


Šସ ൌ ଶ ሺ•൅’௫ ԟ’௬ ԟ’௭ ሻ(12.6)

These orbitals can couple with the 1s orbitals of four hydrogen atoms to form CH4. The first
process is called “promotion”, the second, which is mathematical, is called “hybridization”.
The corresponding energy level diagram is shown in Figure 12.1, and may be compared
with Figure 11.6. Both approaches have their merits. Figure 12.1 brings out the equivalence
of the bonds; Figure 11.6 ties in with the photoelectron spectrum which has two peaks in
the appropriate energy range with intensities in the ratio 1:3.

57
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Valence-bond (VB) theory

======
(antibonding)

======
C 2p

Energy ‒‒‒‒‒‒‒
C 2s

======
4H 1s
======
(bonding)

Figure 12.1 Valence-bond energy-level diagram for methane.

Other hybrid orbitals are set out in Table 12.1.

Exercise

Show that the hybrids in equations (12.3) – (12.6) are normalized and orthogonal. You can
assume that the functions s, px, py, and pz have these properties. Use the conditions ‫ ߰ ׬‬ଶ †߬ ൌ ͳ
and ‫߰߰ ׬‬Ԣ†߬ ൌ Ͳ (Sect. 6).

Type Components Shape Angle

sp s, pz linear 180°

sp2 s, px, py trigonal planar 120°

sp3 s, px, py, pz tetrahedral 109.5°

Table 12.1 Some hybrid orbitals

Ethylene molecule
VB theory can readily be applied to other organic molecules. For example, ethylene can be
formulated with ʹ•ʹ’௫ ʹ’௬ hybrids on the carbon atoms (Table 12.1). These are at 120°
and form the ɐ bonds:

H H
C—C
H H

58
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Valence-bond (VB) theory

The ʹ’௭ orbitals are at 90° to these, and combine to form a π bond.

Benzene molecule
The Kekulé structures for benzene (German benzol) can be formulated in a similar way to
ethylene.

These structures can be mathematically combined to give a structure with a lower energy
and with all the C—C bonds the same:

߰ ൌ ܰሺ߰ଵ ൅ ߰ଶ ሻ(12.7)

Following Linus Pauling, the effect of combining valence bond structures is called “resonance”,
and the structures that are combined are called “resonance structures”.

Oxygen molecule
This can be formulated in the same way as ethylene, only with lone pairs in place of the
bonds to hydrogen. This gives a bond number of two, as does MO theory (Table 11.2).
Unlike MO theory, however, it predicts that all the electrons will be paired, whereas MO
theory predicts that two will be unpaired. The paramagnetism of oxygen confirms the latter.

59
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Ligand-field theory

13 LIGAND-FIELD THEORY
Ligand-field theory is a theory of transition-metal compounds. These have d electrons in
their valence shells. There are five d orbitals, with ݉௟ ൌ Ͳǡ േͳǡ േʹ (Sect. 7). These can be
combined to give the orbitals shown in Figure 13.1.

Figure 13.1 Diagrams showing the shapes of d orbitals


[chemeddl]

The orbitals in Figure 13.1 combine with orbitals of ligands to give molecular orbitals.
How they combine depends on the shape of the metal-ligand system. If it is octahedral, as

Challenge the way we run

EXPERIENCE THE POWER OF


FULL ENGAGEMENT…

RUN FASTER.
RUN LONGER.. READ MORE & PRE-ORDER TODAY
RUN EASIER… WWW.GAITEYE.COM

1349906_A6_4+0.indd 1 22-08-2014 12:56:57

60
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Ligand-field theory

it usually is, the d orbitals split into two groups. The orbitals †௭ మ and †௫ మି௬మ (labelled eg)
point directly towards the ligands and interact with them to give bonding and antibonding
MOs. The bonding orbitals are occupied by ligand electrons, leaving the antibonding orbitals
available for metal d electrons. The remaining d orbitals (– ଶ‰) point between the ligands,
and either do not interact with them or interact relatively weakly. The orbitals available for
metal d electrons are therefore as shown in Figure 13.2.

—— —— (from ‡‰ )
߂‘…–
—— —— —— (– ଶ‰ )

Figure 13.2 Orbitals available to metal d


electrons in an octahedral species

The splitting in Figure 13.2 is called the “ligand-field splitting”. This gives rise to a band or
bands in the absorption spectrum, from which it can be calculated. For an octahedral species,
the splitting is denoted ߂‘…– . If ߂‘…– is small, the metal d electrons distribute themselves
among the orbitals available to them in the same way as in a free metal ion in the same
oxidation state. If, however, ߂‘…– is large, the electrons fill the lower level preferentially. This
leads to the formation of high-spin or low-spin species.

Consider, for example, the ferric complexes [FeF6]3‒ and [Fe(CN)6]3‒. The corresponding

free ion (Fe3+) has five 3d electrons, which occupy the 3d orbitals singly (total spin ൌ ).

The complex [FeF6]3‒ has the same arrangement as the free ion and is high-spin. On the
other hand, the complex [Fe(CN)6]3‒ has its 3d electrons in the lower set of orbitals, leaving

only one unpaired electron (total spin ൌ ). It is therefore low-spin. This difference shows

itself in the magnetic properties of the two complexes.

Ligand-field theory rationalizes a great deal of transition-metal chemistry.

61
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Density functional theory (DFT)

14 DENSITY FUNCTIONAL
THEORY (DFT)
Calculations on many-electron systems can be simplified by making use of the theorem that
the properties of such systems are determined by the overall electron density, ߩ (Hohenberg
and Kohn 1964). This makes the energy a “functional” of ߩ, i.e. a function of a function
( ߩ being a function of position). This means that the energy can be written

‫ ܧ‬ሾߩሿ ൌ ܶሾߩሿ ൅  ܸ‡ ሾߩሿ ൅ ܸ‡‡ ሾߩሿ(14.1)

Here ܶ is the total kinetic energy of the electrons, ܸ‡ is the total energy of attraction between
the electrons and the nuclei, and the ܸ‡‡ total energy of repulsion between the electrons.
The last term can be further written

ܸ‡‡ ሾߩሿ ൌ ‫ܬ‬ሾߩሿ ൅ ‫ ܧ‬ሾߩሿ(14.2)

where ‫ ܬ‬is the classical energy of repulsion between the electrons (the Coulomb energy)
and ‫ ܧ‬covers exchange and correlation (Sect. 10).

Now according to a second Hohenberg-Kohn theorem, the lowest value of ‫ ܧ‬ሾߩሿ is the
true energy. This provides a variation method for determining ‫ܧ‬. This is simpler than the
variation method for solving the Schrödinger equation because, while ߰ depends on the
coordinates of all the electrons, ߩ only depends on ‫ݔ‬, ‫ݕ‬, and ‫ݖ‬. In what follows, I will
represent ‫ݔ‬, ‫ݕ‬, ‫ ݖ‬by ߬.

To minimize equation (14.1), we need expressions for the terms in it. For ܸ‡ and ‫ܬ‬, we
can use the classical equations

௘మ ௓ఘሺఛሻ
ܸ‡ ሾߩሿ ൌ െ ቀସɎఌ ቁ σ ‫׬‬ ௥
†߬ (14.3)

and

௘మ ଵ ఘሺఛభ ሻఘሺఛమ ሻ
‫ܬ‬ሾߩሿ ൌ ቀସɎఌ ቁ ଶ ‫׭‬ ௥భమ
†߬ଵ †߬ଶ(14.4)

In equation (14.3),  labels the nuclei and ‫ ݎ‬is the distance of any given point in the electron
density distribution from nucleus . In equation (14.4), the numbers 1 and 2 specify two
different points in the electron density distribution, a distance ‫ݎ‬ଵଶ apart.

62
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Density functional theory (DFT)

For ܶ it is usual to go back to orbitals (Kohn and Sham 1965). From equation (6.12),

௛మ ‫׏‬మ ట
ܶ ൌ െ ଼Ɏమ௠ (14.5)
‡ ట

However, workers are trying to find simpler functionals for ܶ, inspired by that for a free
electron gas:

మ ఱ

ܶሾߩሿ ൌ ଵ଴ ሺ͵Ɏଶ ሻయ ‫ߩ ׬‬య †߬ (14.6)

Finally, a suitable functional has to be found for ‫ ܧ‬The simplest is to approximate it to


that for a free electron gas:

‫ ܧ‬ሾߩሿ ൎ ‫ ߝ ׬‬ሺߩሻߩ†߬(14.7)

where ߝ is the exchange-correlation per electron in a homogeneous gas of constant density.
This is called the “local-density approximation”. This reproduces energies with an accuracy of
10‒20%, and bond lengths with an accuracy of 2%. Other functionals have been proposed,
which improve the accuracy.

This e-book
is made with SETASIGN
SetaPDF

PDF components for PHP developers

www.setasign.com

63
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Perspective on the quantum theory

15 PERSPECTIVE ON THE
QUANTUM THEORY
The quantum theory is remarkably successful. It has been tested in many ways and so far,
has passed every test. Yet the physical basis of the theory remains obscure, even after many
years of enquiry. Scientists differ widely over their understanding of it.

In particular, some scientists worry that the theory describes physical systems like a
single hydrogen atom probabilistically. They wonder whether there are “hidden variables”
determining the outcome of quantum events. I hinted at this when I introduced the idea
of a sub-electronic medium in Section 5. One sceptic was Einstein, who argued, “God does
not play dice with the universe”.

As I mentioned in Section 7, physicists have suggested other interpretations of ߰ from the


one I have presented here. I survey these in my article:

“How Do Electrons Get Across Nodes? A Problem in the


Interpretation of the Quantum Theory”,

– Journal of Chemical Education, 1990, Vol. 67, pp. 643–647.

Of particular interest is the proposal of Bohm (1952). He recast Schrödinger’s equation so


that the kinetic energy is now classical and the potential energy non-classical. The latter has
a term added to it called the “quantum potential”. The result for a hydrogen atom is that
the electron in the ground state is stationary. This outcome is quite close to Lewis’ picture
of electrons in atoms and molecules being in fixed positions and Coulomb’s law breaking
down (Sect. 2). Bohm and Vigier (1954) associate the quantum potential with the motion
of a sub-electronic fluid, but other physicists dismiss it as a construct of the mathematics.

More work on the quantum theory remains to be done!

64
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Further reading

16 FURTHER READING
There are several textbooks on the application of the quantum theory to chemistry. These
include Peter W. Atkins and Ronald S. Friedman, Molecular Quantum Mechanics, 4th edn.,
Oxford University Press, 2004; Christopher J. Cramer, Essentials of Computational Chemistry,
2nd edn., Wiley, 2004. Applications of basic theories are also given in advanced textbooks
of organic chemistry and inorganic chemistry.

65
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Questions

QUESTIONS
The questions are graded; Questions 7‒10 are more difficult.

1. The spectrum of sunlight consists of the colours of the rainbow with series of dark
lines. Some of these lines (C, F, G) satisfy equation (4.2). What are they?

[Wikipedia]


2. The potential energy of a simple harmonic oscillator is given by ܸ ൌ ଶ ݇‫ ݔ‬ଶ, where
݇ is a constant (the “force constant”). Write down the Schrödinger equation for
this system.
3. What is the electronic configuration of (a) a Co atom, (b) a Co2+ ion, (c) a Co3+ ion?
4. Describe the bonding in the acetylene molecule according to VB theory.

Free eBook on
Learning & Development
By the Chief Learning Officer of McKinsey

Download Now

66
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Questions

5. Predict the number of unpaired electrons possessed by the cobaltic complexes


[CoF6]3‒ and [Co(CN)6]3‒. You may assume that the splittings are similar to those
for the corresponding ferric complexes.
6. Calculate the dissociation energy of the ଶା ion in the LCAO approximation using
the values ߙ ൌ െͳ͵Ǥͳ eV, ߚ ൌ െͳͳǤͶ eV, and ܵ ൌ ͲǤͷͻ.

7. The function ߰ ൌ ܰ‡ି௔௫ is a solution to the Schrödinger equation in Question 2,
Ɏ † †௙
where ܽ ൌ ௛ ξ݉݇. Determine ‫ ܧ‬for this solution. [Hint: ‡௙ ൌ ‡௙ ]
†௫ †௫
8. Derive a possible wave function for a particle on a ring of radius r from equation
(6.7) for a particle in a box. Let x be distance round the circle, hence ߠ ൌ ‫ݔ‬Ȁ‫ݎ‬.
[Hint: make sure ߰ is single valued.]

ߠ ‫ݔ‬
r

9. Calculate the energy of the wave-function in Question 8.


10. The ଷା ion occurs in hydrogen discharges and interstellar space. It is triangular in
shape (angles 120°). Calculate the energy of the bonding molecular orbital formed
by adding 1s orbitals centred on the three nuclei (A, B, and C).

67
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Answers

ANSWERS
1. The lines arise from hydrogen atoms in the sun absorbing energy from the sun’s
light. Compare Table 4.2, where C, F, and G are the first three lines.
†మ ట ଼Ɏమ ௠ ͳ
2. †௫ మ + ቀ‫ܧ‬ െ ݇‫ ʹݔ‬ቁ ߰ = 0
௛మ ʹ
3. (a) [Ar]4s23d7; (b) [Ar]3d7; (c) [Ar]3d6
4. The carbon atoms are •’௭ hybridized to form the ɐ bonds; the ’௫ and ’௬ orbitals
combine to form Ɏ௫ and Ɏ௬ bonds at 90° to the ɐ bonds and to each other.
5. [CoF6]3‒, 4; [Co(CN)6]3‒, 0
6. Use equation (11.9), ‫ܧ‬ሺ ሻ ൌ  െͳ͵Ǥ͸ eV (Sect. 4), and ‫ܧ‬ሺ ା ሻ ൌ Ͳ. These give ‫ ܦ‬ൌ
‫ܧ‬ሺ ሻ ൅ ‫ܧ‬ሺ ା ሻ െ ‫ܧ‬ሺ ଶା ሻ ൎ െͳ͵Ǥ͸ െ ሺെͳͷǤͶሻ ൌ ͳǤͺ eV. The experimental value is
2.65 eV.
ଵ ଵ ௞ ଵ
7. ൌ ଶ ݄ߥ, where ߥ ൌ ଶɎ ට௠ . The general solution is ൌሺ݊ ൅ ଶሻ݄ߥ (݊ ൌ Ͳǡ ͳǡ ʹ ǥ) .
8. In the case of a one-dimensional box, we had to fit a whole number of half
wavelengths in the box (Fig. 6.1). For a particle confined to a circle, fitting a
whole number of half wavelengths into the circumference makes the wave continue
round the circle with the opposite sign, rendering ߰ ൌ …‘• ݊ߠ
double valued. To avoid this,
a whole number of full wavelengths must be fitted round the circumference. This

gives ߰ ൌ •‹ ݊ ൌ •‹ ݊ߠ with ݊ ൌ ͳǡ ʹ ǥ Another possible solution is ߰ ൌ …‘• ݊ߠ
௥ ᇲ
with ݊ ൌ Ͳǡ ͳǡ ʹ ǥ These can be combined using complex numbers to give ߰ ൌ ݁ ‹௡ ఏ
with ݊ ᇱ ൌ Ͳǡ േͳǡ േʹǡ ǥ
௡మ ௛మ
9. Use equation (6.9) with V = 0. This gives ‫ ܧ‬ൌ ଼Ɏమ ூ , where I (the moment of
inertia) = mr2.
10. The orbital is ߰ ൎ ܰሺ߮ ൅ ߮ ൅ ߮ ሻ. Substitute this into equation (6.18), multiply
ఈାଶఉ
the result by ߮, and integrate. Answer: ‫ ܧ‬ൎ ଵାଶௌ

68
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Appendix

APPENDIX

Relativity Equations
We start from equation (B1) in Box 1. This describes how the mass (m) of an object varies
with its speed (‫)ݒ‬:

௠బ
݉ൌ మ (1)
ටଵିೡమ

where ݉଴ is the rest mass and c the speed of light. Squaring this equation gives:

‫ ݒ‬ଶ ݉ଶ ൅ ܿ ଶ ݉଴ଶ ൌ ܿ ଶ ݉ଶ(2)

We shall require the differential of this:

݉‫ ݒ†ݒ‬൅ ‫ ݒ‬ଶ †݉ ൌ ܿ ଶ †݉ (3)

www.sylvania.com

We do not reinvent
the wheel we reinvent
light.
Fascinating lighting offers an infinite spectrum of
possibilities: Innovative technologies and new
markets provide both opportunities and challenges.
An environment in which your expertise is in high
demand. Enjoy the supportive working atmosphere
within our global group and benefit from international
career paths. Implement sustainable ideas in close
cooperation with other specialists and contribute to
influencing our future. Come and join us in reinventing
light every day.

Light is OSRAM

69
AN INTRODUCTION TO THE
QUANTUM THEORY FOR CHEMISTS:
AN EXPERIMENTAL APPROACH Appendix

Now according to classical mechanics, the force (F) required to increase the speed of an
object is given by:

†௣
F = rate of change of momentum = (4)
†௧

where ‫ ݌‬ൌ ݉‫ݒ‬. The corresponding increase in energy (E) is given by:

dE = force × distance = ‫(ݔ†ܨ‬5)

Combining equations (4) and (5) gives:

†௣ †௫
†‫ ܧ‬ൌ ቀ †௧ ቁ †‫ ݔ‬ൌ ቀ †௧ ቁ †‫ ݌‬ൌ ‫ ݌†ݒ‬ൌ ‫†ݒ‬ሺ݉‫ݒ‬ሻ ൌ ݉‫ ݒ†ݒ‬൅ ‫ ݒ‬ଶ †݉ (6)

Combining equations (3) and (6) thus gives:

†‫ ܧ‬ൌ ܿ ଶ †݉ (7)

Integrating this from E = 0 when m = 0 gives Einstein’s celebrated equation,

‫ ܧ‬ൌ ݉ܿ ଶ (8)

This equation can be recast by first squaring it:

‫ ܧ‬ଶ ൌ ݉ଶ ܿ ସ ൌ ݉଴ଶ ܿ ସ ൅  ሺ݉ଶ ԟ ݉଴ଶ ሻܿ ସ(9)

From equation (5),

௠మ௩ మ
݉ଶ ԟ ݉଴ଶ ൌ (10)
௖మ

Substituting this into equation (9) gives

‫ ܧ‬ଶ ൌ ݉଴ଶ ܿ ସ ൅  ݉ଶ ‫ ݒ‬ଶ ܿ ଶ ൌ ݉଴ଶ ܿ ସ ൅  ‫݌‬ଶ ܿ ଶ(11)

Since the energy E is entirely kinetic, the square root of this gives equation (8.1) in the text:

ܶ ൌ ඥ‫݌‬ଶ ܿ ଶ ൅ ݉଴ଶ ܿ ସ (12)

70

Potrebbero piacerti anche