You are on page 1of 458

Table of Contents

Cover
Title Page
Copyright
Preface
Introduction
References
1 Transition from Classical Physics to Quantum Mechanics
1.1 Description of Light as an Electromagnetic Wave
1.2 Blackbody Radiation
1.3 The Photoelectric Effect
1.4 Hydrogen Atom Absorption and Emission Spectra
1.5 Molecular Spectroscopy
1.6 Summary
References
Problems
2 Principles of Quantum Mechanics
2.1 Postulates of Quantum Mechanics
2.2 The Potential Energy and Potential Functions
2.3 Demonstration of Quantum Mechanical Principles for a
Simple, One‐Dimensional, One‐Electron Model System: The
Particle in a Box
2.4 The Particle in a Two‐Dimensional Box, the Unbound
Particle, and the Particle in a Box with Finite Energy
Barriers
2.5 Real‐World PiBs: Conjugated Polyenes, Quantum Dots,
and Quantum Cascade Lasers
References
Problems
3 Perturbation of Stationary States by Electromagnetic
Radiation
3.1 Time‐Dependent Perturbation Treatment of Stationary‐
State Systems by Electromagnetic Radiation
3.2 Dipole‐Allowed Absorption and Emission Transitions
and Selection Rules for the Particle in a Box
3.3 Einstein Coefficients for the Absorption and Emission of
Light
3.4 Lasers
References
Problems
Note
4 The Harmonic Oscillator, a Model System for the Vibrations of
Diatomic Molecules
4.1 Classical Description of a Vibrating Diatomic Model
System
4.2 The Harmonic Oscillator Schrödinger Equation, Energy
Eigenvalues, and Wavefunctions
4.3 The Transition Moment and Selection Rules for
Absorption for the Harmonic Oscillator
4.4 The Anharmonic Oscillator
4.5 Vibrational Spectroscopy of Diatomic Molecules
4.6 Summary
References
Problems
5 Vibrational Infrared and Raman Spectroscopy of Polyatomic
Molecules
5.1 Vibrational Energy of Polyatomic Molecules: Normal
Coordinates and Normal Modes of Vibration
5.2 Quantum Mechanical Description of Molecular
Vibrations in Polyatomic Molecules
5.3 Infrared Absorption Spectroscopy
5.4 Raman Spectroscopy
5.5 Selection Rules for IR and Raman Spectroscopy of
Polyatomic Molecules
5.6 Relationship between Infrared and Raman Spectra:
Chloroform
5.7 Summary: Molecular Vibrations in Science and
Technology
References
Problems
6 Rotation of Molecules and Rotational Spectroscopy
6.1 Classical Rotational Energy of Diatomic and Polyatomic
Molecules
6.2 Quantum Mechanical Description of the Angular
Momentum Operator
6.3 The Rotational Schrödinger Equation, Eigenfunctions,
and Rotational Energy Eigenvalues
6.4 Selection Rules for Rotational Transitions
6.5 Rotational Absorption (Microwave) Spectra
6.6 Rot–Vibrational Transitions
References
Problems
7 Atomic Structure: The Hydrogen Atom
7.1 The Hydrogen Atom Schrödinger Equation
7.2 Solutions of the Hydrogen Atom Schrödinger Equation
7.3 Dipole Allowed Transitions for the Hydrogen Atom
7.4 Discussion of the Hydrogen Atom Results
7.5 Electron Spin
7.6 Spatial Quantization of Angular Momentum
References
Problems
Note
8 Nuclear Magnetic Resonance (NMR) Spectroscopy
8.1 General Remarks
8.2 Review of Electron Angular Momentum and Spin
Angular Momentum
8.3 Nuclear Spin
8.4 Selection Rules, Transition Energies, Magnetization,
and Spin State Population
8.5 Chemical Shift
8.6 Multispin Systems
8.7 Pulse FT NMR Spectroscopy
References
Problems
9 Atomic Structure: Multi‐electron Systems
9.1 The Two‐electron Hamiltonian, Shielding, and Effective
Nuclear Charge
9.2 The Pauli Principle
9.3 The Aufbau Principle
9.4 Periodic Properties of Elements
9.5 Atomic Energy Levels
9.6 Atomic Spectroscopy
9.7 Atomic Spectroscopy in Analytical Chemistry
References
Problems
10 Electronic States and Spectroscopy of Polyatomic Molecules
10.1 Molecular Orbitals and Chemical Bonding in the H2+
Molecular Ion
10.2 Molecular Orbital Theory for Homonuclear Diatomic
Molecules
10.3 Term Symbols and Selection Rules for Homonuclear
Diatomic Molecules
10.4 Electronic Spectra of Diatomic Molecules
10.5 Qualitative Description of Electronic Spectra of
Polyatomic Molecules
10.6 Fluorescence Spectroscopy
10.7 Optical Activity: Electronic Circular Dichroism and
Optical Rotation
References
Problems
Note
11 Group Theory and Symmetry
11.1 Symmetry Operations and Symmetry Groups
11.2 Group Representations
11.3 Symmetry Representations of Molecular Vibrations
11.4 Symmetry‐Based Selection Rules for Dipole‐Allowed
Processes
11.5 Selection Rules for Raman Scattering
11.6 Character Tables of a Few Common Point Groups
References
Problems
Appendix 1: Constants and Conversion Factors
Appendix 2: Approximative Methods: Variation and
Perturbation Theory
A2.1 General Remarks
A2.2 Variation Method
A2.3 Time‐independent Perturbation Theory for
Nondegenerate Systems
A2.4 Detailed Example of Time‐independent Perturbation:
The Particle in a Box with a Sloped Potential Function
A2.5 Time‐dependent Perturbation of Molecular Systems by
Electromagnetic Radiation
Reference
Appendix 3: Nonlinear Spectroscopic Techniques
A3.1 General Formulation of Nonlinear Effects
A3.2 Noncoherent Nonlinear Effects: Hyper‐Raman
Spectroscopy
A3.3 Coherent Nonlinear Effects
A3.4 Epilogue
References
Appendix 4: Fourier Transform (FT) Methodology
A4.1 Introduction to Fourier Transform Spectroscopy
A4.2 Data Representation in Different Domains
A4.3 Fourier Series
A4.4 Fourier Transform
A4.5 Discrete and Fast Fourier Transform Algorithms
A4.6 FT Implementation in EXCEL or MATLAB
References
Appendix 5: Description of Spin Wavefunctions by Pauli Spin
Matrices
A5.1 The Formulation of Spin Eigenfunctions α and β as
Vectors
A5.2 Form of the Pauli Spin Matrices
A5.3 Eigenvalues of the Spin Matrices
Reference
Index
End User License Agreement

List of Tables
Chapter 1
Table 1.1 Photon energies and spectroscopic rangesa.
Chapter 5
Table 5.1 Vibrational modes and assignments for
chloroform, HCCl3.
Chapter 8
Table 8.1 Nuclearg‐factors, magnetogyric ratios, and spin
moments for some sp...
Chapter 9
Table 9.1 Symbols of states for differentl and L values.
Table 9.2 Transition, energies, term symbols, and
wavelengths of the prominen...
Chapter 10
Table 10.1 Symbols of states for differentl and L values.

List of Illustrations
Chapter 1
Figure 1.1 Description of the propagation of a linearly
polarized electromag...
Figure 1.2 (a) Plot of the intensity I radiated by a blackbody
source as a f...
Figure 1.3 Portion of the hydrogen atom emission in the
visible spectral ran...
Figure 1.4 Energy level diagram of the hydrogen atom.
Transitions between th...
Chapter 2
Figure 2.1 Potential energy functions and analytical
expressions for (a) mol...
Figure 2.2 Panel (a): Wavefunctions
for n = 1, 2, 3, 4, and 5 drawn at
the...
Figure 2.3 (a) Representation of the particle‐in‐a‐box
wavefunctions shown i...
Figure 2.4 Wavefunctions of the two‐dimensional particle in
a box for (a) n x
Figure 2.5 (a) Particle in a box with infinite potential energy
barrier. (b)...
Figure 2.6 (a) Structure of 1,6‐diphenyl‐1,3,5‐hexatriene to
be used as an e...
Figure 2.7 Absorption spectra of nanoparticles as a function
of particle siz...
Figure 2.8 (a) An individual energy well with finite barrier
height and a sl...
Chapter 3
Figure 3.1 Two state energy level diagrams used for the
discussion of time‐d...

Figure 3.2 Plot of the PiB ground‐state (trace a) and first


excited‐state (t...

Figure 3.3 Panel (a): Schematic energy level diagram of a 3‐


level system in ...
Figure 3.4 Schematic of a gas laser, consisting of the
resonator structure, ...
Chapter 4
Figure 4.1 Definition of a diatomic harmonic oscillator of
masses m 1 and m 2 ...
Figure 4.2 Quadratic potential energy function V = ½ kx 2
for a diatomic mole...
Figure 4.3 Schematic of allowed (solid arrows) and
forbidden (dashed arrows)...
Figure 4.4 Graphical representation of the orthogonality of
vibrational wave...
Figure 4.5 Potential energy function of a real diatomic
molecule with dissoc...
Figure 4.6 Comparison of energy levels for harmonic and
anharmonic oscillato...
Figure 4.7 (a) Raman spectrum of Br2. (b) Expanded region
of the fundamental...
Chapter 5
Figure 5.1 Depiction of the atomic displacement vectors q i
for the three nor...
Figure 5.2 Energy ladder diagram for the water molecule
within the harmonic ...
Figure 5.3 (a) Observed infrared absorption spectrum of
water. (b) Schematic...
Figure 5.4 Depiction of the atomic displacement vectors q i
for the four norm...
Figure 5.5 Gaussian (a) and Lorentzian (b) line profiles.
Notice that the ar...
Figure 5.6 Dispersion of the refractive index (top) within an
absorption pea...
Figure 5.7 (a) Energy level diagram for a Stokes and anti‐
Stokes Raman scatt...
Figure 5.8 (a) Raman spectrum of chloroform as a neat
liquid. (b) Expanded v...
Figure 5.9 Atomic displacement vectors for (a) the
symmetric –CCl3 stretchin...
Chapter 6
Figure 6.1 Definition of spherical polar coordinates.
Figure 6.2 Graphical representation for the condition
T(φ) = T(φ + b 2π
Figure 6.3 (a) Energy level diagram for linear rotors. (b)
Schematic rotatio...
Figure 6.4 Simulated rotational spectrum of 35Cl–F at room
temperature, usin...
Figure 6.5 Schematic of the center‐of‐mass (COM) position
in an oblate (a) a...
Figure 6.6 Energy level diagram for (a) oblate and (b)
prolate top rotors. S...
Figure 6.7 (a) Observed rot–vibrational band envelopes in
the infrared absor...
Figure 6.8 (a) Rot–vibrational energy level diagram for a
harmonic oscillato...
Figure 6.9 Simulated rot–vibrational spectral band profiles
for the deformat...
Chapter 7
Figure 7.1 (a) Plot of radial part of hydrogen wavefunctions
in units of r/a
Figure 7.2 Plot of first few spherical harmonic functions.
Notice that the
Figure 7.3 Orbital energy eigenvalues and degeneracies for
the hydrogen atom...
Figure 7.4 Radial part of the wavefunctions (dashed lines)
and radial distri...
Figure 7.5 Energy level diagram and allowed electronic
transitions for the h...
Figure 7.6 (a) Energy level diagram of the hydrogen atom
orbitals in the pre...
Figure 7.7 Spatial, or orientational quantization of the
orbital angular mom...
Chapter 8
Figure 8.1 (a) Definition of the angular momentum in terms
of radius r and l...
Figure 8.2 Energy of the α and β proton nuclear spin states
as a f...
Figure 8.3 (a) Energy level diagram for two noninteracting
spins with shield...
Figure 8.4 Spectral pattern observed for two interacting
spins at lower (a) ...
Figure 8.5 Spin–spin coupling patterns for (a) J AXX and (b)
J AXXX spin syste...
Figure 8.6 Reorientation of magnetization vector following
a 90° pulse, view...
Figure 8.7 (a) Simulated “free induction decay” (FID) and
Fourier transforme...
Chapter 9
Figure 9.1 Energy level diagram of multi‐electron atoms,
explaining the Aufb...
Figure 9.2 Ionization energies (a) and atomic radii (b) for
main group eleme...
Figure 9.3 Vector addition schemes for (a) the total orbital
angular momenta...
Figure 9.4 Simplified energy level diagram of the Li atom
and transitions in...
Chapter 10
Figure 10.1 (a) Overlap of the two 1s orbitals on nuclei a and
b. The volume...
Figure 10.2 Wavefunctions for the (a) bonding and (b)
antibonding molecular ...
Figure 10.3 (a) Energy level diagram of the MOs formed
from the overlap of 2...
Figure 10.4 Electron and spin populations in the two
MOs of the lowest‐ene...
Figure 10.5 Observed (a) and simulated (b) vibronic
transition of molecular...

Figure 10.E1 See Example 10.1 for details.


Figure 10.6 Vibronic transition between the ground
vibrational state of the ...
Figure 10.7 (a) Approximate MO energy level diagram and
UV transitions for a...
Figure 10.8 (a) Approximate molecular orbital energy level
diagram of the hi...
Figure 10.9 (a) Energy level (Jablonski) diagram for
fluorescence. (b) Energ...
Figure 10.10 Schematic diagrams representing fluorescence
(a), two‐photon fl...
Figure 10.11 (a) Left (top) and right (bottom) circularly
polarized light. (...
Figure 10.12 Relationship between ORD and CD. Notice
that the differential r...
Figure 10.13 (a) CD (top) and UV absorption (bottom)
spectra of an asymmetri...
Figure 10.14 CD (top) and UV absorption (bottom) spectra
of (a) α‐helical, (...
Figure P.1
Figure P.2
Figure P.3
Figure P.4
Figure P.5
Chapter 11
Figure 11.1 Example of a symmetry operation (C 2).
Figure 11.2 (a) Example of one of three σ ν mirror planes in
the mo...
Figure 11.3 Definition of a center of inversion, located at the
coordinate o...
Figure 11.4 Description of an improper rotation operation
(S6) for ethane.
Figure 11.5 Effects of symmetry operations E and σ yz on a
Cartesian coo...
Figure 11.6 (a) Cartesian displacement vectors for the water
molecule. (b) C...
Appendix 2
Figure A2.1 Comparison between the unperturbed (a) and
perturbed (b) particl...
Appendix 3
Figure A3.1 Schematic energy level diagram for degenerate
(a and nondegenera...
Figure A3.2 (a) Phase matching diagram for frequency
doubling (SHG). (b) Pha...
Figure A3.3 (a) Schematic energy level diagram for the
CARS process. (b) Pha...
Figure A3.4 Broadband micro‐CARS spectra of cellular
components: (a) nucleol...
Figure A3.5 Schematic diagram of FSRS. See text for details
(from ref. 7).
Appendix 4
Figure A4.1 Representation of data in different domains. (a)
Graph of the in...
Figure A4.2 (a) Intensity vs. time and (b) intensity vs.
frequency plot of a...
Figure A4.3 Approximation of a square wave function
(heavy black line) by a ...
Figure A4.4 Examples of Fourier transforms (FTs). (a) The
FT of a delta func...
Figure A4.5 Panel (a): Real part of a reverse transform of a
spectrum back t...
Quantum Mechanical
Foundations of Molecular
Spectroscopy

Max Diem
Author

Max Diem, PhD


Professor Emeritus
Department of Chemistry
Northeastern University
Laboratory of Spectral Diagnosis
Boston, MA
USA
Cover iStock 965444768 / © StationaryTraveller
Supplementary material for instructors, including a Solution Manual, available for
download from www.wiley-vch.de/textbooks
All books published by WILEY‐VCH are carefully produced. Nevertheless, authors,
editors, and publisher do not warrant the information contained in these books, including
this book, to be free of errors. Readers are advised to keep in mind that statements, data,
illustrations, procedural details or other items may inadvertently be inaccurate.
Library of Congress Card No.:
applied for
British Library Cataloguing‐in‐Publication Data
A catalogue record for this book is available from the British Library.
Bibliographic information published by the Deutsche Nationalbibliothek
The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data are available on the Internet at <http://dnb.d-nb.de>.
© 2021 WILEY‐VCH GmbH, Boschstr. 12, 69469 Weinheim, Germany
All rights reserved (including those of translation into other languages). No part of this book
may be reproduced in any form – by photoprinting, microfilm, or any other means – nor
transmitted or translated into a machine language without written permission from the
publishers. Registered names, trademarks, etc. used in this book, even when not specifically
marked as such, are not to be considered unprotected by law.
Print ISBN: 978‐3‐527‐34792‐6
ePDF ISBN: 978‐3‐527‐82961‐3
ePub ISBN: 978‐3‐527‐82960‐6
Cover Design SCHULZ Grafik‐Design, Fußgönheim, Germany
Preface
When the author took courses in quantum mechanical principles and
chemical bonding in graduate school in the early 1970s, the course
materials seldomly covered the fascinating interplay between
spectroscopy and quantum mechanics, and textbooks of these days
devoted the majority of space to derivations and mathematical
principles and the discussion of the hydrogen atom and chemical
bonding. While an understanding of these subjects is, of course, a
necessity for further study, this book emphasizes a slightly different
approach to quantum mechanics, namely, one from the viewpoint of
a spectroscopist. In this approach, the existence of stationary energy
states – either electronic, vibrational, rotational, or spin states – is
considered the fundamental concept, since spectroscopy exists
because of transitions between these states. Quantum mechanics
provides the theoretical framework for the interpretation of
experimental data. On the other hand, spectroscopic results provide
the impetus for refining theories that explain the results. Classical
physics cannot provide this framework, since the idea of stationary
energy states violates the laws of classical physics.
Thus, the approach taken here in this book is to present early on, in
Chapter 2, how the application of quantum mechanical principles
leads necessarily to the existence of stationary energy states using
the particle‐in‐a box model system. The third chapter then
introduces the concept of spectroscopic transitions between these
stationary states, using time‐dependent perturbation theory.
The following chapters are presented in order of mathematical
complexity of the Schrödinger equation that describes the problem.
The simplest case, the particle in a box, is discussed in Chapter 2.
The next subject is the simple harmonic oscillator, for which the
eigenfunctions resemble those of the particle in a box, and
transitions can be visualized in terms of the discussion in Chapters 2
and 3. In the following discussions (Chapters 5–10), vibrational,
rotational, atomic, molecular electronic, and spin spectroscopies will
be introduced. These discussions, if possible, start with a classical
description, followed by the quantum mechanical equations for
wavefunctions and eigenvalues, and the derivation of the selection
rules. These selection rules determine the form and information
content of the respective spectroscopic techniques. Although space
limitations prevent in‐depth discussions of spectroscopic
applications to complex molecular systems, all efforts have been
made to include molecular systems larger than diatomic molecules
(the level of molecular complexity where many textbooks capitulate),
since the world we live in mostly consists of more complicated
molecules than diatomics.
Thus, in Chapter 5, the concept of the harmonic oscillator (Chapter
4) will be extended to vibrational (infrared and Raman) spectroscopy
of polyatomic molecules. This chapter introduces concepts of band
shapes, lifetimes, and a quantum mechanical description of
molecular polarizability. Next in complexity are the differential
equations for a rotational molecule that leads to rotational
spectroscopy (Chapter 6). These equations will introduce the
quantum mechanical description of the angular momentum and the
energy levels of simple and more complicated molecules. The results
from the rotational Schrödinger will also be used to solve the radial
part of the hydrogen atom Schrödinger equation (Chapter 7). The
principles learned from the rotational Schrödinger equation will also
be used to introduce the spin eigenfunctions and eigenstates, a
subject that leads directly to spin spectroscopy such as nuclear
magnetic resonance (NMR), which is discussed in Chapter 8.
Next, the structure of atoms and ions containing more than one
electron will be presented. This discussion includes an introduction
to atomic spectroscopy and term symbols of electronic states.
However, since the main theme of this book is molecular
spectroscopy, this chapter only serves as an introduction to these
subjects.
Chapter 10 is devoted to electronic spectroscopy of di‐ and
polyatomic molecules. Again, as in previous chapters, it is necessary
to define the states between which electronic transitions occur. This
leads necessarily to the discussion of chemical bonding in terms of
molecular orbital theory. Chemical bonding will be discussed to the
level that electronic spectra of simple molecules can be explained,
but the interaction between vibrational and electronic wavefunctions
to produce vibronic states will be discussed in more detail to explain
fluorescence phenomena as well as some Raman effects that rely on
transitions into vibronic energy levels. Finally, Chapter 11 introduces
group theory and the symmetry properties of molecules and the
influence of symmetry on the appearance of molecular spectra.
The approach taken here in this book was strongly influenced by an
excellent textbook Physical Chemistry by Engel and Reid [1] that was
used as a required text in undergraduate physical chemistry courses
at Northeastern University. This book emphasizes the
unconventional approach taken by the early theorists who are
responsible for the field of quantum mechanics as we know it. I
gained substantial understanding of the philosophical background of
quantum mechanics from this book. What is presented here in
Quantum Mechanical Foundations of Molecular Spectroscopy is a
similar approach but with much more emphasis on molecular
spectroscopy.
Although the present book emphasizes the relationship between
spectroscopy and quantum mechanics more than other texts, the
author wishes to point out the importance of following up on some
proofs and derivations (omitted here) by studying books on “real”
quantum mechanics or quantum chemistry. In particular, the one‐
and two‐volume treatments by I. Levine [2, 3] are highly
recommended, as well as many other old and new books [4, 5].
The mathematical requirements for understanding this book do not
exceed the level achieved after a three‐semester sequence of calculus,
and all efforts have been made to provide examples and problems
that will illuminate the mathematical steps. Most importantly,
although some derivations are presented, the goal is not to lose sight
of what quantum mechanics does for spectroscopy in the
mathematical complexities.
Boston, August 2019
Literature references for the Preface are at the end of the
Introduction.
Introduction
This book, Quantum Mechanical Foundations of Molecular
Spectroscopy, is based on a graduate‐level course by the same name
that is being offered to first‐year graduate students in chemistry at
the Department of Chemistry and Chemical Biology at Northeastern
University in Boston. When I joined the faculty there in 2005, I
revised the course syllabus to emphasize the philosophical
underpinnings of quantum mechanics and introduce much more of
the quantum mechanics of molecular spectroscopy, rather than
atomic structure, chemical bonding, and what is commonly referred
to as “quantum chemistry.”
As my own appreciation of many aspects of quantum mechanics
evolved, I found it useful to start my lectures in this course with a
quote from a famous researcher and Nobel laureate (1995, for his
work on quantum electrodynamics), the late Professor Richard
Feynman, which – taken slightly out of context – reads [6]:
I think I can safely say that nobody understands quantum
mechanics.
This rather discouraging statement has to be seen from the viewpoint
that, when studying quantum mechanics, one realizes that this
theory is not based on axioms, but on postulates – a very unusual
fact in the sciences. Furthermore, it replaced deterministic results
with probabilistic answers. When exposed to these conundrums,
students will naturally ask the question: “Why bother studying
quantum mechanics, if I will not understand it anyway?” or worse,
“Is quantum mechanics for real, or is it the brainchild of some far‐
out mad scientists?” The answer here is also contained in a quote by
Feynman:
It doesn't matter how beautiful a theory is, …. If it doesn't agree
with experiment, it's wrong.
This statement could also be formulated to imply that a theory that
consistently provides answers that agree with the experiment most
likely is correct. Thus, although nobody may understand quantum
mechanics in its entirety, it gives answers that – over and over –
agree with experiments and in fact provides a mechanism and
framework for explaining the experimental results.
Quantum mechanics originated in the early decades of the twentieth
century, when it was found that some experiment results just could
not be explained by existing laws of physics and, in fact, violated
established physical dogmas. It was these results that gave rise to the
emergence of quantum mechanics that grew out of a patchwork of
ideas aimed at explaining these hitherto unexplainable experimental
results. These ideas coalesced into the field we now refer to as
quantum mechanics. This newly formulated theory was wildly
successful in explaining a myriad of physical and chemical
observations – from the shape and meaning of the periodic chart of
elements to the subject of this book, namely, the interaction of light
with matter that is the basis of spectroscopy.
While many aspects of molecular spectroscopy, such as the rotational
or vibrational energies of a molecule, can be described in classical
terms, the idea that atoms and molecules can exist in quantized,
stationary energy states is a direct result of the postulates of
quantum mechanics. Furthermore, application of the principles of
time‐dependent quantum mechanics explains how electromagnetic
radiation of the correct energy may cause a transition between these
stationary energy states and produce observable spectra. Thus, the
entire field of molecular spectroscopy is a direct result of quantum
mechanics and represents the experimental results that confirms the
theory. The phenomenal growth of all forms of spectroscopy over the
past eight decades has contributed enormously to our understanding
of molecular structure and properties. What started as simple
molecular spectroscopy such as infrared and Raman vibrational
spectroscopy, (microwave) rotational spectroscopy, ultraviolet–
visible absorption, and emission spectroscopy has now bloomed into
a very broad field that includes, for example, the modern magnetic
resonance techniques (including medical magnetic resonance
imaging); nonlinear, laser, and fiber‐based spectroscopy; surface and
surface‐enhanced spectroscopy; pico‐ and femtosecond time‐
resolved spectroscopy, and many more. Spectroscopy is embedded as
a major component in material science, chemistry, physics, and
biology and other branches of scientific and engineering endeavors.
Thus, the quantum mechanical underpinnings of spectroscopy are a
major subject that need to be understood in the pursuit of scientific
efforts.

References
1 Engel, T. and Reid, P. (2010). Physical Chemistry, 2e. Upper
Saddle River, NJ: Pearson Prentice Hall.
2 Levine, I. (1970). Quantum Chemistry, vol. I&II. Boston: Allyn &
Bacon.
3 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.
4 Kauzman, W. (1957). Quantum Chemistry. New York: Academic
Press.
5 Eyring, H., Walter, J., and Kimball, G.E. (1967). Quantum
Chemistry. New Yrok: Wiley.
6 Feynman, R. (1964). Probability and Uncertainty: The Quantum
Mechanical View of Nature ‐ The Character of Physical Law
1964. Cornell University.
1
Transition from Classical Physics to Quantum
Mechanics
At the end of the nineteenth century, classical physics had progressed
to such a level that many scientists thought all problems in physical
science had been solved or were about to be solved. After all, classical
Newtonian mechanics was able to predict the motions of celestial
bodies, electromagnetism was described by Maxwell's equations (for a
review of Maxwell's equations, see [1]), the formulation of the
principles of thermodynamics had led to the understanding of the
interconversion of heat and work and the limitations of this
interconversion, and classical optics allowed the design and
construction of scientific instruments such as the telescope and the
microscope, both of which had advanced the understanding of the
physical world around us.
In chemistry, an experimentally derived classification of elements had
been achieved (the rudimentary periodic table), although the nature
of atoms and molecules and the concept of the electron's involvement
in chemical reactions had not been realized. The experiments by
Rutherford demonstrated that the atom consisted of very small,
positively charged, and heavy nuclei that identify each element and
electrons orbiting the nuclei that provided the negative charge to
produce electrically neutral atoms. At this point, the question
naturally arose: Why don't the electrons fall into the nucleus, given
the fact that opposite electric charges do attract? A planetary‐like
situation where the electrons are held in orbits by centrifugal forces
was not plausible because of the (radiative) energy loss an orbiting
electron would experience. This dilemma was one of the causes for
the development of quantum mechanics.
In addition, there were other experimental results that could not be
explained by classical physics and needed the development of new
theoretical concepts, for example, the inability of classical models to
reproduce the blackbody emission curve, the photoelectric effect, and
the observation of spectral “lines” in the emission (or absorption)
spectra of atomic hydrogen. These experimental results dated back to
the first decade of the twentieth century and caused a nearly explosive
reaction by theoretical physicists in the 1920s that led to the
formulation of quantum mechanics. The names of these physicists –
Planck, Heisenberg, Einstein, Bohr, Born, de Broglie, Dirac, Pauli,
Schrödinger, and others – have become indelibly linked to new
theoretical models that revolutionized physics and chemistry.
This development of quantum theory occupied hundreds of
publications and letters and thousands of pages of printed material
and cannot be covered here in this book. Therefore, this book
presents many of the difficult theoretical derivations as mere facts,
without proof or even the underlying thought processes, since the aim
of the discussion in the following chapters is the application of the
quantum mechanical principles to molecular spectroscopy. Thus,
these discussions should be construed as a guide to twenty‐first‐
century students toward acceptance of quantum mechanical
principles for their work that involves molecular spectroscopy.
Before the three cornerstone experiments that ushered in quantum
mechanics – Planck's blackbody emission curve, the photoelectric
effect, and the observation of spectral “lines” in the hydrogen atomic
spectra – will be discussed, electromagnetic radiation, or light, will be
introduced at the level of a wave model of light, which was the
prevalent way to look at this phenomenon before the twentieth
century.

1.1 Description of Light as an


Electromagnetic Wave
As mentioned above, the description of electromagnetic radiation in
terms of Maxwell's equation was published in the early 1860s. The
solution of these differential equations described light as a transverse
wave of electric and magnetic fields. In the absence of charge and
current, such a wave, propagating in vacuum in the positive z‐
direction, can be described by the following equations:

(1.1)
(1.2)

where the electric field and the magnetic field are perpendicular
to each other, as shown in Figure 1.1, and oscillate in phase at the
angular frequency

(1.3)
where ν is the frequency of the oscillation, measured in units of s−1 =
Hz. In Eqs. (1.1) and (1.2), k is the wave vector (or momentum vector)
of the electromagnetic wave, defined by Eq. (1.4):

(1.4)

Here, λ is the wavelength of the radiation, measured in units of


length, and is defined by the distance between two consecutive peaks
(or troughs) of the electric or magnetic fields. Vector quantities, such
as the electric and magnetic fields, are indicated by an arrow over the
symbol or by bold typeface.
Since light is a wave, it exhibits properties such as constructive and
destructive interference. Thus, when light impinges on a narrow slit,
it shows a diffraction pattern similar to that of a plain water wave that
falls on a barrier with a narrow aperture. These wave properties of
light were well known, and therefore, light was considered to exhibit
wave properties only, as predicted by Maxwell's equation.
Figure 1.1 Description of the propagation of a linearly polarized
electromagnetic wave as oscillation of electric ( ) and magnetic ( )
fields.
In general, any wave motion can be characterized by its wavelength λ,
its frequency ν, and its propagation speed. For light in vacuum, this
propagation speed is the velocity of light c (c = 2.998 × 108 m/s). (For
a list of constants used and their numeric value, see Appendix 1.) In
the context of the discussion in the following chapters, the interaction
of light with matter will be described as the force exerted by the
electric field on the charged particles, atoms, and molecules (see
Chapter 3). This interaction causes a translation of charge. This
description leads to the concept of the “electric transition moment,”
which will be used as the basic quantity to describe the likelihood
(that is, the intensity) of spectral transition.
In other forms of optical spectroscopy (for example, for all
manifestations of optical activity, see Chapter 10), the magnetic
transition moment must be considered as well. This interaction leads
to a coupled translation and rotation of charge, which imparts a
helical motion of charge. This helical motion is the hallmark of optical
activity, since, by definition, a helix can be left‐ or right‐handed.

1.2 Blackbody Radiation


From the viewpoint of a spectroscopist, electromagnetic radiation is
produced by atoms or molecules undergoing transitions between well‐
defined stationary states. This view obviously does not include the
creation of radio waves or other long‐wave phenomena, for example,
in standard antennas in radio technology, but describes ultraviolet,
visible, and infrared radiation, which are the main subjects of this
book. The atomic line spectra that are employed in analytical
chemistry, for example, in a hollow cathode lamp used in atomic
absorption spectroscopy, are due to transitions between electronic
energy states of gaseous metal atoms.
The light created by the hot filament in a standard light bulb is
another example of light emitted by (metal) atoms. However, here,
one needs to deal with a broad distribution of highly excited atoms,
and the description of this so‐called blackbody radiation was one of
the first steps in understanding the quantization of light.
Any material at a temperature T will radiate electromagnetic radiation
according to the blackbody equations. The term “blackbody” refers to
an idealized emitter of electromagnetic radiation with intensity I(λ, T)
or radiation density ρ(T, ν) as a function of wavelength and
temperature. At the beginning of the twentieth century, it was not
possible to describe the experimentally obtained blackbody emission
profile by classical physical models. This profile was shown in
Figure 1.2 for several temperatures between 1000 and 5000 K as a
function of wavelength.
Figure 1.2 (a) Plot of the intensity I radiated by a blackbody source
as a function of wavelength and temperature. (b) Plot of the radiation
density of a blackbody source as a function of frequency and
temperature. The dashed line represents this radiation density
according to Eq. (1.5).
M. Planck attempted to reproduce the observed emission profile using
classical theory, based on atomic dipole oscillators (nuclei and
electrons) in motion. These efforts revealed that the radiation density
ρ emitted by a classical blackbody into a frequency band dν as
function of ν and T would be given by Eq. (1.5):

(1.5)

where the Boltzmann constant k = 1.381 × 10−23 [J/K]. This result


indicated that the total energy radiated by a blackbody according to
this “classical” model would increase with ν2 as shown by the dashed
curve in Figure 1.2b. If this equation were correct, any temperature of
a material above absolute zero would be impossible, since any
material above 0 K would emit radiation according to Eq. (1.5), and
the total energy emitted would be unrestricted and approach infinity.
Particularly, toward higher frequency, more and more radiation
would be emitted, and the blackbody would cool instantaneously to 0
K. Thus, any temperature above 0 K would be impossible. (For a more
detailed discussion on this “ultraviolet catastrophe,” see Engel and
Reid [2].)
This is, of course, in contradiction with experimental results and was
addressed by M. Planck (1901) who solved this conundrum by
introducing the term 1/(ehν/kT − 1) into the blackbody equation, where
h is Planck's constant:

(1.6)

The shape of the modified blackbody emission profile given by Eq.


(1.6) is in agreement with experimental results. The new term
introduced by Planck is basically an exponential decay function,
which forces the overall response profile to approach zero at high
frequency. The numerator of the exponential expression contains the
quantity hν, where h is Planck's constant (h = 6.626 × 10−34 Js). This
numerator implies that light exists as “quanta” of light, or light
particles (photons) with energy E:

(1.7)

This, in itself, was a revolutionary thought since the wave properties


of light had been established more than two centuries earlier and had
been described in the late 1800s by Maxwell's equations in terms of
electric and magnetic field contributions. Here arose for the first time
the realization that two different descriptions of light, in terms of
waves and particles, were appropriate depending on what questions
were asked. A similar “particle–wave duality” was later postulated
and confirmed for matter as well (see below). Thus, the work by
Planck very early in the twentieth century is truly the birth of the
ideas resulting in the formulation of quantum mechanics.

Incidentally, the form of the expression or is fairly

common‐place in classical physical chemistry. It compares the energy


of an event, for example, a molecule leaving the liquid for the gaseous
phase, with the energy content of the surroundings. For example, the
vapor pressure of a pure liquid depends on a term , where
ΔHvap is the enthalpy of vaporization of the liquid, and RT = NkT is
the energy at temperature T, R is the gas constant, and N is
Avogadro's number. Similarly, the dependence of the reaction rate
constant and the equilibrium constant on temperature is given by
equivalent expressions that contain the activation energy or the
reaction enthalpy, respectively, in the numerator of the exponent. In
Eq. (1.6), the photon energy is divided by the energy content of the
material emitting the photon and provides a likelihood of this event
occurring.
Figure 1.2 shows that the overall emitted energy increases with
increasing temperature and that the peak wavelength of maximum
intensity shifts toward lower wavelength (Wien's law). The total
energy W radiated by a blackbody per unit area and unit time into a
solid angle (the irradiance), integrated over all wavelengths, is
proportional to the absolute temperature to the fourth power:

(1.8)

(Stefan–Boltzmann law)

The irradiance is expressed in units of or .

The implication of the aforementioned wave–particle duality will be


discussed in the next section.

1.3 The Photoelectric Effect


In 1905, Einstein reported experimental results that further
demonstrated the energy quantization of light. In the photoelectric
experiment, light of variable color (frequency) illuminated a
photocathode contained in an evacuated tube. An anode in the same
tube was connected externally to the cathode through a current meter
and a source of electric potential (such as a battery). Since the cathode
and anode were separated by vacuum, no current was observed,
unless light with a frequency above a threshold frequency was
illuminating the photocathode. Einstein correctly concluded that light
particles, or photons, with a frequency above this threshold value had
sufficient kinetic energy to knock out electrons from the metal atoms
of the photocathode. These “photoelectrons” left the metal surface
with a kinetic energy given by

(1.9)

where ϕ is the work function, or the energy required to remove an


electron from metal atoms. This energy basically is the atoms'
ionization energy multiplied by Avogadro's number. Furthermore,
Einstein reported that the photocurrent produced by the irradiation
of the photocathode was proportional to the intensity of light, or the
number of photons, but that increasing the intensity of light that had
a frequency below the threshold did not produce any photocurrent.
This provided further proof of Eq. (1.9).
This experiment further demonstrated that light has particle
character with the kinetic energy of the photons given by Eq. (1.7),
which led to the concept of wave–particle duality of light. Later, de
Broglie theorized that the momentum p of a photon was given by

(1.10)
Equation (1.10) is known as the de Broglie equation. The wave–
particle duality was later (1927) confirmed to be true for moving
masses as well by the electron diffraction experiment of Davisson and
Germer [3]. In this experiment, a beam of electrons was diffracted by
an atomic lattice and produced a distinct interference pattern that
suggested that the moving electrons exhibited wave properties. The
particle–wave duality of both photons and moving matter can be
summarized as follows.
For photons, the wave properties are manifested by diffraction
experiments and summarized by Maxwell's equation. As for all wave
propagation, the velocity of light, c, is related to wavelength λ and
frequency ν by

(1.11)
with c = 2.998 × 108 [m/s] and λ expressed in [m] and ν expressed in
[Hz = s−1]. The quantity is referred to as the wavenumber of
radiation (in units of m−1 or cm−1) that indicates how many wave
cycles occur per unit length:

(1.12)
The (kinetic) energy of a photon is given by

(1.13)
with ħ = h/2π and ω, the angular frequency, defined before as ω =
2πν.
From the classical definition of the momentum of matter and light,
respectively,

(1.14)
it follows that the photon mass is given by

(1.15)

Notice that a photon can only move at the velocity of light and the
photon mass can only be defined at the velocity c. Therefore, a photon
has zero rest mass, m0.
Particles of matter, on the other hand, have a nonzero rest mass,
commonly referred to as their mass. This mass, however, is a function
of velocity v and should be referred to as mν, which is given by

(1.16)

Example 1.1 Calculation of the mass of an electron moving at 99.0


% of the velocity of light (such velocities can easily be reached in a
synchrotron).
Answer:
According to Eq. (1.16), the mass mv of an electron at ν = 0.99 c is

(E1.1.1)

The electron at 99 % of the velocity of light has a mass of about seven


times its rest mass.

Equation (1.16) demonstrates that the mass of any matter particle will
reach infinity when accelerated to the velocity of light. Their kinetic
energy at velocity ν (far from the velocity of light) is given by the
classical expression

(1.17)

The discussion of the last paragraphs demonstrates that at the


beginning of the twentieth century, experimental evidence was
amassed that pointed to the necessity to redefine some aspects of
classical physics. The next of these experiments that led to the
formulation of quantum mechanics was the observation of “spectral
lines” in the absorption and emission spectra of the hydrogen atom.

1.4 Hydrogen Atom Absorption and


Emission Spectra
Between the last decades of the nineteenth century and the first
decade of the twentieth century, several researchers discovered that
hydrogen atoms, produced in gas discharge lamps, emit light at
discrete colors, rather than as a broad continuum of light as observed
for a blackbody (Figure 1.2a). These emissions occur in the ultraviolet,
visible, and near‐infrared spectral regions, and a portion of such an
emission spectrum is shown schematically in Figure 1.3. These
observations predate the efforts discussed in the previous two
sections and therefore may be considered the most influential in the
development of the connection between spectroscopy and quantum
mechanics.

Figure 1.3 Portion of the hydrogen atom emission in the visible


spectral range, represented as a “line spectrum” and schematically as
an emission spectrum.
These experiments demonstrated that the H atom can exist in certain
“energy states” or “stationary states.” These states can undergo a
process that is referred to as a “transition.” When the atom undergoes
such a transition from a higher or more excited state to a lower or less
excited state, the energy difference between the states is emitted as a
photon with an energy corresponding to the energy difference
between the states:

(1.18)

where the subscript f and i denote, respectively, the final and initial
(energy) state of the atom (or molecule). Such a process is referred to
as a “emission” of a photon. Similarly, an absorption process is one in
which the atom undergoes a transition from a lower to a higher
energy state, the energy difference being provided by a photon that is
annihilated in the process. Absorption and emission processes are
collectively referred to as “transitions” between stationary states and
are directly related to the annihilation and creation, respectively, of a
photon.
The wavelengths or energies from the hydrogen emission or
absorption experiments were fit by an empirical equation known as
the Rydberg equation, which gave the energy “states” of the hydrogen
atom as

(1.19)
In this equation, n is an integer (>0) “quantum” number, and Ry is
the Rydberg constant, (Ry = 2.179 × 10−18 J). This equation implies
that the energy of the hydrogen atom cannot assume arbitrary energy
values, but only “quantized” levels, E(n). This observation led to the
ideas of electrons in stationary planetary orbits around the nucleus,
which – however – was in contradiction with existing knowledge of
electrodynamics, as discussed in the beginning of this chapter.
The energy level diagram described by Eq. (1.19) is depicted in
Figure 1.4. Here, the sign convention is as follows. For n = ∞, the
energy of interaction between nucleus and electron is zero, since the
electron is no longer associated with the nucleus. The lowest energy
state is given by n = 1, which corresponds to the H atom in its ground
state that has a negative energy of 2.179 × 10−18 J.
Figure 1.4 Energy level diagram of the hydrogen atom. Transitions
between the energy levels are indicated by vertical lines.
Equation (1.19) provided a background framework to explain the
hydrogen atom emission spectrum. According to Eq. (1.19), the
energy of a photon, or the energy difference of the atomic energy
levels, between any two states nf and ni can be written as

(1.20)

At this point, an example may be appropriate to demonstrate how this


empirically derived equation predicts the energy, wavelength, and
wavenumber of light emitted by hydrogen atoms. This example also
introduces a common problem, namely, that of units. Although there
is an international agreement about what units (the système
international, or SI units) are to be used to describe spectral
transitions, the problem is that few people are using them. In this
book, all efforts will be made to use SI units, or at least give the
conversion to other units.
The sign conventions used here are similar to those in
thermodynamics where a process with a final energy state lower than
that of the initial state is called an “exothermic” process, where heat
or energy is lost. In Example 1.2, the energy is lost as a photon and is
called an emission transition. When describing an absorption process,
the energy difference of the atom is negative, ΔEatom < 0, that is, the
atom has gained energy (“endothermic” process in thermodynamics).
Following the procedure outlined in Example 1.2 would lead to a
negative wavelength of the photon, which of course is physically
meaningless, and one has to remember that the negative ΔEatom
implies the absorption of a photon.

Example 1.2 Calculation of the energy, frequency, wavelength, and


wavenumber of a photon emitted by a hydrogen atom undergoing a
transition from n = 6 to n = 2.
Answer:
The energy difference between the two states of the hydrogen atom is
given by

(E1.2.1)

Using the value of the Rydberg constant given above, Ry = 2.179 ×


10−18 J, the energy difference is

(E1.2.2)

Using Eq. (1.12), ΔE = Ephoton = hν = hc/λ, the frequency ν is found to


be

(E1.2.3)

The wavelength of such a photon is given by Eq. (1.7) as

(E1.2.4)

that is, a photon in the ultraviolet wavelength range. Finally, the


wavenumber of this photon is

(E1.2.5)

This is a case where the SI units are used infrequently, and the results
for the wavenumber are usually given by spectroscopists in units of
cm−1, where 1 m−1 = 10−2 cm−1. Accordingly, the results in Eq. E1.5 is
written as
or about 24 380 cm−1.
1.5 Molecular Spectroscopy
Example 1.2 in the previous section describes an emission process in
atomic spectroscopy, a subject covered briefly in Chapter 9.
Molecular spectroscopy is a branch of science in which the
interactions of electromagnetic radiation and molecules are studied,
where the molecules exist in quantized stationary energy states
similar to those discussed in the previous section. However, these
energy states may or may not be due to transitions of electrons into
different energy levels, but due to vibrational, rotational, or spin
energy levels. Thus, molecular spectroscopy often is classified by the
wavelength ranges of the electromagnetic radiation (for example,
microwave or infrared spectroscopies) or changes in energy levels of
the molecular systems. This is summarized in Table 1.1, and the
conversion of wavelengths and energies were discussed in Eqs. (1.11)–
(1.15) and are summarized in Appendix 1.
Table 1.1 Photon energies and spectroscopic rangesa.
νphoton λphoton Ephoton Ephoton Ephoton Transition
[J] [kJ/mol] [m−1]
Radio 750 0.4 m 5×10−25 3×10−4 2.5 NMRb
MHz
Microwave 3 GHz 10 cm 2×10−24 0.001 10 EPRb
Microwave 30 GHz 1 cm 2×10−23 0.012 100 Rotational
Infrared 3×1013 10 μm 2×10−20 12 105 Vibrational
Hz
UV/visible 1015 300 6×10−19 360 3×106 Electronic
nm
X‐ray 1018 0.3 nm 6×10−16 3.6×105 3×109 X‐ray
absorption
a) For energy conversions, see Appendix 1.
b) The resonance frequency in NMR and EPR depends on the magnetic field strength.

In this table, NMR and EPR stand for nuclear magnetic and electron
paramagnetic resonance spectroscopy, respectively. In both these
spectroscopic techniques, the transition energy of a proton or electron
spin depends on the applied magnetic field strength. All techniques
listed in this table can be described by absorption processes although
other descriptions, such as bulk magnetization in NMR, are possible
as well. As seen in Table 1.1, the photon energies are between 10−16
and 10−25 J/photon or about 10−4–105 kJ/(mol photons). Considering
that a bond energy of a typical chemical (single) bond is about 250–
400 kJ/mol, it shows that ultraviolet photons have sufficient energy
to break chemical bonds or ionize molecules. In this book, mostly low
energy photon interactions will be discussed, causing transitions in
spin states, rotational, vibrational, and electronic (vibronic) energy
levels.
Most of the spectroscopic processes discussed are absorption or
emission processes as defined by Eq. (1.18):

(1.18)

However, interactions between light and matter occur even when the
light's wavelength is different from the specific wavelength at which a
transition occurs. Thus, a classification of spectroscopy, which is more
general than that given by the wavelength range alone, would be a
resonance/off‐resonance distinction. Many of the effects described
and discussed in this book are observed as resonance interactions
where the incident light, indeed, possesses the exact energy of the
molecular transition in question. IR and UV/vis absorption
spectroscopy, microwave spectroscopy, and NMR are examples of
such resonance interactions.
The off‐resonance interactions between electromagnetic radiation and
matter give rise to well‐known phenomena such as the refractive
index of dielectric materials. These interactions arise since force is
exerted by the electromagnetic radiation on the charged particles of
matter even at off‐resonance frequencies. This force causes an
increase in the amplitude of the motion of these particles. When the
frequency of light reaches the transition energy between two states,
an effect known as anomalous dispersion of the refractive index takes
place. This anomalous dispersion of the refractive index always
accompanies an absorption process. This phenomenon makes it
possible to observe the interaction of light either in an absorption or
as a dispersion measurement, since the two effects are related to each
other by a mathematical relation known as the Kramers–Kronig
relation. This aspect will be discussed in more detail in Chapter 5.
The normal (nonresonant) Raman effect is a phenomenon that also is
best described in terms of off‐resonance models, since Raman
scattering can be excited by wavelengths that are not being absorbed
by molecules. A discussion of nonresonant effects ties together many
well‐known aspects of classical optics and spectroscopy.

1.6 Summary
The observation of the photoelectric effect and the
absorption/emission spectra of the hydrogen atom and the
modifications required to formulate the blackbody emission theory
were the triggers that forced the development of quantum mechanics.
As pointed out in the introduction, the development of quantum
mechanics is based on postulates, rather than axioms. The form of
some of these postulates can be visualized from other principles, but
their adoption as “the truth” came from the fact that they produced
the correct results.

References
1 Halliday, D. and Resnick, R. (1960). Physics. New York: Wiley.
2 Engel, T. and Reid, P. (2010). Physical Chemistry, 2e. Upper
Saddle River, NJ: Pearson Prentice Hall.
3 Davisson, C.J. and Germer, L.H. (1928). Reflection of electrons by
a crystal of nickel. Proceedings of the National Academy of
Sciences of the United States of America 14 (4): 317–322.

Problems
1. What is the maximum wavelength of electromagnetic radiation
that can ionize an H atom in the n = 2 state?
2. Why is it that any photon with a wavelength below the limiting
value obtained in (1) can ionize the H atom, whereas in standard
spectroscopy, only a photon with the correct energy can cause a
transition?
3. Assume that you carry out the experiment in (1) with light with a
wavelength of 10 nm less than calculated in (1). What is the
kinetic energy of the photoelectron created?
4. What is the velocity of the electron in Problem 3?
5. Using the de Broglie relation for matter waves, calculate the
velocity to which an electron needs to be accelerated such that its
wavelength is 10 nm.
6. What percentage of the velocity of light is the velocity in (5)?
7. What is the relativistic mass of this electron?
8. At what velocity is the wavelength of an electron 30 nm?
9. What is the momentum of such the electron in (8)?
10. What is the mass of a photon with a wavelength of 30 nm?
11. What is the momentum of the photon in (10)?
12. Compare and comment on the masses and momenta of the
moving particles in Problems (8)–(11).
13. “Frequency doubling” or “second harmonic generation (SHG)” is
a little optical trick (Appendix 3) in which two photons of the
same wavelength are squashed into a new photon, while the
energy is conserved. Calculate the wavelength of the photon
created from frequency doubling of two photons with λ = 1064
nm.
14. “Sum frequency generation (SFG)” is another optical trick
(Appendix 3) in which two photons of different wavelengths are
squashed into a new photon while the energy is conserved.
Calculate the wavelength of SFG photon created from combining
two photons with λ1 = 1064 nm and λ2 = 783 nm.
15. The value of the Rydberg constant, Ry, can be calculated
according to where e' = e/√4πεo and where mR is the
reduced mass of electron and proton. Perform an analysis of the
units of Ry.
16. Which two experiments demonstrate that light has wave and
particle character?
17. Which two experiments demonstrate that moving electrons have
wave and particle character?
2
Principles of Quantum Mechanics
Quantum mechanics presents an approach to describe the behavior
of microscopic systems. Whereas in classical mechanics the position
and momentum of a moving particle can be established
simultaneously, Heisenberg's uncertainty principle prohibits the
simultaneous determination of those two quantities. This is
manifested by Eq. (2.1):

(2.1)

which implies that the uncertainty in the momentum and position


always exceeds ħ/2, where ħ is Planck's constant divided by 2π.
Mathematically, Eq. (2.1) follows from the fact that the operators
responsible for defining position and momentum, and , do not
commutate; that is, . (This aspect will be discussed in
more detail at the end of Section 2.1.) As we shall see later (Chapter
5), the uncertainty principle also can be rewritten in terms of the
uncertainty in energy and lifetime of a spectroscopic state or in
frequency and time of a wave.
The incorporation of this uncertainty into the picture of the motion
of microscopic particles leads to discrepancies between classical and
quantum mechanics: classical physics has a deterministic outcome,
which implies that if the position and velocity (trajectory) of a
moving body are established, it is possible to predict with certainty
where it is going to be found in the future. This principle certainly
holds at the macroscopic scale: if the position and trajectory of a
macroscopic body, for example, the moon, are known, it is certainly
possible to calculate its position six days from now and to send a
spaceship to this predicted position.
Quantum mechanical systems, on the other hand, obey a
probabilistic behavior. Since the position and momentum can never
be determined simultaneously at any point in time, the position (or
momentum) in the future cannot be precisely predicted, only the
probability of either of them. This is manifested in the postulate that
all properties, present or future, of a particle are contained in a
quantity known as the wavefunction Ψ of a system. This function, in
general, depends on spatial coordinates and time; thus, for a one‐
dimensional motion (to be discussed first), the wavefunction is
written as Ψ(x, t). The probability of finding a quantum mechanical
system at any time is given by the integral of the square of this
wavefunction: ∫Ψ(x, t)2 dx. This is, in fact, one of the “postulates” on
which quantum mechanics is based to be discussed next. Different
authors list these postulates in different orders and include different
postulates necessary for the description of quantum mechanical
systems [1]. Quantum mechanics is unusual in that it is based on
postulates, whereas science, in general, is axiom‐based.

2.1 Postulates of Quantum Mechanics


Postulate 1: The state of a quantum mechanical system is completely
defined by a wavefunction Ψ(x, t). The square of this function, or in
the case of complex wavefunction, the product Ψ*(x, t) Ψ(x, t),
integrated over a volume element dτ (= dx dy dz in Cartesian
coordinates or sin2θ dθ dφ in spherical polar coordinates) gives the
probability of finding a system in the volume element dτ. Here, Ψ*(x,
t) is the complex conjugate of the function Ψ(x, t). This postulate
contains the transition from a deterministic to probabilistic
description of a quantum mechanical system. The wavefunctions
must be mathematically well behaved, that is, they must be single‐
valued, continuous, having a continuous first derivative, and
integratable (so they can be normalized).
Postulate 2: The classical linear momentum expression, p = mv, is
substituted in quantum mechanics by the differential operator ,
defined by

(2.2)
operating (or being applied to) the wavefunction Ψ(x, t). In Eq. (2.2),
i is the imaginary unit, defined by Equation (2.2) often is
considered the central postulate of QM.
The form of Eq. (2.2) can be made plausible from equations of
classical wave mechanics, de Broglie's equation (Eq. [1.10]) and
Planck's equation (Eq. [1.7]), but cannot be derived axiomatically. It
was the genius of E. Schrödinger to realize that the substitution
described in Eq. (2.2) yields differential equations that had long been
known and had solutions that agreed with experiments. In the
Schrödinger equations to be discussed explicitly in the next chapters
(for the H atom, the vibrations and rotations of molecules, and
molecular electronic energies), the classical kinetic energy T given by

(2.3)
is, therefore, substituted by

(2.4)

which is, of course, obtained by inserting Eq. (2.2) into Eq. (2.3). The
total energy of a system is given as the sum of the potential energy V
and the kinetic energy T:

(2.5)

Postulate 3: All experimental results are referred to as observables


that must be real (not imaginary or complex). An observable is
associated with (or is the “eigenvalue” of) a quantum mechanical
operator . This can be written as

(2.6)

where a are the eigenvalues and ϕ the corresponding eigenfunctions.


The terms “operator,” “eigenvalues,” and “eigenfunctions” are
terminology from linear algebra and will be further explained in
Section 2.3 where the first real eigenvalue problem, the particle in a
box, will be discussed. Notice that the eigenfunctions often are
polynomials, and each of these eigenfunctions has its corresponding
eigenvalue.
In this book, following generally accepted notations, the total energy
operator is generally identified by the symbol and referred to as
the Hamilton operator, or the Hamiltonian, of the system. With the
definition of the Hamiltonian above, it is customary to write the total
energy equation of the system as

(2.7)

Equation (2.7) implies that the energy “eigenvalues” E are obtained


by applying the operator on a set of (still unknown)
eigenfunctions ψ that are here assumed to be time‐independent and
a function of spatial coordinates x only, ψ(x). Solving the differential
equations given by Eq. (2.7) yields the eigenfunctions ψi and their
associated energy eigenvalues Ei.
Postulate 4: The expectation value of an observable a, associated
with an operator , for repeated measurements, is given by

(2.8)

If the wavefunctions Ψ(x, t) are normalized, Eq. (2.7) simplifies to

(2.9)

since the denominator in Eq. (2.8) equals 1. This expectation value


may be viewed as an expected average of many independent
measurements and embodies the probabilistic nature of quantum
mechanics.
Postulate 5: The eigenfunctions ϕi, which are the solutions of the
equation , form a complete orthogonal set of functions or,
in other words, define a vector space. This, again, will be
demonstrated in Section 2.3 for the particle‐in‐a‐box wavefunctions,
which are all orthogonal to each other and therefore may be
considered unit vectors in a vector space.
When evaluating the expectation values
(Eq. [2.9]), the functions ψ(x) may or may not be eigenfunctions of
because the real eigenfunctions ϕ(x) form a complete vector
space. Functions that are not eigenfunctions of can be written as
linear combinations of the basis functions ϕ(x). Thus, any arbitrary
wavefunction ψ of a system can be written in terms of a series
expansion of the true eigenfunctions ϕ(x) as follows:

(2.10)

The expansion coefficients an indicate how much each wavefunction


contributes to, or resembles, the true eigenfunction of the operator.
This aspect is particularly important for the approximate methods
for solving the Schrödinger equation discussed in Appendix 2.
Postulate 6: Time‐dependent systems are described by the time‐
dependent Schrödinger equation

(2.11)

where the time‐dependent wavefunctions are the product of a time‐


independent part, ψ(x), and a time evolution part:

(2.12)

We shall encounter the time‐dependent Schrödinger equation mainly


in processes where molecular systems are subject to a perturbation
by electromagnetic radiation (i.e. in spectroscopy) and shall develop
the formalism that predicts whether or not the incident radiation will
cause a transition in the molecule between two states with energy
difference ΔE = h ν = ħ ω.
Next, a simple operator/eigenvalue example will be presented to
illustrate some of the mathematical aspects.

Example 2.1 Operator/eigenvalue problem


Show that the function is an eigenfunction of the
operator , that is, show that

Answer:

(E2.1.1)

The function is an eigenfunction of the operator. The


eigenvalue c = −1.

Postulate 7: In many‐electron atoms, no two electrons can have


identically the same set of quantum numbers. This postulate is
known as the Pauli exclusion principle. It is also formulated as
follows: the product wavefunction for all electrons in an atom must
be antisymmetric with respect to interchange of two electrons. This
postulate leads to the formulation of the product wavefunction in the
form of Slater determinants (see Section 9.2) in many‐electron
systems. The value of a determinant is zero when two rows or two
columns are equal; thus, an atomic system where any electrons have
exactly the same four quantum numbers would have an undefined
product wavefunction. Furthermore, exchange of two rows (or
columns) leads to a sign change of the value of the determinant. This
last statement implies the antisymmetric property of the product
wavefunction that changes its sign upon exchange of two electrons.
Commutation of operators: Although not really a postulate of
quantum mechanics (since it follows from well‐defined mathematical
principles), a discussion of the effects of operator commutation is
included here. In physics, one often wishes to determine several
quantities simultaneously, such as the position and momentum of a
moving object or the x, y, and z components of the angular
momentum. Since Postulate 3 above states that every observable is
associated with a quantum mechanical operator, one has to
investigate the case of solving for the eigenvalues of two operators
simultaneously.
Let and be two operators such that

(2.13)

where a and b are the eigenvalues and φ and ϕ the eigenfunctions of


and , respectively. These eigenvalues can be determined
simultaneously in the same vector space if and only if the operators
commutate, that is, if the order of application of the operators on the
eigenfunction is immaterial. This commutator of two operators is
written as

(2.14)

or abbreviated as . If the operators commutate


and can be determined simultaneously; if the commutator is not
zero, then the eigenvalues cannot be determined simultaneously.
This case will be demonstrated in Example 2.2.

Example 2.2 Determine the commutator of the


momentum operator and the position operator when applied
to a function f(x), i.e. determine

(E2.2.1)
(E2.2.2)

Answer:
The derivative of the product needs to be evaluated using
the product rule of differentiation. Thus,

(E2.2.3)

(E2.2.4)

Thus, the commutator

(E2.2.5)

which predicts that the position and momentum of a moving particle


cannot be determined simultaneously. This was stated earlier in Eq.
(2.1) as the Heisenberg uncertainty principle as

(2.1)

To show the equivalency of Eqs. (E2.2.5) and (2.1), one has to


determine the standard deviations in momentum and position σp
and σx that can be related to the uncertainties Δpx and Δx.
Figure 2.1 Potential energy functions and analytical expressions
for (a) molecular vibrations and (b) an electron in the field of a
nucleus. Here, f is a force constant, k is the Coulombic constant, and
e is the electron charge.

2.2 The Potential Energy and Potential


Functions
In Postulate 2, the kinetic energy T was substituted by the operator

(2.4)

but the potential energy V was left unchanged, since it does not
include the momentum of a moving particle. The potential energy,
however, depends on the particular interactions describing the
problem, for example, the potential energy an electron experiences in
the field of a nucleus or the potential energy exerted by a chemical
bond between two vibrating nuclei. The shape of these potential
energy curves are shown in Figure 2.1 along with the potential energy
equations.
When these potential energy expressions are substituted into the
Schrödinger equation
(2.7)

one obtains a differential equation:

(2.15)

for the harmonic oscillation of a diatomic molecule and

(2.16)

for the electron in a hydrogen atom. In Eqs. (2.15) and (2.16), f and k
are constants that will be introduced later, and e is the electronic
charge, e = 1.602 × 10−19 [C]. Equation (2.16) is not strictly correct
since the potential energy is a spherical function in the distance r
from the nucleus, but is presented here and in Figure 2.1 as a one‐
dimensional quantity. Also, the mass in the denominator of the
kinetic energy operator needs to be substituted by the reduced mass
to be introduced later.
Due to the difficulties in solving equations such as Eqs. (2.15) and
(2.16), a much simpler potential energy function will be used for the
initial example of a quantum mechanical system, namely, a
rectangular box function. The ensuing particle in a box is an artificial
example but is pedagogically extremely useful and presents simple
differential equations while offering real physical applications; see
Section 2.5.

2.3 Demonstration of Quantum Mechanical


Principles for a Simple, One‐Dimensional,
One‐Electron Model System: The Particle in
a Box
Real quantum mechanical systems have the tendency to become
mathematically quite complicated due to the complexity of the
differential equations introduced in the previous section. Thus, a
simple model system will be presented here to illustrate the
principles of quantum mechanics introduced in Sections 2.1 and 2.2.
The model system to be presented is the so‐called particle in a box
(henceforth referred to as “PiB”) in which the potential energy
expression is simplified but still has with wide‐ranging analogies to
real systems. This model is very instructive, since it shows in detail
how the quantum mechanical formalism works in a situation that is
sufficiently simple to carry out the calculations step by step while
providing results that much resemble the results in a more realistic
model. This is exemplified by the overall similarity such as the
symmetry (parity) of the PiB wavefunctions when compared with
that of the harmonic oscillator wavefunctions discussed in Chapter 4.

2.3.1 Definition of the Model System


The PiB model assumes that a particle, such as an electron, is placed
into a potential energy well or confinement shown in Figure 2.2. This
confinement (the “box”) has zero potential energy for 0 ≤ x ≤ L,
where L is the length of the box. Outside the box, i.e. for x < 0 and for
x > L, the potential energy is assumed to be infinite. Thus, once the
electron is placed inside the box, it has no chance to escape, and one
knows for certain that the electron is in the box.
As discussed earlier, the total energy is written as the sum of the
kinetic and potential energies, T and V, respectively:

(2.17)
As before, the kinetic energy of the particle is given by

(2.3)

where m is the mass of the electron. Substituting the quantum


mechanical momentum operator,
(2.4)

into Eq. (2.3), the kinetic energy operator can be written as

(2.5)

Figure 2.2 Panel (a): Wavefunctions


for n = 1, 2, 3, 4, and 5 drawn at their appropriate energy levels.
Energy given in units of h2/8mL2. Panel (b): Plot of the square of the
wavefunctions shown in (a).
The potential energy inside the box is zero; thus, the total energy of
the particle inside the box is

(2.18)

Since the potential energy outside the box is infinitely high, the
electron cannot be there, and the discussion henceforth will deal
with the inside of the box. Thus, one may write the total Hamiltonian
of the system as

(2.19)

In the notation of linear algebra, this


operator/eigenvector/eigenvalue problem is written as

(2.20)

Equation (2.20) instructs to apply the Hamiltonian of Eq. (2.19) to a


set of yet unknown eigenfunctions to obtain the desired energy
eigenvalues. The eigenfunctions typically form an n‐dimensional
vector space in which the eigenvalues appear along the diagonal.
Thus, Eq. (2.20) implies

(2.21)

that is, the Hamiltonian operating on a set of eigenfunctions such


that
; ; ; and so forth that is,
of course, obtained by carrying out the matrix multiplication
indicated in Eq. (2.21).
2.3.2 Solution of the Particle‐in‐a‐Box Schrödinger
Equation
Rearranging Eqs. (2.19) and (2.20) yields

(2.22)

which is a simple differential equation that can be used to obtain the


eigenfunctions ψ(x):

(2.23)

Any functions fulfilling Eq. (2.23) must be of the form that their
second derivative equals to the original function multiplied by a
constant. For example, the function

(2.24)
could be solution of the differential Eq. (2.23),
since

(2.25)

Here, the term b2 would correspond to 2mE / ħ2, and A is a yet


undefined amplitude factor. Similarly,

(2.26)
or the sum of Eqs. (2.24) and (2.26) could be acceptable solutions.
For the time being, and for reasons that will become obvious shortly,
Eq. (2.26) will be used as a trial function to fulfill Eq. (2.23):
(2.27)

and

(2.28)

At this point, it should be pointed out that the solutions of any


differential equation depend to a large extent on the boundary
conditions: the general solution of the differential equation may or
may not describe the physical reality of the system, and it is the
boundary conditions that force the solutions to be physically
meaningful. In the case of the PiB, the boundary conditions are
determined by one of the postulates of quantum mechanics that
requires that wavefunctions are continuous. Thus, if the
wavefunction outside the box is zero (since the potential energy
outside to box is infinitely high and, therefore, the probability of
finding the particle outside the box is zero), the wavefunction inside
the box also must be zero at the boundaries of the box. Thus, one
may write the boundary conditions for the PiB differential equation
as

(2.29)
Because of these conditions, the cosine function proposed as possible
solutions (Eq. [2.24]) of Eq. (2.23) was rejected, since the cosine
function is nonzero at x = 0. Because of the required continuity at x =
L, the value of the function

must be zero at x = L as well. This can happen in two ways: The first
possibility occurs if the amplitude A is zero. This case is of no further
interest, since a zero amplitude of the wavefunction would imply that
the particle is not inside the box. The second possibility for the
wavefunction to be zero at x = L occurs if

(2.30)

Since the sine function is zero at multiples of π radians, it follows


that

(2.31)

Solving Eq. (2.31) for E yields the energy eigenvalues

(2.32)

Equation (2.32) reveals that the energy levels of the particle in a box
are quantized, that is, the energy can no longer assume any arbitrary
values, but only values of and so on. This is the
first appearance of the concept of quantized energy levels in a model
system and represents a step of enormous importance for the
understanding of quantum mechanics and spectroscopy: by
substituting the classical momentum with the momentum operator,
quantized energy levels (or stationary states) were obtained. This
quantization is a direct consequence of the boundary conditions,
which required wavefunctions to be zero at the edge of the box. Since
the energy depends on this quantum number n, one usually writes
Eq. (2.32) as

(2.33)

Substituting these energy eigenvalues back into Eq. (2.27)


(2.27)

one obtains

(2.34)

which are the wave functions for the PiB.

2.3.3 Normalization and Orthogonality of the PiB


Wavefunctions
In Eq. (2.34), “A” is an amplitude factor still undefined at this point.
To determine “A,” one argues as follows: since the square of the
wavefunction is defined as the probability of finding the particle, the
square of the wavefunction written in Eq. (2.34), integrated over the
length of the box, must be unity, since the particle is known to be in
the box. This leads to the normalization condition

(2.35)

Using the integral relationship

(2.36)

the amplitude A is obtained as follows:


(2.37)

Thus, the normalized stationary‐state wavefunctions for the particle


in a box can be written in a final form as

(2.38)

The stationary‐state (time‐independent) wavefunctions and energies


are depicted in Figure 2.2, panel (a). Although one refers to these
wavefunctions as time‐independent, they may be considered as
standing waves in which the amplitudes oscillate between the
extremes as shown in Figure 2.3 and resemble the motion of a
plugged string. Time independency then refers to the fact that the
system will stay in one of these standing wave patterns forever or
until perturbed by electromagnetic radiation.
The probability of finding the particle at any given position x is
shown in Figure 2.2, panel (b). These traces are the squares of the
wavefunctions and depict that for higher levels of n, the probability
of finding the particle moves away from the center to the periphery
of the box.
The PiB wavefunctions form an orthonormal vector space, which
implies that

(2.39)

δmn in Eq. (2.39) is referred to as the Kronecker symbol that has the
value of 1 if n = m and is zero otherwise. The wavefunctions'
normality was established above by normalizing them (Eqs. (2.36)
and (2.37)); in order to demonstrate that they are orthogonal, the
integral
(2.40)

Figure 2.3 (a) Representation of the particle‐in‐a‐box


wavefunctions shown in Figure 2.2 as standing waves. (b)
Visualization of the orthogonality of the first two PiB wavefunctions.
See text for detail.
needs to be evaluated. This can be accomplished using the integral
relationship

(2.41)

For any two adjacent wavefunction, say, m = 1 and n = 2 or m = 2 and


n = 3, the numerator of the first term in Eq. (2.41) contains the sine
function of odd multiples of π, whereas the numerator of the second
term will contain the sine function of even multiples of π. Since the
sine function of odd and even multiples of π is zero, the total integral
described by Eq. (2.41) is zero. This argument holds for any case
where n ≠ m.
This can also be visualized graphically, as shown in Figure 2.3b for
the first two PiB wavefunctions for n = 1 (curve a) and m = 2 (curve
b). When multiplied, curve c is obtained. The shaded areas above and
below the abscissa of curve c represent the integral in Eq. (2.40) for
n = 1 and m = 2 and are equal; therefore, the area under the product
curve c is zero.
Figure 2.3a also shows that the wavefunctions for the states with
quantum number larger than 1 have nodal points, or points with no
amplitude. This is familiar from classical wave behavior, for example,
for a vibrating string. Since the meaning of the squared amplitude of
the wavefunction can be visualized for the particle in a box as the
probability of finding the electron, these nodal points represent
regions in which the electron is not found.

Example 2.3

a. What is the probability P of finding a PiB in the center third of


the box for n = 1?
b. What is P for the same range for a classical particle?

Answer:

a. The probability P of finding a quantum mechanical particle–


wave is given by the square of the amplitude of the
wavefunction. Thus,

(E2.3.1)

The integral over the sin2 function can be evaluated using

(E2.3.2)
Then the probability is

(E.2.3.3)

b. A classical particle would be found with equal probability


anywhere in the box; thus, the probability of finding it in the
center third would just 1/3. Note that for higher values of n, the
probability of finding it in the center third will decrease.

2.4 The Particle in a Two‐Dimensional Box,


the Unbound Particle, and the Particle in a
Box with Finite Energy Barriers
2.4.1 Particle in a 2D Box
The principles derived in the previous section can easily by expanded
to a two‐dimensional (2D) case. Here, an electron would be confined
in a box with dimensions Lx in the x‐direction and Ly in the y‐
direction, with zero potential energy inside the box and infinitely
high potential energy outside the box:

(2.42)

The Hamiltonian for this system is

(2.43)

and the total wavefunction ψx, y can be written as


(2.44)

where A as before is an amplitude (normalization) constant. The


total energy of the system is

(2.45)

Figure 2.4 Wavefunctions of the two‐dimensional particle in a


box for (a) nx = 1 and ny = 2 and (b) nx = 2 and ny = 1.
For a square box with Lx = Ly = L, the energy expression simplifies to

(2.46)

The wavefunctions can now be represented as shown in Figure 2.4


for the cases nx = 2 and ny = 1 and nx = 1 and ny = 2. These
wavefunctions represent the standing wave on a square drum. Notice
that the energy eigenvalues for these two cases are the same:

(2.47)

When two or more energy eigenvalues for different combination of


quantum numbers are the same, these energy states are said to be
degenerate. Here, for nx = 2 and ny = 1 and nx = 1 and ny = 2, the
same energy eigenvalues are obtained; consequently, E21 and E12 are
degenerate. This is a common occurrence in quantum mechanics, as
will be seen later in the discussion of the hydrogen atom (Chapter 7),
where all the three 2p orbitals, the five 3d orbitals, and the seven 4f
orbitals are found to be degenerate.

2.4.2 The Unbound Particle


Next, the case of a system without the restriction of the boundary
conditions (an unbound particle) will be discussed. This discussion
starts with the same Hamiltonian used before:

(2.23)

When this differential equation is solved without the previously used


boundary conditions

(2.29)
the new solutions represent a particle–wave that travels along the
positive or negative x‐direction. The most general solution of the
differential Eq. (2.23) is

(2.48)
where b is a constant.
The second derivative of Eq. (2.48) is given by

(2.49)

with

(2.50)

or
(2.51)

Equation (2.51) was obtained by substituting

(2.52)

into Eq. (2.50). Thus, the unbound particle can be described by a


traveling wave (as opposed to a standing wave)

(2.53)

carrying a momentum

(2.54)

into the positive or negative x‐direction. k is the wave vector defined


in Eq. (1.6).

2.4.3 The Particle in a Box with Finite Energy Barriers


Finally, the particle in a box with a finite energy barrier, V0, will be
discussed qualitatively. This is a situation where the particle is no
longer strictly forbidden outside the confinement box and leads to
the concept of tunneling, that is, the probability of the electron found
outside the box. The shape of the potential function is shown in
Figure 2.5b.
The potential energy for this case is written as

(2.55)

and
(2.56)

(Notice that the boundaries of the box were shifted from 0 to L to


−L/2 to +L/2 for symmetry reasons that will be taken up again in
Section 3.2.) The Schrödinger equation is written in two parts: Inside
the box, where the potential energy is zero, the same equation holds
that was used earlier:

(2.23)

Outside the box, the Schrödinger equation is

(2.57)
Figure 2.5 (a) Particle in a box with infinite potential energy
barrier. (b) Particle in a box with infinite potential energy barrier.
The solutions of this equation will be of the form

(2.58)

(2.59)
where

(2.60)
For is an exponential decay function, and for
is an exponential growth function. This is shown in
Figure 2.5b to the right and left of the potential energy box,
respectively. This represents the probability of finding the electron
outside the box, a process that is known as “tunneling.” Inside the
box, the solutions of Eq. (2.23) resemble the bound wavefunctions of
the particle in a box, except that the amplitude at the boundary is no
longer zero, but must meet with the wavefunction outside the box.
This is depicted in Figure 2.5. Bound states exist for energies
E(n) < V0 only; for E(n) > V0, the electron exists as a traveling wave
as discussed before for unbound states.
The concept of tunneling may seem esoteric at first, but it has
interesting consequences. For example, a technique exists that is
known as “tunneling electron microscopy (TEM)” where a very sharp
metal tip is moved very close (within fractions of a nanometer) to the
surface of the analyte, which is at a positive potential with respect to
the metal tip. A tunneling current is observed between the tip and
the analyte that is due to electrons tunneling from the tip to the
analyte. As the substrate is moved laterally under the tip, the tip is
lowered or raised to keep the tunneling current constant. In this way,
an “image” of the morphology of the analyte can be obtained.
Tunneling may also play a role in certain chemical reactions that
depend on electron transfer from a donor to a receptor; some of
these reactions are faster than expected from computations of the
reaction rate from the activation energy. It is thought that in these
reactions, the electron may tunnel from donor to receptor at a very
fast rate. Finally, in the last example of “real‐world PIBs” in the
section below, tunneling plays a major role.

2.5 Real‐World PiBs: Conjugated Polyenes,


Quantum Dots, and Quantum Cascade
Lasers
2.5.1 Transitions in a Conjugated Polyene
Although the PiB was introduced here as a model to demonstrate
quantum mechanical principles in a mathematically manageable
system, there are several physical examples that can be treated
adequately using the PiB formalism. One of these is frequently
incorporated as an experiment in physical chemistry laboratories [2]
and involves a conjugated dye such as 1,6‐diphenyl‐1,3,5‐hexatriene,
shown in Figure 2.6a. In this molecule, the Lewis structure suggests
three double and four single bonds in the link between the two
phenyl groups. From the viewpoint of the PiB formalism, one may
consider the polyene framework the length of the box, indicated by
the straight line connecting the two phenyl groups, and the six π‐
electrons to be delocalized over the entire conjugated length and
constituting six electrons in a box in three electron pairs. A
schematic of the π‐bonding scheme is shown in Figure 2.6b. The
three electron pairs would, in this model, occupy the n = 1, n = 2, and
n = 3 levels, as indicated by the up/down arrows in each of these
levels. The absorption spectrum in the visible range shows one
absorption peak that is, in this approximation, assigned as a PiB
transition of one electron from the highest occupied molecular
orbital (HOMO) with n = 3 to the lowest unoccupied molecular
orbital (LUMO) with n = 4, as indicated by the heavy up arrow. In
Example 2.4, the wavelength of this absorption will be calculated.
This example may be a bit premature in this chapter, because it
introduced aspects of transitions between energy levels, principles of
bonding orbitals, and so forth. These subjects will be taken up in
later chapters in more detail, but this example shows quite nicely
that the PiB formalism can be applied to real systems. In some
experiments in physical chemistry laboratory, the dependence of the
absorption wavelength on the “length of the box,” that is, the
conjugated length, has been described.
Figure 2.6 (a) Structure of 1,6‐diphenyl‐1,3,5‐hexatriene to be
used as an example for the PiB calculations. (b) Energy level
diagram, based on the PiB formalism, showing the three lowest
energy levels occupied by the π‐electrons.

Example 2.4 Calculation of the energy difference between n = 3


and n = 4 energy levels for the 1,6‐diphenyl‐1,3,5‐hexatriene system,
shown in Figure 2.6, assuming that the electrons obey the particle in
a box formalism. What is the wavelength of a photon that causes this
transition?
Answer:

a. Estimation of the conjugated length. Since the single and double


bonds, with bond lengths of 154 pm and 130 pm, respectively,
are approximately 120o from each other, one can approximate
the length of the box as

(E2.4.1)
b. Calculation of the energy difference between n = 3 and n = 4.
Use me = 9.1×10−31 [kg] and h = 6.6 × 10−34 [Js] for the electron
mass and Planck's constant. Since the length of the box was
estimated to 2 significant figures, the entire computation is
carried out with 2 significant figures:

Analysis of units:

(E2.4.2)

ΔE = 3.7 × 10−19 [J]


Figure 2.7 Absorption spectra of nanoparticles as a function of
particle size. As expected, the larger particles exhibit lower energy
(longer wavelength) transitions.

c. ΔE = hc/λ or λ = hc/ΔE

(E2.4.3)

2.5.2 Quantum Dots


Certain quantum dot structures can also be modeled by a 2D particle
in a box. Quantum dots may be manufactured by creating small
circular or square semiconductor deposits on a substrate that is an
electric insulator. The electrons of the semiconductor spots are free
to move over the entire size of the dot, and the energy levels of the
free electrons follow a 2D PiB model [3]. Consequently, the color of
electronic transitions can be tuned by varying the size of the
quantum dot.
Closely related to these 2‐dimensional quantum dots are 3‐
dimensional nanoparticles, such as spheres of metallic or
semiconductor materials. The free electrons on the surface of such
spheres can assume wave patterns known as “spherical harmonic
functions” that are the solutions for particle on a sphere. Similar
functions will be discussed in Chapter 7 during the treatment of the
hydrogen atom wavefunctions. Again, the optical properties of such
nanoparticles can be tuned by adjusting the size of the nanoparticle.
This is shown in Figure 2.7.

2.5.3 Quantum Cascade Lasers


Finally, an example of a commercial application of the particle in a
box will be discussed, namely, that of a solid‐state infrared laser
known as the quantum cascade laser (QCL) [4]. In QCLs, the “box” is
constructed by creating a semiconductor “superlattice” potential
functions that mimics a PiB with finite potential energy barriers.
Furthermore, the bottom of the energy well is not flat, but at a slant,
as shown in Figure 2.8a. Both the barriers and the slant of the
bottom of the energy well can be achieved in the fabrication process
by vaporizing different composition of semiconductor materials
(doping).
Figure 2.8 (a) An individual energy well with finite barrier height
and a sloping energy bottom with the two lowest energy states and
wavefunctions. (b) The superlattice structure in a quantum cascade
laser modeled by successive PiB potential energy levels.
The slant of the energy well bottom has the effect of distorting the
PiB wavefunctions as shown exaggerated for the two lowest energy
states in Figure 2.8a. (The computation of wavefunctions in the
presence of a sloping baseline will be discussed in Appendix 2,
perturbation methods.) The distortion causes the amplitude of the
wavefunction to shift toward lower potential energy with the
consequence that an electron in the ground state of the well has a
higher probability of tunneling through the barrier, due to their
higher amplitude at the right side of the well. The “superlattice” is
formed by having a large number of these energy wells arranged as
shown in Figure 2.8b.
During the operation of the QCL, electrons are injected, via an
electric current, at a potential energy marked by the * symbol in
Figure 2.8b into a highly excited energy state and undergo a
transition as indicated by the leftmost down arrow. During this
transition, an (infrared) photon is emitted. Subsequently, the
electron in the ground state may tunnel through the finite‐height
barrier and arrives in the next quantum well and undergoes another
transition. The emission and tunneling processes are repeated as
many times as there are quantum wells in the superstructure. The
term “cascade” in QCL is due to the fact that one electron can
undergo many consecutive emission processes in the superlattice
structure. By placing the superlattice crystal into an optic cavity,
stimulated emission (see Chapter 3) from the excited states into the
ground states of each well can be achieved.

References
1 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.
2 Shoemaker, D.P., Garland, C.W., and Nibler, J.W. (1996).
Experiments in Physical Chemistry, 6e. New York: McGraw‐Hill.

3 Banin, U. et al. (1999). Identification of atomic‐like electronic


states in indium arsenide nanocrystal quantum dots. Nature 400
(6744): 542–544.
4 Faist, J. et al. (1994). Quantum Cascade Laser. Science 264
(5158): 553–556.

Problems
The following trigonometric integral relationships are needed for
these problems:
1. Show that the function f(x) = cos(bx) is an eigenfunction of the
operator d2/dx2. What is the eigenvalue?
2. Show that the function e−x2/2 is an eigenfunction of the operator
(d2/dx2) − x2. What is the eigenvalue?
3. Show that the function e−4ix is an eigenfunction of the operator
d2/dx2. What is the eigenvalue?
4. What is the probability P of finding a ground‐state PiB in the
center third of the box? What is P for the same range for a
classical particle?
5. For the PiB in the ground state, determine the expectation
values of x and px.
6. What is the expectation value of the kinetic energy operator T
for the ground‐state PiB?
7. What is the probability P of finding a particle in the first excited
state in the left half of the box with length L within the PiB
approximation?
8. Consider an electron in a one‐dimensional box with a length of
0.1 nm.
a. Calculate the energy of the 1st, 2nd, and 3rd energy levels
for this electron.
b. Calculate the wavelength of a photon required to promote
the electron from the 2nd to the 3rd energy level.
9. Describe in your own words why the particle‐in‐a‐box model
results in quantized energy levels.
10. What is quantum mechanical tunneling?
11. Calculate the commutator [Tx, px] where Tx is the kinetic energy
operator in the x‐direction and px is the momentum operator in
the x‐direction. Can the kinetic energy and the momentum be
determined simultaneously in a quantum mechanical system?
12. Show by analytical integration that ψ1(x) and ψ2(x) are
orthogonal for the one‐dimensional PiB.
13. Using a graphics program such as Excel, plot the first and
second wavefunctions for a particle in a box. Show by graphical
integration that these functions are orthogonal.
3
Perturbation of Stationary States by
Electromagnetic Radiation
In the previous chapter, the principles of quantum mechanics were
introduced at the level of an artificially conceived model system, the
particle in a box. Before investigating real situations, such as the
vibration or rotation of molecules, we shall investigate how the
simple particle in a box model system will behave when perturbed by
electromagnetic radiation and will further develop the ideas of
spectral transitions that were introduced qualitatively in
Example 2.4.
The effect of electromagnetic radiation on a quantum mechanical
system is presented in the form of perturbation theory, in general, is
discussed in Appendix A2.5. Perturbation theory is an approach used
when the quantum mechanical equations to be solved are so
complicated that they cannot be evaluated directly. In perturbation
theory, one assumes that the perturbation to the system is
sufficiently small that the solutions can be approximated by the
unperturbed wavefunctions, and an energy correction is computed
based on the unperturbed set of basis functions.

3.1 Time‐Dependent Perturbation Treatment


of Stationary‐State Systems by
Electromagnetic Radiation
Time‐independent quantum mechanics introduced in the previous
sections describes the energy expressions and wavefunctions of
stationary states, that is, states that do not change with time.
Stationary‐state wavefunctions and energies are obtained by solving
the appropriate Schrödinger equation. Next, a description is needed
that describes how a system can transition from one stationary state
to another when a perturbation, typically electromagnetic radiation,
is applied. For this, one needs to invoke the time‐dependent
Schrödinger equation, which for the one‐dimensional case is

(2.11)

When electromagnetic radiation impinges on a system described by


the stationary‐state wavefunctions, one describes the perturbation in
terms of a perturbation operator such that

(3.1)
One assumes that there exist exact eigenfunctions for the operator
and that the perturbation due to is small. If the perturbation
applied to the system is due to electromagnetic radiation, one may
write the perturbation operator as

(3.2)

In Eq. (3.2), the expression introduced previously (Eq. [1.2]) for the
electric field in electromagnetic radiation propagating in the z‐
direction is assumed. The electric field will exert a force

(3.3)
on particles with charge e. If the molecular system consists of i
charged particles found at positions xi, one defines the “electric
dipole moment” μ according to

(3.4)

and rewrites Eq. (3.2) as


(3.5)

The time‐dependent Schrödinger equation then appears as

(3.6)

and subsequently is solved with the unperturbed eigenfunctions ψ of


the operator

(3.7)

which may be, for example, the time‐independent (unperturbed)


wavefunctions of the particle in a box or the harmonic oscillator (see
Chapter 4). The time dependence of each of the wavefunctions is
introduced as follows:

(3.8)
where

(3.9)

Equation (3.8) is the general definition of a time‐dependent


wavefunction that consists of a stationary, time‐independent part
ψ(x) and the time evolution of this wavefunction given by φ(t) = eiωt.
The time‐dependent wavefunctions Ψ(x, t) of the system undergoing
a transition subsequently are expressed in terms of time‐dependent
coefficients ck(t) and the time‐dependent wavefunctions Ψ(x, t)
according to

(3.10)
The coefficients ck(t) describe the time‐dependent response of the
quantum mechanical system to the perturbation – in particular, the
change in population of the excited state in response to the
perturbation. An example may serve to illustrate the procedure
invoked so far. Consider a two‐state system in the absence of a
perturbation, as shown in Figure 3.1.

Figure 3.1 Two state energy level diagrams used for the discussion
of time‐dependent wavefunction during a transition.

The system is in the ground state and can be described


by cg(t) = 1 and ce(t) = 0 or

(3.11)
In Eq. (3.11), the subscripts “g” and “e” denote the ground and
excited states, respectively. After a perturbation is applied, the
coefficients cg(t) and ce(t) change to account for the system
undergoing a transition into the excited state that can be described
by

(3.12)
Thus, the overall time‐dependent Schrödinger equation that
accounts for the response of the system is

(3.13)

The solution of this equation can be found in many texts on quantum


mechanics [1] (vol. II, Chapter 2) and proceeds by taking the
necessary derivatives and integrating the resulting terms between
time 0 and the duration of the perturbation. This procedure yields an
expression for the time evolution of the expansion coefficients cm(t):

(3.14)

Equation (3.14) is one of the most important equations for the


understanding of spectroscopic processes since it outlines three
major features of the response of a molecule when exposed to
electromagnetic radiation.
First, it implies that the term

(3.15)

known as the transition moment must be nonzero for a transition


between the two states n and m to occur. This statement holds for
one‐photon absorption and emission processes only; later, another
transition mechanism will be introduced (Raman scattering; see
Chapter 5) that is a two‐photon process and obeys different selection
rules. The transition moment describes the action of the dipole
operator μ defined in Eq. (3.4) on the stationary‐state wavefunctions
n and m between which the transition is induced. The transition
moment, in general, determines the selection rules, depending on
the exact nature of the wavefunctions and their symmetries. This
aspect will be discussed in Section 3.2.
Second, the term containing the amplitude of the electric field, ,
indicates that the “light must be on” for a transition to occur. Third,
the part in the square brackets in Eq. (3.14) describes how the system
responds to electromagnetic radiation of different frequency or
wavelength. It is this term that prescribes that the frequency of the
light must match the energy difference between the molecular energy
levels for the transition to occur. This can be seen from the following
discussion. The second term in the square bracket in Eq. (3.14)
becomes very large at the resonance condition

(3.16)
This implies that when the frequency ω of the incident radiation is
equal to, or very close to, the energy difference ωnm between states n
and m, a transition between these states may occur, and a photon
with the corresponding energy ħω may be absorbed, if the transition
moment is nonzero.
Similarly, the first term in the square bracket in Eq. (3.14) becomes
very large if

(3.17)
This case corresponds to the situation of stimulated emission, where
a photon of the proper energy impinges onto a molecular or atomic
system in the excited state and causes this state to emit a photon,
thereby returning to the lower energy state. This time‐dependent
part of Eq. (3.14) also contains explicitly the expressions needed to
explain certain off‐resonance phenomena, such as molecular
polarizability, to be discussed in Chapter 5. The magnitude of
resonance versus off‐resonance effects (see Chapter 1) can be
estimated from the expression in square brackets as well.
Equation (3.14) holds for one‐photon absorption and emission
situations that include standard infrared (vibrational), microwave
(rotational), and visible/ultraviolet (electronic) absorption
spectroscopy. The time‐dependent part in the square bracket of Eq.
(3.14) can be summarized as

(3.18)

This equation was first mentioned in the introduction and


corresponds exactly to the condition described above by Eqs. (3.16,
3.17):

(3.19)

3.2 Dipole‐Allowed Absorption and


Emission Transitions and Selection Rules for
the Particle in a Box
The dipole moment for a transition from state m to state n, given by
the expectation value of the dipole operator,

(3.20)

must be nonzero for a transition to be allowed, and its square is


proportional to the intensity of the transition. Whether or not the
dipole transition moment is nonzero depends on the symmetry, or
parity, of the wavefunctions involved. This will be demonstrated for
the model system introduced earlier, the particle in a box.
For a one‐dimensional, one‐electron system, the dipole operator is
written as

(3.4)
where e is the electronic charge. Thus, for the transition from n = 1 to
n = 2 for the particle‐in‐a‐box wavefunctions, one needs to evaluate
the expression

(3.21)

This integral can be solved using the Mathematica or similar


software by entering
Integrate[Sin[2 Pi x /L]*x*Sin [Pi x/L], [2]]
which gave the result −8L2/9π2. After multiplication by the
normalization factor, the result given in Eq. (3.21) was obtained.
To investigate whether or not the n = 1 to n = 3 transition for the
particle in a box is allowed, the integral
Integrate[Sin[3 Pi x /L]*x*Sin[Pi x/L], [2]] = 0
was solved and gave a zero transition moment; that is, this transition
is not allowed.
Here, one encounters, for the first time, the situation that a
transition is “allowed” or “forbidden” based on the symmetry of the
transition. This can best be visualized graphically. Figure 3.2 shows a
plot of the 1st and 2nd wavefunctions for the particle in a box, along
with the transition operator that, according to Eq. (3.4), is a straight
line. Notice that in Figure 3.2, the center of the box was shifted to be
at x = 0 and the length of the box was adjusted to run from −L/2 to
L/2. This was done for symmetry reasons. The goal of the zero‐point
shift was to make the ground‐state wavefunction “symmetric” with
respect to the dotted line and the first excited‐state wavefunction
“antisymmetric” to this axis. In this view, the dipole operator is
antisymmetric as well.
Figure 3.2 Plot of the PiB ground‐state (trace a) and first excited‐
state (trace c) wavefunctions. Trace (b) represents a plot of the
transition operator. Trace (d) represents the transition moment, the
product of traces a, b, and c. The integral of this product (shaded
area under trace d) is nonzero.
In Figure 3.2, the functions ψ1 (trace a) and ψ2 (trace c) and the
transition operator (trace b) are shown. The product 〈ψ2 ∣ μ ∣ ψ1〉 is
shown by curve “d,” and the area under the curve when integrated
from −L/2 to L/2 is nonzero. This result can also be visualized from a
parity or symmetry argument: since ψ2 and the dipole operator are
odd functions, their product will be even. This product, multiplied by
the ground state ψ1, which has even parity, will result in an overall
transition moment with even parity.
By the same argument, the transition from n = 1 to n = 3 is forbidden
since the n = 3 state has even parity. Thus, the product of the dipole
operator and the ground‐state and excited‐state functions will have
odd parity, and the integral will be zero.
For the particle in a box, this leads to the following selection rules:
transitions with Δn = ±1, ±3, ±5, and so on are allowed, whereas
transitions with Δn = ±2, ±4, and so on are forbidden because the
transition moment integrals are zero. Thus, one encounters here that
transitions are allowed or forbidden depending on the symmetry of
the wavefunctions. The transition moment is the quantity that needs
to be evaluated in order to determine whether or not a transition
may occur.

3.3 Einstein Coefficients for the Absorption


and Emission of Light
When electromagnetic radiation passes from vacuum into a medium,
there will be two major interactions that are governed by the
properties of the medium: one will be a surface effect, reflection, and
refraction at the interface. The other will be an attenuation of the
intensity of the light within the medium known as absorption. Both
these effects are described by the complex refractive index, η:

(3.22)
In Eq. (3.22), n is the real refractive index used in classical optics to
describe reflection and refraction, and κ is the absorptivity discussed
in a moment. The real part of the refractive index n can be used at
conditions far from an absorption band to determine the amount of
light being reflected and refracted and determines the angle of
reflection and refraction under these conditions. In classical optics,
for example, refraction of light to describe lenses can be adequately
treated by the real part of the refractive index since glass is
transparent in the visible spectrum, and there are no nearby
absorption bands. However, even in classical optics, the achromatic
behavior of lenses results from the fact that the refractive index is not
constant with wavelength, but generally increases toward the short
wavelength region of light. This effect will be discussed in more
detail in Chapter 5 where absorption and dispersion of light will be
covered.
The absorptivity κ is related to the molar extinction coefficient ε as
follows:

(3.23)

ε is defined in Beer–Lambert law, which states that the absorbance


(attenuation) A of light in a medium is given by

(3.24)

In Eq. (3.24), C is the concentration, in mol/L, of the absorbing


species; l is the path length the light travels through the medium,
expressed in cm;1 and I0 and I are the incident light intensity and the
intensity after passing a path length l into the sample, respectively.
The Beer–Lambert law often is written in the form

(3.25)

This form better relates to the fact that, if ε is nonzero, the light
intensity decreases exponentially with the path the light travels
through the medium.
For the rest of this section and the next one on laser theory, it may be
best to visualize the stationary states as those of an atomic species,
for example, the ground state and first excited states of a noble gas
atom, such as neon, although the discussion is true for any atomic or
molecular states. Furthermore, for the discussion of the population
of the states below, let us assume that the system can only exist in
two states, |ψ1〉 and |ψ2〉. The dipole transition moment would then
be written as 〈ψ2 ∣ μ ∣ ψ1〉. However, this transition moment does not
express the rate of transitions between the two states. This rate
depends on the population of the originating state and the number of
photons impinging onto the system, in the case here the neon gas
exposed to light. The rate of absorption, i.e. the number of photons
being absorbed by the gaseous molecules undergoing a transition
from the ground state to the excited state |ψ2〉 per unit time, is given
by the expression

(3.26)
Here, N1 denotes the population of the ground state; ρ(ν12) is the
radiation density, i.e. the number of photons at frequency ν12 that
have the correct energy to cause the transition from the ground state
to the excited state; and B12 is the Einstein coefficient for absorption
given by

(3.27)

Eq. (3.26) gives a rate for the absorption of photons or the creation
of excited‐state species and therefore the depopulation of the lower
energy state. By the same token, photons of frequency ν12 may
encounter an atom in an excited state (from a previous absorption
process) and may stimulate this excited atom to release its excitation.
The rate for this “stimulated emission” process is given in complete
analogy as

(3.28)
where N2 is the population of the excited state and B21 is the Einstein
coefficient for stimulated emission. Since the squares of the
transition moments for emission and absorption are equal

(3.29)

it follows that

(3.30)
However, the rates of the absorption and emission process are vastly
different, since they depend on the populations N1 and N2 of the
ground and excited states. The ratio of the populations N2–N1 is
given by the Boltzmann equation

(3.31)

In the following example, this ratio will be calculated for two


different energy spacings between the states involved, one for a
typical electronic transition where the energy difference corresponds
to a visible photon and one for a vibrational transition that
corresponds to an infrared photon.

Example 3.1 Calculation of population ratios for a two‐state


system with an energy spacing between the energy levels of the
following:

a. ΔE = 15 000 cm−1 (=1.0 × 106 m−1) corresponding to a visible


photon with λ = 666 nm.
b. ΔE = 1500 cm−1 (=1.0 × 105 m−1) corresponding to an infrared
photon with λ = 6.66 μm at room temperature (T = 298 K).

Answer:

a. Using Eq. (3.31), with a numerical value of R = 0.698 cm−1/K−1


mol−1 (see Appendix 1), one obtains
(E3.1.1)

b. (E3.1.2)

The two population ratios are obviously different by many orders of


magnitude. While in case (a) roughly one atom out of 109 mol (!!)
would be in the excited state, in the second case, nearly 0.1 % of the
atoms are in the excited state.

As demonstrated in Example 3.1, the Boltzmann distribution


determines that the population in the lower energy state is always
higher than that in the higher energy state, that is,

(3.32)
Therefore, the rate expression given by Eq. (3.26) will always be
larger than the rate expression given by Eq. (3.28), i.e.

(3.33)
since the two Einstein coefficients are the same (see Eq. [3.30]).
The two processes introduced so far, absorption and stimulated
emission, depend on the presence of electromagnetic radiation.
However, another deactivation process, known as “spontaneous
emission,” can happen. The rate of this step depends on the
expression

(3.34)
where A21 is the Einstein coefficient for spontaneous emission that
also depends on the transition moment but is about a millionfold
smaller than B21 for energy differences corresponding to visible
photons.
Thus, the rate of light absorption is given by Eq. (3.26), whereas for
the combined deactivation process, Eq. (3.35) holds
(3.35)
The absorptivity κ, introduced in Eqs. (3.22) and (3.23), is defined in
terms of the Einstein coefficients as

(3.36)

that is, the difference of the absorption and stimulated emission rate
equations. Since the first term in the parentheses is larger than the
second term (see Eq. [3.33]), κ always will be a positive number for a
two‐state energy system discussed here. Consequently, ε also will be
positive, and, according to Eq. (3.24), light will always experience an
attenuation when passing through a medium within an absorption
region.

3.4 Lasers
The last sentence of the previous paragraph only is correct for a
system with two energy levels. However, light amplification in a
medium is possible under certain circumstances, as demonstrated by
the existence of lasers, which are very real objects. The term “laser” is
an acronym for light amplification by stimulated emission of
radiation, and this name implies that there is a way that light,
passing through a medium, is amplified, rather than attenuated. For
this to be possible, both κ and ε need to be negative such that the
exponent of Eq. (3.24) becomes positive, and the intensity of light
increases when passing through the medium.
According to Eq. (3.36), κ can only be negative if

(3.37)
that is, if a condition known as population inversion exists. This is
thermodynamically impossible in a two‐state system described by
Eq. (3.31) since it would require negative (absolute) temperatures
(which is a serious no–no).
Figure 3.3 Panel (a): Schematic energy level diagram of a 3‐level
system in which population inversion can be achieved. (b) Simplified
energy level diagram of the He–Ne laser.
The thermodynamic limitations of achieving a population inversion
can be bypassed using kinetic principles in a system that
incorporates at least three energy states. This is equivalent to certain
chemical reactions that should proceed to a thermodynamically most
stable product (“thermodynamic control”) but rather proceed to a
thermodynamically less stable product because the reaction rate is
faster (“kinetic control”). In analogy, a state may be populated by a
very fast process, but the deactivation process from this state is slow,
leading to a population inversion. This is shown in Figure 3.3a.
In the case shown in Figure 3.3a, state 3 is populated by
electromagnetic radiation at a frequency at ν31 or any other methods,
such as collisional energy transfer to be discussed later. If the
Einstein coefficient for spontaneous emission (depopulation) from
state 2 to state 1, A21, is much larger than that from state 3 to state 2,
A32, the population in state 2 is drained fast, and a population
inversion between states 3 and 2 can occur. Laser action may happen
between these states; this action is initiated when a photon at ν32 is
created to be spontaneous emission and subsequently starts to
stimulate species in state 3. Then, amplification of the radiation field
at ν32 occurs.
Among the many laser types that now exist – gas lasers, solid state
lasers, diode lasers, fiber lasers, etc., gas lasers will be used to
elaborate upon the principles of laser operation since the principles
here are most easily understood. One of the most commonly
produced and used gas lasers is the He‐Ne laser that emits a bright
red laser line at 632.8 nm.
A laser, in general, consists of a gain medium and a resonator
structure, as shown in Figure 3.4. The resonator structure consists of
mirrors that reflect the light created in the gain medium back and
forth to stimulate excited species in the gain medium to emit. The
resonator structure is often a flat and a spherical mirror to focus the
light into the gain medium. One of the mirrors may be partially
transparent to allow some of the light, typically 1 %, to escape from
the resonator and create the laser beam.

Figure 3.4 Schematic of a gas laser, consisting of the resonator


structure, defined by two mirrors and the gain medium.
In the He–Ne laser, the gain medium is a mixture of helium and
neon, at a ratio of about a 10 : 1, contained in a gas cell shown in
Figure 3.4 fitted with Brewster angle windows to minimize reflection
losses and to linearly polarize the laser output. A DC of 10–20 mA at
a few hundred volt is passed through the gas mixture that ionizes the
He atoms to produce He ions and electrons. Collisions between
electrons and He atoms cause transitions of the He atoms into a
highly energetic state. These excited He atoms, in turn, collide with
Ne atoms, promoting one of the 2p electrons into a 5s orbital; see
Figure 3.3b. This energy level corresponds to energy level 3 in
Figure 3.3a. The deactivation of the 2p53s1 states is very fast, thus
allowing the population inversion between the 2p55s1 and 2p53s1
states to build. Laser action occurs, among other transitions,
between the 2p55s1 and (still excited) 2p53s1 states. This transition is
responsible for the 632.8 nm laser line. He‐Ne lasers produce highly
monochromatic light typically in the 0.1–50 mW power regime.
Although the gain medium varies between the various laser types,
the principle is always the same in that a population inversion
between certain states is created. Light amplification occurs when a
photon, created by spontaneous emission, encounters excited species
that exist in the inverted population and stimulates these species to
emit. The photon created in this process is coherent (in phase) with
the photon that created it.

References
1 Levine, I. (1970). Quantum Chemistry, vol. I&II. Boston: Allyn &
Bacon.
2 Eyring, H., Walter, J., and Kimball, G.E. (1967). Quantum
Chemistry. New Yrok: Wiley.

Problems
1. Define the electric dipole transition moment and operator both
mathematically and descriptively.
2. Show graphically, using Excel or a similar program, that the
dipole moment operator‐mediated transition from n = 1 to n = 2
is allowed for PiB wavefunctions by plotting the PiB
wavefunction and the dipole operator. Shift the origin of the
potential well to the center of the box.
3. Certain two‐photon spectroscopic effects depend on the square
of the dipole moment operator, which for a single charged
particle would be e2x2. For a particle in a box (with the origin
shifted to the middle of the box), plot the wavefunctions and
transition operators and discuss their parity:
a. ∫ sin(x) sin(2x) dx (orthogonality)
b. ∫ sin(x) x sin(2x) dx (one‐photon dipole transition)

c. ∫ sin(x) x2 sin(2x) dx (forbidden 2‐photon 1 →2 transition)

d. ∫ sin(x) x2 sin(3x) dx (allowed 2‐photon 1 →3 transition)

4. Consider a two‐state quantum mechanical system with two


energies E2 and E1 (E2 > E1). The energy difference between the
two states is 600 cm−1.
a. At what temperature will the ratio of the population of the
higher state, n2, and the lower state, n1, be 0.3? (k = 0.698
cm−1/K)
b. What will be the ratio n2/n1 at room temperature?
5. What is the fundamental difference between spontaneous and
stimulated emission of radiation?
6. Discuss the sine qua non condition for laser action to occur.
How is this condition related to quantities that determine
absorption or emission of radiation?

Note
1 Although the unit of length should be the meter, molar extinction
coefficients are generally reported in units of L/(mol cm).
4
The Harmonic Oscillator, a Model System for
the Vibrations of Diatomic Molecules
The atoms in molecules are in constant vibrational motion that is a
manifestation of the heat content or temperature of matter. These
vibrational motions appear to be random but can be decomposed
into movement along certain coordinates, which will be discussed in
Chapter 5. The amplitude of these motions increases with increasing
temperature. Completely stopping this motion is impossible, as we
shall see in this and the following chapter, and also would violate the
third law of thermodynamics as it could create a complete order and
absolute‐zero temperature.
This chapter will introduce the next logical model system commonly
discussed in quantum mechanics, the vibrational motion of a
diatomic molecule obeying the harmonic oscillator approximation.
This model, for which the Schrödinger equation still can be solved
exactly, already is much closer to real molecular systems than the
particle‐in‐a‐box model introduced in Chapter 2 and describes the
molecular vibration of diatomic molecules adequately. Further
modification to this model (by introducing the anharmonicity)
refines it to the point where real diatomic molecules can be described
with very high accuracy.

4.1 Classical Description of a Vibrating


Diatomic Model System
The harmonic oscillator model is usually introduced for a diatomic
molecule, as shown in Figure 4.1. Here, two atoms with masses m1
and m2 are connected by a covalent chemical bond along the x‐
direction. This bond can be thought of as a spring with a restoring
force F obeying Hook's law:
(4.1)
where k is the force constant of the spring, expressed in units of
[N/m]. The potential energy V created when extending or
compressing this spring by the distant dx is obtained by integrating
the force over this distance:

(4.2)

Figure 4.1 Definition of a diatomic harmonic oscillator of masses


m1 and m2 that are at a distance x1 and x2 from the center of mass
(COM).
The shape of this potential energy curve was shown in Figure 2.1a.
The name of the model system “harmonic oscillator” is derived from
the expression in Eq. (4.2) that implies a harmonic or quadratic
potential energy function.
With Newton's second law of motion,

(4.3)

one can write the equation of motion for a diatomic molecule as

(4.4)

In Eqs. (4.2) and (4.3), mR is the reduced mass defined by


(4.5)

with m1 and m2 the individual masses of the two atoms. The reduced
mass is used to create a vibrational motion in which the center of
mass of the diatomic molecule is stationary, that is, when using the
reduced mass, the vibrational motion is decoupled from any
translational motion of the molecule.
One valid solution of the differential equation of motion Eq. (4.4) is

(4.6)
where

(4.7)

or

(4.8)

where ω is the angular frequency. Notice that for a classical


vibrational problem, the amplitude A is arbitrary but that the
frequency is defined by Eq. (4.7). This implies that for larger
amplitudes, the velocity of the motion of the particles increases but
the frequency remains constant. These equations also express the
motion of a mass suspended from a solid beam by a spring.
At this point, a quick analysis of magnitudes and units is
appropriate. The magnitude of the force constant k acting between
the atoms in a diatomic molecule such as gaseous HCl was found
experimentally to be about 450 [N/m] = 450 [kg/s2], corresponding
to a relatively stiff spring in classical mechanics. This allows us to
calculate the classical vibrational frequency of the diatomic molecule
HCl as will be demonstrated in Example 4.1.
Example 4.1 Calculation of the classical vibrational frequency of
the HCl molecule
Answer:
The reduced mass of an HCl molecule is, according to Eq. (4.5),
approximately

(E4.1.1)
Thus, the vibrational frequency for the HCl molecule is found to be

(E4.1.2)

Using the frequency to wavenumber conversion gives a value


close to the observed stretching frequency for gaseous HCl of 2.85 ×
103 cm−1 or 2.85 × 105 m−1.

Notice that, as pointed out in earlier chapters, this result implies that
a molecule such as diatomic HCl has a characteristic vibrational
frequency, but this leaves no room for the concept that
electromagnetic radiation causes a transition to a more highly
excited state. In a classical system, the energy can increase in
infinitesimally small increments by increasing the amplitude of the
vibration, whereas in the quantum mechanical and experimentally
verified situation, the energy can only increase in certain quantized
increments, leading to the absorption and annihilation of a photon.
This aspect will be discussed next.

4.2 The Harmonic Oscillator Schrödinger


Equation, Energy Eigenvalues, and
Wavefunctions
Assuming that the chemical bond in a diatomic molecule obeys
Hook's law, the vibrational Schrödinger equation for a harmonic
oscillator with one degree of freedom (x) is then

(4.9)

This follows from Eqs. (2.5) and (4.2). (In Eq. [4.9] and the following
discussion, the subscript “R” in mR for the reduced mass has been
dropped to simplify the notation.) This differential equation is
known as “Hermite's” differential equation, in which the
wavefunctions ψ(x) are the time‐independent (stationary state)
vibrational wavefunctions and E denotes the energy of the
vibrational states. Equation (4.9) is a typical operator–eigenvalue
equation notation commonly found in linear algebra. This formalism
is an instruction to operate with an operator, here, the vibrational
Hamiltonian

(4.10)

on a set of (yet unknown) functions to obtain the eigenvalues.


Substituting the eigenvalues into the trial solution and considering
the boundary conditions yield the eigenfunctions ψ(x). Detailed
methods for solving the vibrational Schrödinger Eq. (4.9) can be
found in many books on vibrational spectroscopy or quantum
chemistry textbooks [1]. Here, the approach to solve Eq. (4.9) is only
outlined to demonstrate how involved such a solution is.
Equation (4.9) is reformatted to read

(4.11)

Here, the results from Eq. (4.8) were used:

(4.12)
Equation (4.10) is solved under the assumption that for large values
of x, the following condition holds:

(4.13)

This results in a simplified equation

(4.14)

which has the approximate solutions

(4.15)

that are known as the “asymptotic solutions”where

(4.16)

Next, one assumes that the final results are of the form

(4.17)

i.e. Eq. (4.15), multiplied by a yet unknown function f(x). In Eq.


(4.17), the Gaussian part of the harmonic oscillator wavefunctions is
already in place, but the derivation of the functions f(x) is rather
complicated and involves a series expansion of f(x) according to

(4.18)

After taken the necessary derivatives required in Eq. (4.11), one


obtains
(4.19)

Here, the expansion coefficients “a” for the series expansion of f(x)
are given by the recursion formula

(4.20)

Equation (4.20) leads to two different infinite series expansions for


f(x), one with odd, the other with even expansion coefficients. Next,
one has to investigate how the boundary conditions affect these
expansion coefficients. This involves rather tricky arguments [1]
about the values of successive expansion coefficients, and it appears
that the wavefunction will become infinite as x goes to infinity. For
the wavefunctions ψ(x) to be finite, the summation in Eq. (4.19)
must be terminated at some finite value of n. Then, the numerator of
the right side of Eq. (4.20) must be zero:

(4.21)

This leads to

(4.22)

With Eq. (4.16), the energy eigenvalues

(4.23)

are obtained. Quantization of the energy in Eq. (4.22) leads to


wavefunctions that are a product of the Gaussian factor and a series
expansion in x with the expansion coefficients given by Eq. (4.20).
The resulting polynomials are referred to as the Hermite polynomials
These polynomials are related to each other by a
recursion formula (see below) that originates from Eq. (4.20).
It is very interesting to realize that the solution of a differential
equation that appears quite simple requires many steps and quite a
few assumptions, as discussed above. The author, as a spectroscopist
and not a mathematician, always marvels at the insight and
ingenuity of the mathematicians who first proposed solutions to the
equations encountered in physical chemistry. In this case, the
differential equation was solved by a nineteenth‐century French
mathematician Charles Hermite.
The vibrational wavefunctions resulting from the discussion above
then are of the form

(4.24)

where N is a normalization constant, , and

are the Hermite polynomials of order n in the variable . The


vibrational quantum number n is an integer that may take values
from zero to infinity, and α was defined in Eq. (4.16).
For simplicity's sake, setting , the Hermite polynomials
in the variable z can be written as

(4.25)

The order n of the Hermite polynomials determines the highest


power in which the variable z occurs in each polynomial. Thus, more
highly excited states (i.e. those with higher quantum number n)
correspond to Hermite polynomials with higher power of z. This
aspect is important since the power of z determines the shape of the
wavefunctions.
The higher members of the Hermite polynomials can be derived
from the recursion formula:

(4.26)

Thus, the Hermite polynomial of degree n is related to the previous


and subsequent polynomial by Eq. (4.26).

Example 4.2 Using the recursion formula and the analytical


expressions for H2(z) and H3(z), derive the form of H4(z).
Answer:
We rewrite Eq. (4.24) to read

(E4.2.1)

Then, H4(z) = 2zH3(z) − 6 H2(z) = 2z(8 z3 − 12 z) − 6(4 z2 − 2) =

(E4.2.2)

Thus, the first few vibrational wavefunctions are

(4.27)

These functions, which are plotted in Figure 4.2, form an


orthonormal vector space. Orthonormality implies (see Eqs. [2.35]–
[2.41]) that
(4.28)

Here, δij represents the Kronecker delta, which equals to one if i = j


and zero otherwise.
For example,

(4.29)

and

(4.30)

In Eq. (4.29), the integral relationship was


used.
The eigenvalues of the vibrational Schrödinger equation are given by

(4.23)
Figure 4.2 Quadratic potential energy function V = ½ kx2 for a
diatomic molecule and the resulting quantum mechanical vibrational
wavefunctions and energy states (a) and the square of the
wavefunctions (b).
In Eq. (4.23), the frequency ν is written in units of s−1 such that the
term hν has units of energy [J]. Vibrational spectroscopists, however,
prefer to use wavenumber (cf. Eq. [1.12]) as a unit of energy; thus, in
the remainder of the book, molecular vibrational energies are
expressed as

(4.31)

and the corresponding photon energy as .


The harmonic oscillator wavefunctions and energy eigenvalues are
shown in Figure 4.2a, along with the quadratic potential energy
function (Eq. [4.2]) and the energy levels corresponding to Eq.
(4.23). There are several interesting facts about the wavefunctions
and energy level plot. Firstly, Eq. (4.23) and Figure 4.3a demonstrate
that even in the vibrational ground state with n = 0, the system is not
at zero energy, but rather, at an energy

(4.32)

which is referred to as the zero‐point energy. This zero‐point energy


accounts for the fact that even in its vibrational ground state, the
atoms in a molecule undergo continuous vibrational motion. This
zero‐point vibrational energy also accounts for the 3rd law of
thermodynamics, which states that absolute zero temperature is
unattainable. This is because atomic “standstill” is impossible due to
the residual vibrational energy. This is also in line with Heisenberg's
uncertainty principle (2.1), since a vibrational amplitude of zero
would define both position and momentum simultaneously.
Secondly, Figure 4.2 also indicates some degree of “tunneling,” or a
finite probability of the oscillating system to be found outside the
potential energy curve. Thirdly, the wavefunctions are symmetric
(n = 0, 2, 4,…) or antisymmetric (n = 1, 3, 5,…) with respect to the y‐
axis in Figure 4.2a. In other words, the wavefunctions with even
quantum numbers have even parity, that is, f(x) = f(−x), while the
wavefunctions with odd quantum numbers have odd parity with f(x)
= −f(−x). This aspect will become particularly important in the
discussion of the allowed and forbidden transitions in the harmonic
oscillator approximation (see below). Fourthly, the spacing between
the energy levels is given by the classical expression

(4.7)
Figure 4.3 Schematic of allowed (solid arrows) and forbidden
(dashed arrows) absorption and emission transitions for the
harmonic oscillator.
and agrees with what is expected from classical mechanics.
Finally, the quadratic potential depicted in Figure 4.3 does not
explain bond breakage at sufficiently high energy, since the potential
function – the restoring force between the oscillating atoms –
increases steadily in the “harmonic approximation.” Therefore, the
concept of an anharmonic potential needs to be introduced (see
Section 4.4).
4.3 The Transition Moment and Selection
Rules for Absorption for the Harmonic
Oscillator
In Chapter 3, we demonstrated that three conditions are necessary
for an absorption transition to occur in an atomic or a molecular
system under the influence of a perturbation by electromagnetic
radiation. First, radiation must impinge on the molecular system (Eo
≠ 0); second, the radiation must possess the proper energy, or
frequency, corresponding to the energy difference between the
molecular or atomic states. Third, the dipole transition moment
must be nonzero:

(4.20)

Here, ψn and ψm would be the vibrational wavefunctions defined by


Eq. (4.23). Rather than evaluating the integral in Eq. (4.20) for
different values of n and m by analytical integration, the recursion
formula given in Eq. (4.25) allows a rather painless way to assess
whether or not the transition moment is zero for any values of n and
m.
As established earlier, the vibrational wavefunctions are of the form

(4.24)

The Gaussian part of the wavefunction has even parity; thus, it


does not affect the parity of the integral described by Eq. (4.20), and
the following discussion can concentrate on the parity of the Hermite
polynomials alone. Thus, the transition moment given by Eq. (4.20)
can be simplified to

(4.33)
where the factor was set to 1 since it is a constant. Recalling the
recursion formula for the Hermite polynomials,

(4.26)

the term x Hm(x) in Eq. (4.33) can be substituted by the right‐hand


side of Eq. (4.26) to yield

(4.34)

(4.35)

Since the Hermite polynomials are orthogonal, the two integrals in


Eq. (4.35) are nonzero if and only if

(4.36)

Equation (4.36) implies that transitions are allowed only if the


vibrational quantum number n changes by one unit; that is, the
selection rule for absorption (and emission) for the harmonic
oscillator is

(4.37)
Equation (4.36) also infers that

(4.38)
This selection rule implies that electric dipole transitions are allowed
only between adjacent energy levels for the harmonic oscillator. This
is shown schematically in Figure 4.3 where solid arrows indicate
allowed transitions and dashed arrows indicate forbidden
transitions. Incidentally, the transition energy between adjacent
energy levels is

(4.39)
since adjacent energy levels are equidistant and differ by the energy
obtained by the classical vibrational frequency. Thus, the classical
vibrational frequency of a harmonic oscillator given by Eq. (4.7) is
identically the same as the one predicted by the quantum mechanical
model.
As in the case of the transition moment for the particle in a box, the
transition moment (and the orthogonality of the harmonic oscillator
wavefunctions) can be demonstrated graphically as well. Figure 4.4a
demonstrates by graphical integration that the integral
is indeed zero, whereas Panel (b) shows the
same result for Thus, it can be seen easily that
the vibrational wavefunctions are orthogonal.
Figure 4.4 Graphical representation of the orthogonality of
vibrational wavefunctions and the vibrational transition moment. (a)
Product of ψ0·ψ1. The light and dark gray regions under the curves
have equal areas; thus, integration along x results in zero net area,
and the functions are orthogonal. (b) Product of ψ1·ψ2. The same
argument demonstrates that the functions are orthogonal. (c) Plot of
ψ0 (gray), ψ1 (light gray), and the dipole operator μ = e x (black). (d)
Integration of 〈ψ1 ∣ μ ∣ ψ0〉 along x‐axis yields a nonzero transition
moment.
Panels (c) and (d) show that the transition moment integral
〈ψ1∣μ∣ψ0〉 is nonzero. Panel (C) shows the two wavefunctions and the
transition operator and Panel (D) the product of the three functions.
The area under the curve is nonzero.

4.4 The Anharmonic Oscillator


The potential energy used up to this point was the harmonic
potential, which is a crude approximation to the real potential
function in a diatomic molecule. At elevated temperatures, all
molecules tend to fall apart into atoms; thus, when molecules are in
highly excited states, bond breakage will occur. The harmonic
potential used so far does not predict bond breakage ever to happen,
since the quadratic potential function increases steadily with
increasing values of x, whereas a real potential function must predict
that at a sufficiently high value of energy, bond breakage will occur.
The real potential function can be obtained by detailed quantum
mechanical calculations, in which the electronic energy is computed
as a function of the internuclear distance, and is shown schematically
in Figure 4.5. This potential energy can be approximated by the
Morse potential, given by

(4.40)

with
Figure 4.5 Potential energy function of a real diatomic molecule
with dissociation energy De.
The function has a minimum at the bond equilibrium distance xo.
When compressing the bond beyond xo, the potential energy rises
sharply due to the repulsion of the two atoms. When the bond is
elongated toward large interatomic distances, the potential function
eventually levels out, and the bond breaks. One normally defines the
potential energy at very large interatomic distances as the zero
energy (no bonding interaction takes place at large distances); thus,
the potential energy of the bond is at a negative minimum at the
equilibrium distance. The energy difference between zero potential
energy and the minimum potential energy at point xo is referred as
the bond dissociation energy, De.
Solving the quantum mechanical equations for the vibrations of a
diatomic molecule with the potential function shown in Figure 4.5
would be difficult. Thus, one approximates the shape of the potential
function V(x) in the vicinity of the potential energy minimum by a
power series expansion about the equilibrium distance:

(4.41)

V(xo) is an offset along the y‐axis and does not affect the curvature of
the potential energy. The term containing the first derivative of the
potential energy with respect to x is zero since the equilibrium
geometry corresponds to an energy minimum. The quadratic
expression in Eq. (4.41)

(4.42)

is the harmonic potential energy function used so far. The cubic term
gives the next level of approximation, and an anharmonic force
constant is defined as follows:

(4.43)

The anharmonic vibrational Schrödinger equation thus is written as

(4.44)

and is solved by perturbation methods (see Appendix 2.3). The


perturbed energy eigenvalues for an anharmonic diatomic molecule
are given by

(4.45)

with the anharmonicity constant given by

(4.46)

χ is always a positive number; thus, the energy levels of the


anharmonic case are always lowered as compared with the harmonic
oscillator; this lowering increases with the square of the quantum
number n. This is depicted in Figure 4.6, which shows a comparison
between harmonic and anharmonic oscillator energy levels.
Figure 4.6 Comparison of energy levels for harmonic and
anharmonic oscillators.
One further effect of the anharmonic potential is the fact that the
observed fundamental transition from n = 0 to n = 1 is no longer the
harmonic frequency. Within the harmonic approximation, the energy
of the n = 0 to n = 1 transition was given by

(4.47)
Taken into account the anharmonicity, the energy difference
between the ground and first excited state is given by

(4.48)

In addition to lowering the energy values, the wavefunctions –


although similar in shape to the harmonic oscillator wavefunctions –
will be shifted toward longer internuclear distances; consequently,
the symmetry of the wavefunctions changes; thus, they are no longer
symmetric or antisymmetric with respect to the xo position.
The direct consequences of the asymmetry of the wavefunction are
that the selection rules change somewhat. Whereas in the harmonic
oscillator approach transitions with

(4.49)
are strictly forbidden, these transitions are weakly allowed in the
case of an anharmonic oscillator. That implies that a weak
absorption band is observed at

(4.50)

i.e. just under twice the frequency of the fundamental transition. The
n = 0 to n = 2 transition is also referred to as the overtone (or the 1st
harmonic) of the fundamental.
In addition, the n = 0 to n = 1 transition and the n = 1 to n = 2
transition will no longer have the same energy, since the spacing
between adjacent energy levels is no longer constant:

(4.51)

Thus, the n = 1 to n = 2 transition will appear on the low frequency


side of the n = 0 to n = 1 transition with a much lower intensity,
which is determined primarily by the low population of the n = 1
state at room temperature. At elevated temperatures, the n = 1 state
becomes more populated, and therefore, the n = 1 to n = 2 transition
will become more intense. This transition is therefore referred to as a
“hot band.”
4.5 Vibrational Spectroscopy of Diatomic
Molecules
The discussion above will now be illustrated using the vibrational
spectra of selected molecules, namely, bromine, Br2, and iodine
bromide, I‐Br. The choice of good examples is actually quite hard,
since many diatomic molecules do not exhibit an infrared absorption
spectrum due to the lack of a permanent dipole moment, which is a
prerequisite for diatomic molecules to have an infrared absorption
spectrum. Since all homonuclear diatomic molecules such as H2, N2,
F2, Cl2, O2, etc. are nonpolar, the vibrational transitions are not
allowed in absorption. However, they are allowed in an alternate
method of vibrational spectroscopy, Raman scattering, which will be
discussed in detail in the next chapter. The vibrational energy levels
are the same in both forms of vibrational spectroscopy, but the
Raman process has different selection rules. Thus, homonuclear
diatomic molecules do exhibit a Raman vibrational spectrum.
Therefore, one of the examples presented below will use Raman
spectra to illustrate the energy levels and transitions for a
homonuclear diatomic species, Br2. These data are taken from the
literature [2].
Due to the high mass of the bromine atoms, the molecule has low
vibrational transition frequencies, and the fundamental stretching
frequency is observed at c. 315 cm− 1 . Since the frequency is so low,
the higher harmonics can readily be observed and appear in the
spectrum as a series of nearly equidistant peaks (see Figure 4.7a)
marked by the change in quantum number Δn = 1, 2, 3,….
The spectrum shown in Figure 4.7 further demonstrates a number of
important features of vibrational spectroscopy. The inset (b) shows
an expanded region of the fundamental transition that is split into
several peaks. These are, in part, due to the fact that Br2 is a mixture
of isotopic species: 79 Br2, 79 Br‐, 81 Br, and 81 Br2 with an abundance
ratio of 1:2:1. Thus, the fundamental transitions exhibit these peaks
at 322, 320, and 318 cm− 1 (features f, e, and d in the insert of
Figure 4.8b). The effects of isotopic species will be discussed further
in Example 4.3. Furthermore, the spectrum contains the “hot bands”
of the three isotopic species, since the population of excited states is
quite high for molecular species with low‐lying energy states. These
hot bands (features a, b, and c) are superimposed for different
isotopic species and for different levels of “n” from which they
originate. The energy difference between the fundamental and the
hot bands is only a few wavenumbers, since anharmonicity effects
are small for strongly bonded covalent molecules. This will be
demonstrated in Example 4.3.

Figure 4.7 (a) Raman spectrum of Br2. (b) Expanded region of the
fundamental with overtones and hot bands (adapted from [2]). See
text for details.

Example 4.3 Heteronuclear diatomic molecules such as HF, HCl,


ClF, ClBr, CO, NO, CN−, etc. do exhibit infrared absorption spectra.
Many of these are gaseous at standard conditions and are
complicated by the interaction between vibrational and rotational
transitions, which, in turn, complicates the observed spectra. This
aspect will be discussed in Section 6.5. Thus, we shall discuss the
infrared spectrum of I‐Br, which is a solid with a melting point of 27
°C, has low‐lying vibrational energy levels, and has a similarly strong
bond as Br2 discussed before.

The vibrational constants for I‐Br are given in literature [3] in


wavenumber [cm−1] units:
. Here, is the
harmonic transition energy in wavenumber units for the 127I‐79Br
isotopic species. Since the mass of iodine is even larger than that of
Br, the vibrational frequency is significantly lower than for Br2.
First, we shall compute the frequency of the observed fundamental.
For this, we reformat Eq. (4.47) by dividing by the velocity of light to
get the transition energy in wavenumber units:

(4.47)

(E4.3.1)

(E4.3.2)

This result shows that the observed transition energy is close to the
harmonic value, indicating that the molecule vibrates in a steep and
deep energy well for which the anharmonic perturbation is small.
This is in agreement with the reported bond dissociation energy
value of 177.8 [kJ/mol] for the homolytic bond breakage of I‐Br. This
energy, when expressed in wavenumber per molecule, comes out to
be 14 656 cm−1 (see Appendix 1 for energy conversion factors). Thus,
the energy well is many vibrational quanta deep, and the
fundamental transition occurs far away from the anharmonic
perturbation.
The partially allowed 2←0 overtone will be observed at

(E4.3.3)

and the 3←0 overtone at

(E4.3.4)

These frequencies correspond to the overtone spectrum shown in


Figure 4.8a. The 2←1 hot band frequency is obtained by a similar
approach:

(E4.3.5)

In Eq. (E4.3.2), the observed fundamental frequency was calculated


to be 267 cm−1; therefore, the hot band will appear as a low
wavenumber shoulder at 265.3 cm−1.
Finally, we shall investigate the isotopic splitting for this molecule.
To this end, we assume that the force constant is the same for all
isotopic species. This is a fair assumption, since the isotopes of atoms
differ only by additional neutrons in the nucleus. There are two
dominant isotopic species, 127I‐79Br and 127I‐81Br with molar masses
of 206 and 208, respectively.
Using

(4.41)

we form the ratio of the vibrational harmonic frequencies of the two


isotopic species:

(E4.3.6)

where mR 206 and mR 208 are the reduced masses of the two species.
These are given by

Thus, the isotopic shift observed for the two I‐Br species, 2.3 [cm−1],
is of the same magnitude as the anharmonicity correction. This fact
was seen before in the example of Br2, where the hot bands and the
isotopic species appeared under one overlapping band envelope.

4.6 Summary
The discussion in Chapter 4 bridged from a very simplistic example,
the particle in a box, to a model where the potential function closely
resembles a real potential function, namely, the anharmonic
diatomic oscillator. The mathematics becomes rather involved when
using realistic potential function, but the results are in excellent
agreement with experimental data. Thus, one can say that the
approach taken and alluded to in the introduction (“It doesn't matter
how beautiful a theory is…. If it doesn't agree with experiment, it's
wrong”) about the interplay between theory and experiment is in line
with the philosophy of science, in general: one starts with a theory as
simplistic as possible and continues a refinement process of the
theory until agreement between experiment and theory is achieved.
In the case of the harmonic oscillator, many aspects can be derived
even for this approximate model, for example, the odd/even parity of
wavefunctions and its effect on allowed transitions and the transition
frequency that approximately equals that of the classical model
(Eqs. [4.6, 4.7]). The concept of bond dissociation, as well as the
observation of overtones and hot bands, requires refinement of the
model, and the introduction of the anharmonicity provides this
refinement.
The visualization of the meaning of the wavefunctions plotted in
Figure 4.3 requires a little more thought. In the case of the particle in
a box, the squared amplitude of the wavefunction simply implied the
probability of finding the electron at a given value of x inside the box.
In the case of the harmonic oscillator, we are dealing with two atoms
vibrating about a fixed center of mass. Thus, the square of the
amplitude of the wavefunction shown in Figure 4.2b indicates that in
the ground state, the harmonic oscillator vibrates about x0 or, in
other words, the most likely distance between the two atoms is x0.
This is harder to visualize for the first excited state, where there are
two most likely distances between the atoms and there is a node at
the distance x0. In the case of the particle in a box, the number of
nodal points was given by n−1, whereas it equals to n in the case of
the harmonic oscillator.
Further refinements, to be introduced in later chapters, will take into
account that during rotational motion of a diatomic molecule, the
bond between the atoms will stretch due to centrifugal forces. This
effect is observed in the pure rotational and rot‐vibrational spectra
and demonstrates that further refinements of any model may be
necessary the more sophisticated the experimental methods become.
This does not invalidate prior models, but rather enhances them.
Unfortunately, anti‐scientific opinions often do not understand this
interplay between more detailed experimental methods and
enhanced scientific descriptions.

References
1 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.
2 Baierl, P. and Kiefer, W. (1975). Hot band and isotopic structure
in the resonance Raman spectra of bromine vapor. The Journal of
Chemical Physics 62: 306–308.
3 NIST. Chemistry WebBook, SRD 69. National Institute of
Standards and Technology
https://webbook.nist.gov/cgi/cbook.cgi?
ID=C7789335&Mask=1000.

Problems
1. From the recursion formula for the Hermite polynomials,
determine H5(z) and H6(z).
2. Plot the functions H5(z) and H6(z) between z = −3 and z = +3
(Excel will work fine).
3. Plot unnormalized harmonic oscillator wavefunctions ψ5(z) and
ψ6(z) between z = −5 and z = +5.
4. Plot unnormalized harmonic oscillator wavefunctions ψ52(z) and
ψ62(z) between z = −5 and z = +5.
5. What features (parity, nodal points, intensity distributions, etc.)
can you observe?
6. Show graphically (analog to Figure 4.4) that ψ2(z) and ψ3(z) are
orthogonal and that the transition from n = 2 to n = 3 is electric
dipole allowed in the harmonic oscillator approximation. You
can use unnormalized wavefunctions.
7. In analogy to Problem (6) in Chapter 2, solve for the expectation
value of the total energy operator for the harmonic oscillator for
n = 0.
The following problems deal with the vibration of the carbon
monoxide molecule. Assume that the vibration of the carbon–
oxygen triple bond in CO follows the anharmonic oscillator
formalism, unless stated otherwise.
8. Calculate the force constant k [in N/m] for CO from the
observed stretching frequency, which is 2169 cm−1.
9. What is the stretching frequency [in cm−1] for the isotopic
species 13C18O, assuming that the potential energy does not
depend on the mass.
10. What is the energy [in cm−1] of the n = 0 vibrational state of CO
(i.e. what is the vibrational zero‐point energy)?
11. Calculate the anharmonicity constant χ from the bond
dissociation energy (1077 kJ/mol).
12. Given that the observed frequency (2169 cm−1) is the
anharmonic transition and using the anharmonicity constant
from Problem (11), calculate the transition frequencies (in cm−1)
for
a. n = 1 ← n = 0 harmonic transition
b. n = 2 ← n = 1 hot band
c. weakly allowed n = 2 ← n = 0 overtone
5
Vibrational Infrared and Raman
Spectroscopy of Polyatomic Molecules
Vibrational spectroscopy as an analytical method was established in
the late 1940s mostly by researchers in polymer chemistry when it
was discovered that vibrational (then mostly infrared [IR])
spectroscopy provides important information on polymer chain
crosslinking. Later, the oil industry further developed the
methodology for the analysis of oil products, in particnular the
dependence of the spectra of hydrocarbons on chain length and
saturation. These efforts produced the first software, then confined
to mainframe computers, to perform what was referred to “normal
coordinate analysis” in which the normal modes of vibration were
calculated. This approach uses a strictly classical method in which
the force field, that is, a matrix of all the forces acting between the
atoms in a molecule, was calculated by empirical fitting. In addition,
these calculations revealed the atomic motions during the normal
modes of vibration.
These empirical calculations have now been superseded by quantum
mechanical computations in which the force field is calculated as the
partial derivatives of the total energy E of a molecule with respect to
the Cartesian displacement coordinates (the Hessian matrix [1]). The
total energy is obtained by detailed molecular orbital calculations
using extended basis sets or density functional theory (DFT). This
example again shows the interplay between spectroscopy and
quantum mechanics: the observed vibrational frequencies can be
used for the refinement of the computational methods of molecular
energy. Although these calculations are still computationally
involved, they have produced excellent agreement between observed
and computed vibrational frequencies and intensities.
5.1 Vibrational Energy of Polyatomic
Molecules: Normal Coordinates and Normal
Modes of Vibration
As in the case of the harmonic oscillator, the treatment of vibrational
spectroscopy of polyatomic molecules starts with a classical
description of the vibrational energy. However, this description
requires the derivation of the “normal modes of vibration” of a
molecule. This derivation was well established in classical mechanics
for masses connected by springs with known force constants but is
quite involved for a set of atomic masses where the forces acting
between atoms are unknown. Thus, the derivation of the classical
normal modes of vibration of a molecule requires many steps and
will be presented here in a highly abbreviated form. For step‐by‐step
derivations, the reader is referred to more specialized resources on
vibrational spectroscopy [2].
The classical description is based on the assumption that a molecule
consisting of N atoms can be described by N point masses connected
by springs. In the original efforts in vibrational analysis, the force
constants of these springs were transferred between different
molecules, or fitted to experimental data, first by manual
computation, later by a set of early programs developed for normal
coordinate analysis.
As a starting point for the computation of the vibrational energies of
a polyatomic molecule, Cartesian displacement coordinates xi are
attached to all atoms, since the motions of atoms during a normal
mode can be decomposed into these Cartesian displacement
coordinates. Furthermore, these coordinates are “mass‐weighted”
according to

(5.1)

In mass‐weighted Cartesian displacement coordinates, the


amplitudes of vibration of each atom are properly accounted for by
weighting them by the atomic masses. In terms of these mass‐
weighted coordinates, the kinetic energy is written as

(5.2)

where denotes the time derivative of the coordinate. Notice that


Eq. (5.2) is just the classical definition of kinetic energy (Eq. [2.3]),
summed over all particles.
The potential energy for motions along each Cartesian displacement
coordinate is given by

(5.3)

where the terms fij are the Cartesian components of the force
constants. Again, this expression corresponds to the definition of the
harmonic potential energy introduced by Eq. (4.2) but decomposed
along the 3N Cartesian displacement directions.
The sums in Eqs. (5.2) and (5.3) are overall 3N Cartesian
displacement components. Note that Eqs. (5.2) and (5.3) correspond
exactly to Eq. (4.4) for a one‐dimensional case: in diatomic
molecules, there is just one degree of vibrational freedom – the
stretching of the bond connecting the two atoms. In a polyatomic
molecule with N atoms, there will be 3N degrees of freedom. This
number of degrees of vibrational freedom is based on the concept
that each of the N atoms can move in three independent directions –
the x, y, and z directions. Furthermore, one assumes that every bond
stretching or angle deformation in a polyatomic molecule obeys a
harmonic oscillator formalism or, in other words, all atoms may be
assumed to be connected by springs which obey Hook's law.
When the expressions for the kinetic and potential energy given by
Eqs. (5.2) and (5.3) are substituted into Lagrange's equation of
motion
(5.4)

one obtains a set of 3N differential equations

(5.5)

Here, denotes the second derivative of q with respect to time.


Equation (5.5) is a short form for a set of 3N simultaneous
differential equations, with the index i is running from 1 to 3N.
Following steps described in the literature [2], one can show that
there are 3N solutions to these simultaneous, linear differential
equations which are given by

(5.6)

where the Ai are amplitude factors, and λi are quantities related to


the vibrational frequency and determined by the force constants. The
amplitude factors Ai give relative magnitude of the displacement
vectors that provide a view of the relative amplitudes of all atomic
motions. Notice that Eq. (5.6) is the equivalent of Eq. (4.5) in the
case of a diatomic molecule. Furthermore, six of these equations will
have zero eigenvalues, as will be discussed next.
Molecular vibrations depend on a restoring force to bring the atoms
of a molecule back to their equilibrium position, as discussed before,
for diatomic molecules. Thus, if all atoms in a molecule move
simultaneously in the x‐direction, by the same amount, no bonds are
being compressed or elongated. Thus, this motion is not that of an
internal vibrational coordinate, but that of a translation. Therefore,
the total number of degrees of freedom is reduced by three degrees.
Similarly, a rotation of a molecule without any changes in bond
length or bond angles would not change the potential energy. Thus,
one needs to eliminate three more degrees of freedom, and one
arrives at 3N–6 vibrational degrees of freedom for a nonlinear
polyatomic molecule.
Next, the solutions described in Eq. (5.6) may be recast in a new
system of 3N–6 coordinates, the so‐called normal coordinates Q.
Normal coordinates are linear combinations of the mass‐weighted
Cartesian displacement coordinates, as shown in Figure 5.1, given in
matrix notation as

(5.7)

Figure 5.1 Depiction of the atomic displacement vectors qi for the


three normal modes Q1 (antisymmetric stretching mode), Q2
(symmetric stretching) and Q3 (symmetric deformation mode) of the
water molecule. The magnitude of the displacement vectors is not
known, but the relative displacements are drawn approximately to
scale.
There are 3N–6 normal coordinates that are corresponding to the
3N–6 degrees of vibrational freedom and to the 3N–6 normal modes
of vibration that are associated with the quantum mechanical
observables for the system. In a normal mode of vibration, all atoms
oscillate with the same frequency and in‐phase but with different
amplitudes. This definition implies that all atoms are in motion
during a normal mode of vibration, which is required to maintain the
center of mass of the molecule. The normal modes for a simple
molecule, such as water, are shown in Figure 5.1.
The vibrational frequencies and depictions of the normal modes are
obtained in the classical calculations by transforming the vibrational
problem into normal coordinate space. For that, one recasts Eqs.
(5.2) and (5.3) into matrix notation:
(5.8)

and

(5.9)

Here, the superscript T denotes the transpose of a matrix; thus, the


column vector becomes a row vector upon transposition. The dot
implies, as before, the time derivative of the coordinates. F denotes
the matrix of mass‐weighted Cartesian force constants.

Subsequently, one applies the transformation matrix between mass‐


weighted Cartesian displacement coordinates and normal
coordinates (Eq. [5.7]) and obtains

and

(5.10)

Here, the diagonal matrix Λ = LT F L is obtained by numerically


diagonalizing the force constant matrix F (in mass‐weighted
Cartesian displacement coordinates). The (eigenvector) matrix L that
diagonalizes the potential energy matrix, F, is the matrix that
transforms from the mass‐weighted Cartesian displacement space
into normal coordinate space according to

(5.7)
Λ is the diagonal matrix of the values λk that are the vibrational
frequencies for each normal mode:
(5.11)

A more detailed description of the process of classical normal


coordinate analysis can found in the literature [3]. These
computations are based strictly on a classical model but use the
concept discussed in Chapter 4 that the natural frequencies of
vibration are the same as the quantum mechanically allowed
transition frequencies (see discussion with Eq. [4.39]).

5.2 Quantum Mechanical Description of


Molecular Vibrations in Polyatomic
Molecules
The vibrational Schrödinger equation for a polyatomic molecule is
written in terms of the normal coordinates, since in this coordinate
system, the vibrational frequencies are the quantum mechanical
observables. As in Chapter 4, where the vibrational Schrödinger
equation in the harmonic oscillator approximation was written as

(4.9)

one writes the vibrational Schrödinger equation for a polyatomic


molecule (in the harmonic approximation) in terms of the normal
coordinates as

(5.12)

where ψvib is the total vibrational wavefunction of the molecule


which is a product of the wavefunctions along each of the 3N–6
normal coordinates:

(5.13)
This definition of the total vibrational wavefunction as products of
wavefunctions associated with one and only one normal coordinate
succeeds since the expressions for kinetic and potential energy are
both diagonal in normal coordinate space (cf. Eq. [5.10]).
Substitution of Eq. (5.13) into Eq. (5.12) yields the Schrödinger
equation in terms of the 3N–6 normal coordinates:

(5.14)

Since the normal coordinates Qk are orthogonal functions, Eq. (5.14)


can be separated into 3N–6 individual differential equations of the
form

(5.15)

Equation (5.15) is the 3N–6‐ dimensional equivalent of Eq. (4.8) of


the harmonic oscillator discussion, and assumes a harmonic
potential for each of the 3N–6 normal coordinates. The solutions of
Eq. (5.15) for each normal coordinate are the eigenfunctions

(15.17)

in complete analogy with the harmonic oscillator situation. The total


vibrational wavefunction of an N‐atomic molecule is then given by
Eq. (5.13). A similar situation was encountered before in the two‐
dimensional particle in a box, where the wavefunctions were the
product of the two one‐dimensional wavefunctions, and the total
energy simply the sum of the energies given by the respective one‐
dimensional energy expressions.
The energy eigenvalues (in wavenumber units) for each normal
mode (within the harmonic approximation) is given by
(5.18)

and the total vibrational energy is

(5.19)

The results of this quantum mechanical description are graphically


shown in Figure 5.2 for the water molecule. Each of the three normal
modes of vibration has its own energy ladder, and the vibrational
transitions occur (within the harmonic approximation) within each
ladder, that is, the selection rules Δnk = ±1 apply within each energy
ladder. The vibrational energy spacing for the water molecule
(expressed in wavenumber units; see Eq. [1.11]) associated with each
normal coordinate are approximately

Note that each ladder starts at the zero‐point energy

(5.20)

that is, at approximately 1875, 1825, and 810 cm−1 for Q1, Q2, and Q3,
respectively. Thus, the total zero‐point vibrational energy of water is
given by

(5.21)

or approximately 4510 cm−1. In any of the energy ladders, a


transition to the more highly excited vibrational state may occur if a
photon of proper wavenumber (3750, 1650, or 1620 cm−1) interacts
with the sample, and certain requirements are fulfilled (see also
Figure 5.3). Whether or not a transition occurs depends not only on
the presence of photons with the correct energy but also on some
symmetry considerations discussed in Chapter 11.

Figure 5.2 Energy ladder diagram for the water molecule within
the harmonic oscillator approximation.
Figure 5.3 (a) Observed infrared absorption spectrum of water.
(b) Schematic of the three fundamentals observed in panel (a). (c)
Energy level diagram of the observed transitions.
Linear molecules exhibit 3N–5, rather than 3N–6, degrees of
vibrational modes. The additional degree of vibrational freedom
results from the fact that linear molecules are assumed to have one
less rotational degree of freedom, since rotation along the direction
connecting all atoms has a zero moment of inertia (assuming the
atoms can be described as point masses). Thus, only three
translational and two rotational degrees of freedom need to be
subtracted to arrive at the number of vibrational degrees of freedom.
At this point, a comment about the actual atomic motions in a
molecule is appropriate. These motions are random and increase
with temperature. The familiar “thermal ellipsoids” observed in X‐
ray crystallography of molecular crystals are manifestations of this
random motion. Keeping the temperature low during X‐ray
diffraction data acquisition reduces the size of the thermal ellipsoids.
The random atomic motion, however, can be decomposed into
contributions from the normal modes of vibration. When a
vibrational transition into one of these normal modes occurs, the
random motion along this coordinate increases in amplitude.
The rate of chemical reactions that depend on an initial breakage of a
bond can be enhanced by increasing the vibrational amplitude of the
corresponding bond, either by a non‐selective method – increasing
the reaction temperature – or by illuminating the molecule with light
that is absorbed and increases the amplitude of a highly anharmonic
excited state. This method worked well for the photo‐decomposition
of SF6, where illumination with an IR laser into one of the
antisymmetric S–F stretching vibration actually accelerated the
decomposition [4]. The goal of this work was the development of an
optical method for the separation of uranium isotopes by
illuminating a mixture of UF6 isotopic species with narrow‐band
laser radiation such that only one of the species' decomposition was
accelerated by the IR light. While this method worked for the
separation of 34SF6 from 32SF6, it did not work for uranium
hexafluoride, presumably due to the rapid deactivation of the excited
vibrational states, or the overall density of available states in a
complicated reaction mixture which makes selective excitation of one
highly excited state very difficult.

5.3 Infrared Absorption Spectroscopy


In IR spectroscopy, absorptions of IR photons by the sample are
observed, which cause excitation along one normal coordinate of
vibration into a higher vibrational energy level. Such a process can be
described by Eq. (5.22) which states that the quantum mechanical
transition moment determines the molar extinction coefficient
as follows

(5.22)

which, in turn, determines the likelihood that a photon of


wavenumber is absorbed by the sample. A plot of the molar
extinction coefficient against the wavenumber of radiation gives the
desired IR spectrum.
5.3.1 Symmetry Considerations for Dipole‐Allowed
Transitions
Similar to the case of the diatomic harmonic oscillator, the transition
moment given in Eq. (5.22) must be determined for each of the 3N–6
normal coordinates for a polyatomic molecule. In the case of the
harmonic diatomic molecule, the necessary condition for a
vibrational transition to occur was that the molecule must possess a
permanent dipole moment. In a polyatomic molecule, this rule is
modified to read that the dipole moment along a normal coordinate
must change for a vibration to be allowed to occur in IR absorption
spectroscopy:

(5.23)

This will be discussed in more detail in Chapter 11 on molecular


symmetry but can be viewed as a special condition arising from
parity considerations. At this point, a simple qualitative discussion of
the symmetry requirements will be presented. Transitions along the
three vibrational coordinates shown for water in Figure 5.1 all are
allowed since the water molecule possesses a dipole moment, and
this dipole moment changes for each of the three vibrational
coordinates. Therefore, the water molecule exhibits an IR absorption
spectrum shown schematically in Figure 5.3. The intensity of each
transition shown reflects the magnitude of the dipole change induced
by the corresponding normal mode.
A symmetric linear molecule such as liquid carbon disulfide presents
quite different a situation. Its 3N–5, or four normal modes, are
shown in Figure 5.4. Since the molecule is linear, the two bond
dipole moments of each carbon–sulfur double bond cancel, and the
molecular is non‐polar. The vibrational mode designated as Q2, the
symmetric stretching mode, does not change the dipole moment of
the molecule, and therefore is forbidden in IR absorption. The mode
designated as Q1, the antisymmetric stretching mode, on the other
hand, creates a dipole moment since the two C=S band lengths and
bond dipole moments are different in the vibrationally excited state.
Thus, this transition is allowed. Similarly, the two degenerate
bending modes create a structure that has a dipole moment in the
excited state.

Figure 5.4 Depiction of the atomic displacement vectors qi for the


four normal modes Q1 (a, antisymmetric stretching mode), Q2 (b,
symmetric stretching) and Q3 and Q4 (c, degenerate deformation
modes) of the CS2 molecule. The magnitude of the displacement
vectors is unknown, but the relative displacements are drawn
approximately to scale.
As in the case of diatomic molecules, a further refinement of the
model presented so far would be the inclusion of anharmonicity into
the vibrational force field. Anharmonicity relaxes the selection rules,
as seen before, and weakly allows overtones and combinations and
difference bands. Combination bands are those in which the
quantum number in two different normal modes change
simultaneously; that is, the normal modes of vibration of these
modes interact. There are certain symmetry requirements for this to
happen: the two vibrational transitions must belong to the same
symmetry species (see Chapter 11). Thus, in the water molecule, a
combination band Q2 + Q3 would be allowed, and is observed at
about 5200 cm−1. Incidentally, the slight blue color of water,
observed at long pathlength of visible light through water, is due to a
high overtone 3Q1 + Q2, which occurs at about 14 300 cm−1 or about
700 nm, in the red. Thus, blue light is absorbed less, and water
appears blue at long pathlength.
More examples of vibrational IR spectra will be presented after the
discussion of Raman spectra of polyatomic molecules.

5.3.2 Line Shapes for Absorption and Anomalous


Dispersion
5.3.2.1 Line Shapes and Lifetimes
The width and shapes of the observed IR bands will be discussed
next. A spectral “band” is a plot of the variation of molar extinction
coefficient, or the absorbance of the sample, in the neighborhood of
the band center, as shown in Figure 5.3a. However, the observed
water vibrational spectrum shown in Figure 5.3 has exceptionally
broad IR bands, measured as the full width at half maximum
(FWHM). For the deformation mode Q3 the FWHM is nearly 200
cm−1. This halfwidth is unusually large for a small molecule because
water exists as relatively poorly defined hydrogen‐bonded clusters.
Monomeric water, for example, as a diluted solution in non‐polar
organic solvents, exhibits much narrower peaks, with an FWHM of
about 20 cm−1, and characteristic band shapes shown schematically
in Figure 5.5. The width and shape of observed bands in
spectroscopy is quite intriguing and will be discussed next.

Figure 5.5 Gaussian (a) and Lorentzian (b) line profiles. Notice
that the areas under the line profiles are unequal; however, both
bands have a full width “a” at half maximum (FWHM) of 5 and an
intensity I = 1.0.
As we have seen in the discussion of the perturbation treatment of
stationary states by electromagnetic radiation (ending in Eq. [3.14]),
an absorption transition occurs if the radiation incident on the
sample has a frequency of
(3.16)
However, if this condition is fulfilled exactly, the denominator in the
expression

(5.24)

(Eq. [3.14]) would become zero, and the expression would be


undefined. In reality, the denominator does not become exactly zero
since it contains a damping term that depends on the lifetime of the
excited state. When the damping term is included, the denominator
of the expression in Eq. (5.24) is usually written as

(5.25)

This expression can be derived from classical oscillator theory [5, 6],
with γ the damping term, related to the lifetime of the excited state.
When a photon is absorbed to create an excited vibrational state, the
molecules remain in this excited state for a certain time, referred to
the lifetime of this state. In IR vibrational spectroscopy, this lifetime
is about 10 ps = 10−11 [s]. Given that a vibration with a wavenumber
of 1000 cm−1 has a frequency (cf. Eq. [1.11]) of ν ≈ 3 × 1013 [Hz], or
a period of about 3 × 10−14 [s], the molecule remains in the excited
state for hundreds of complete vibrational cycles, before it
deactivates, either by spontaneous emission, by collisional
deactivation, or by internal energy conversion into other vibrational
energy modes.
The finite lifetime of the excited state determines the width of an
observed absorption peak as follows. The uncertainty principle
introduced in Chapter 2 (Eq. [2.1]) can also be written in terms of
energy and time as follows:

(5.26)

Thus, if the uncertainty in the lifetime by spontaneous emission is


about 10 ps, the corresponding uncertainty in the energy of the state
would be about 5 × 10−23 [J], or about 2.5 cm−1. This uncertainty in
the energy of the state leads to an inherent minimal linewidth of a
few wavenumbers, also referred to as “natural broadening” of a
transition, and the deactivation process is also referred to as the
longitudinal relaxation. Collisional deactivation and internal energy
conversion have similar inherent lifetimes and follow damping
equations of similar form to Eq. (5.25).
The shape of an absorption band with the denominator given by Eq.
(5.25) is referred to as a Lorentzian band that can be described by the
general equation

(5.27)

In Eq. (5.27), x0 is the peak position, “I” the intensity at the band
maximum, and “a” is the FWHM. A Lorentzian band shape is shown
in Figure 5.5b. Thus, the inherent band shape of a transition will be a
Lorentzian band with an FWHM of a few wavenumbers.
Thermal motion of the molecules, particularly in the gas phase,
contributes to another mechanism of band broadening via the
Doppler effect. This effect, which influences the transition frequency
of a molecule, depends on whether it moves toward or away from the
detector of the radiation. This will cause a band broadening that has
a Gaussian profile since the thermal motion itself follows a Gaussian
distribution. The broadening of the band shape by a Gaussian
mechanism produces a line shape given by

(5.28)

shown in Figure 5.5a. The symbols used in Eq. (5.28) are the same as
in Eq. (5.27).
Thus, there are several distinct mechanisms that give raise to the
broadening of spectral bands. Depending on which mechanism
dominates, Gaussian, Lorentzian or mixed bands shapes will be
observed in the IR spectra. These mixed band shape functions often
are described by another function that is a convolution of Gaussian
and Lorentzian band shapes in Fourier space, known as the Voigt
function. Due to the complexity of computing this band shape, it is
often approximated by a “pseudo‐Voight” function that is just a
mixture of Gaussian and Lorentzian band shapes.
In addition, to these “physical” causes of line broadening, there are
several more “chemical” mechanisms, such as the aforementioned
hydrogen bonding, solvation in general, the presence of molecular
interactions, and so forth. Asymmetries in the observed band shapes
at the low wavenumber side can be due to the presence of low‐
intensity hot bands (see Section 4.4 4.4). Many other causes of line
broadening and distortion can be found in the literature on
vibrational spectroscopy [7].

5.3.2.2 Anomalous Dispersion


In the last section, the band shapes created during an absorption
process – that is, a change in the extinction coefficient over a
wavenumber interval – was discussed. The extinction coefficient is
only one part of two quantities that change within an absorption
band. The other quantity that changes simultaneously is the
refractive index. Both of them are combined in the complex index of
refraction η, defined as

(3.22)
in which n is the real refractive index familiar from classical optics,
and κ is known as the absorption index which is related to the
extinction coefficient ε by

(3.23)

Equation (3.22) demonstrates one of the fundamental aspects of


spectroscopy: that the refraction and absorption of light are coupled
processes. That absorption of light always is accompanied by changes
in the refractive index of the medium. Interestingly, the connection
expressed in Eq. (3.22) is often ignored in basic courses in geometric
optics since optical materials from which lenses and prisms are
produced generally do not have any absorptions in the spectral
region for which they are manufactured. Therefore, the refractive
index often is treated as independent of wavelength, or at best,
slowly varying with wavelength. However, at the wavelength at which
any absorption process occurs, the refractive index undergoes what
is known as “anomalous dispersion”, indicated by a deflection point
in the otherwise smooth n(λ) curve. This is shown schematically in
Figure 5.6. At the center of the absorption curve (or at the deflection
point of the dispersive curve), the phase between the electromagnetic
radiation and the response of the medium changes as well. These
effects are well understood and discussed in books on classical
optics, such as the monograph by Born and Wolf [5]. Notice that the
refractive index is always a positive number, even within the regions
of anomalous dispersion.
The coupling of absorption and dispersion is expressed
mathematically by the Kramers–Kronig transform, written here in
terms of the angular frequency ω of the light, rather than its
wavelength λ:

(5.29)

(5.30)
Figure 5.6 Dispersion of the refractive index (top) within an
absorption peak (bottom).
In these equations, ω0 describes the frequency of light at which a
transition occurs. Eqs. (5.29) and (5.30) indicate that knowing one of
the quantities, for example n(ω), uniquely defines the other, ε(ω),
and vice versa. The integrations in Eqs. (5.28) and (5.29) are over a
singularity at ω = ω0, which requires that the principal value of this
Cauchy integral is evaluated. The relationship between the refractive
index and the extinction coefficient is also well‐known in other areas
of spectroscopy: optical rotatory dispersion, which is based on the
differential refractive index of a sample toward left and right
circularly polarized light is related by the Kramers–Kronig transform
to circular dichroism, which is the differential absorption of a sample
toward left and right circularly polarized light (see Section 10.7.2).

5.4 Raman Spectroscopy


5.4.1 General Aspects of Raman Spectroscopy
As indicated in Sections 4.5 and 5.3, vibrational spectral information
can be collected by direct absorption of IR photons or by light
scattering mechanism known as the Raman effect. Light scattering at
a microscopic (molecular) level can easily be visualized: A strong,
focused visible light laser beam travelling through a clean,
transparent liquid sample can be seen with the naked eye since the
molecules in the sample scatter photons in the direction of the
observer. This scattering occurs at the same wavelength as the
incident light and involves a change in the direction, but not the
momentum and energy of the scattered photon. Therefore, this
process is often termed elastic scattering. The scattering cross‐
section (efficiency) of molecular scattering depends on the fourth
power of the frequency of the incident light.
Scattering is a relatively weak process, and only about 1 in 106
photons traveling through a medium will undergo elastic scattering
that is also referred to as Rayleigh scattering. In addition to this
molecular form of elastic scattering, inelastic scattering also may
occur, but with even lower efficiency than elastic scattering. In this
inelastic scattering, both the direction and energy of the photon
change. This scattering process is known as Raman scattering,
named after Sir Chandrashekhar Raman, who first observed this
effect experimentally in 1928 after A. Smekal had predicted it in
1923.

5.4.2 Macroscopic Description of Polarizability


Whereas IR spectroscopy is caused by direct absorption of a photon
to promote the molecule into a vibrationally excited state and
requires a change of the dipole moment along a vibrational
coordinate, Raman spectroscopy requires the polarizability of a
molecule to change along a vibrational coordinate. The polarizability
of a molecule is its response to the incident radiation far from an
electronic transition. This can be visualized as follows. In a clear,
colorless liquid, for example, there are no electronic transitions
between ca. 400 and 750 nm, or ca. 25 000 and 13 000 cm−1 (hence
the material is colorless). Visible electromagnetic radiation, however,
still can interact with the molecule by setting in motion the electron
clouds (particularly those in multiple bonds) by inducing a dipole
moment, μind, mediated by the polarizability α:

(5.31)
Here, E denotes the strength E of the electromagnetic field. Since
both the induced dipole moment and the electric field are vectors,
the polarizability is actually a tensor.
Thus, Eq. (5.31) can be written as a vector equation:

(5.32)
where both μ and E are vectors, and is represented by a 3 × 3
matrix:

(5.33)

This polarizability varies as the molecule oscillates along its normal


coordinates, Qk, since the polarizability (that is, the ease with which
electrons can be moved around) depends very much on the nuclear
coordinates, and thereby on the vibrational modes of the molecule.
Thus, one may expand the polarizability in a Taylor series about the
equilibrium position according to

(5.34)

where ωk is the vibrational frequency of normal mode Qk.


The electric field exciting the Raman scattering is usually provided
by laser radiation at ωL, one can represent this radiation by

(5.35)
Any oscillating dipole, whether induced or permanent, emits
radiation of intensity I into all space according

(5.36)

Equation (5.36) shows that the scattered light intensity depends on


the fourth power of the frequency, and on the square of the electric
field strength and the polarizability. The polarizability itself is
modulated by the molecular vibrational motion according to Eq.
(5.33); thus, the induced dipole moment contains two frequency
components, ωk and ωL. The radiation emitted by the induced dipole
contains frequency components at the frequency of the exciting light
ωL, and at the beat frequencies between the molecular frequency and
the frequency of the incident radiation. The beat frequencies ωL + ωk
and ωL − ωk arise from the product of the two cosine functions in
Eqs. (5.33) and (5.34) and are given by a well‐known trigonometric
identity

(5.37)

The beat frequencies are the anti‐Stokes and Stokes Raman


frequencies, respectively. Thus, a simple, classical description of the
off‐resonance interaction of light with a polarizable molecular system
can explain some aspects of Raman scattering.

5.4.3 Quantum Mechanical Description of Polarizability


A basic quantum mechanical description of the polarizability can be
found in reference [8]. In short, this derivation proceeds as follows.
One writes the induced dipole moment, μind, as the expectation value
of the operator, μ:

(5.38)

where Ψ denotes the time‐dependent wavefunction defined in Eq.


(3.10):

(5.39)

As before, the wavefunctions ψ(x) are the stationary state


wavefunctions, φ(t) the time‐dependent wavefunctions, and ck(t) the
time‐dependent expansion coefficients that describe how the system
transitions from one state to another. These coefficients are given by
Eq. (3.14).
Substituting Eqs. (3.14) and (5.38) into Eq. (5.37), and assuming that
the time dependence of the electric field can be written as

(5.35)
the induced dipole moment is given by

(5.40)

Here, one assumes that the molecule is in the ground state |ψ0〉
before any interaction with the radiation occurs. The first term in Eq.
(5.40) describes the permanent dipole moment, the second term a
component of the induced dipole that oscillates at the same
frequency as the incident light, and the third term an induced dipole
that oscillates at the transition frequency ω0m and is clearly not in
phase with the frequency of the light inducing the dipole moment.
The second term is called the polarizability, and a comparison
between Eqs. (5.35) and (5.40) reveals that

(5.41)

Equation (5.40) implies that any medium exposed to


electromagnetic radiation will undergo some, albeit small, change,
even when the frequency of the radiation is far from any transition
frequency, ω0m. This change can be viewed as a small, induced
oscillatory motion of the electrons in a molecule; tightly bound inner
electrons will respond to a lesser degree than more loosely bound
electrons such as π‐electrons in a double bond. This interaction also
determines the refractive index, n, and the relation between
polarizability and refractive index is given by

(5.42)

Here, N is Avogadro's number. Thus, one finds that off‐resonance


spectroscopic properties, such as the dielectric constant or the
refractive index, are related to the transition moment via the
polarizability. Thus, the proximity of an electronic transition will
increase the refractive index. In the visible region of the spectrum,
one can, therefore, predict that acetone has a higher refractive index
than water, since its closest UV transition occurs around 280 nm,
whereas the closest transition for water lies below 200 nm in the far
UV region.
Raman scattering can then be described as a process in which an
incident photon, typically from a laser and with an angular frequency
ωL interacts momentarily with the sample by creating a virtual state,
indicated by the dashed line in Figure 5.7a. When it is scattered from
this state, a photon of the same frequency is emitted (Rayleigh
scattering), or a photon with angular frequency ωL – ωM is scattered
(the Stokes Raman photon) where ωM is the frequency of a molecular
vibration. The Raman scattering leaves the molecule in the same
excited state that is reached by direct absorption of a photon with
angular frequency ωM. It is also possible that the original laser
photon encounters the molecule in the vibrationally excited state.
The virtual state now is higher in energy, and a photon with an
angular frequency of ωL + ωM is scattered, a so‐called anti‐Stokes
Raman photon. Anti‐Stokes Raman intensities are lower than those
of the Stokes spectrum since the former depends on the population
of the vibrationally excited state from which they originate. This
population depends, via the Boltzmann equation (Eq. [3.31]) on the
absolute temperature and the energy of the vibrational level. Thus,
anti‐Stokes intensities are highest at high temperature, and for those
vibrational energy levels with the lowest vibrational frequencies.

Figure 5.7 (a) Energy level diagram for a Stokes and anti‐Stokes
Raman scattering process involving a virtual state (dashed line). (b)
Energy level diagram to model polarizability in terms of sum of
excited state transitions. See text for details.
The virtual state depends on the energy (frequency) of the incident
photon, and not on a real molecular energy level since this level is
created by the photon itself via the polarizability. According to Eq.
(5.40), the polarizability can be viewed as the sum of all the
electronic transition moments in a molecule, each one weighted by
an energy term in the denominator. This is shown in Figure 5.7b for
a hypothetical molecule with two real electronic excited states
defined by the time‐dependent vibronic wavefunctions Ψe ′ v and Ψe "
v. Here, as well as in the discussion below, we use the subscript “e”
for an electronic and “v” for a vibrational state. A prime or double
prime on either subscript indicates an excited state. As shown in
Figure 5.7b, the transitions into the first real excited state can
happen from the ground vibrational state of the ground electronic
state, Ψev or from a vibrationally excited state from the ground
electronic state . The polarizability matrix elements for both
these transitions are given, divided by the energy difference between
the transition energy ωee' minus the energy of the laser photon.
Equivalent expressions would be written for the matrix elements into
the other electronic state.
This is summarized by the definition of the polarizability tensor
elements as follows:

(5.43)

The subscripts α and β in Eq. (5.43) indicate the Cartesian


coordinates of the polarization tensor. This equation also indicates
that Raman spectroscopy is a two‐photon process since there are two
electronic transition moment expressions contained in the definition
of the polarizability. Thus, the selection rules for the Raman effect,
depending on the binary components x·x, x·y, x·z, etc. of the
Cartesian displacement coordinates, which are listed in the character
tables (see Chapter 11).
Equation (5.43) holds for the so‐called “far from resonance” (FFR)
case in which the photons of the incident light do not have energies
close to those of an electronic transition. Therefore, the contribution
of a given transition moment is small. However, if the frequency of
the exciting light approaches the energy difference between two real
stationary states, that is, if , the corresponding term in the
sum expressed by Eq. (5.42), gets very large. In this case, one
observes an enormous increase in Raman scattered intensities. This
phenomenon is referred to as “resonance Raman” scattering that will
be discussed next.
Resonance Raman spectroscopy was experimentally verified in 1972
[9] and has had a profound impact on the study of the structure and
dynamics of biophysical systems. In addition to the enhancement of
the scattering cross‐section by several orders of magnitude,
Resonance Raman spectroscopy offers the enormous advantage that
only groups on which the electronic transition is located will
experience the resonance enhancement.
The basic theory of resonance Raman intensities usually is discussed
in terms of the scattering tensor, rather than the polarizability, which
has the form

(5.44)

As in Eq. (5.43), the subscripts α and β of the scattering tensor α


refer to all permutations of the Cartesian coordinates x, y, and z. In
Eq. (5.44), r is the intermediate (real or virtual) state, and each
element of the scattering tensor is defined as the sum over all
vibronic states of the molecule. The subscripts n and m denote the
final and original states of the system.
The scattering tensor equation explicitly contains a damping term iΓ
in the denominator, which was neglected in the previous discussion
of non‐resonant Raman spectroscopy. This damping term physically
is the lifetime of the intermediate state and prevents the
denominator from becoming exactly zero at the resonance condition.
In the “far‐from‐resonance” case, and if the molecule is initially in its
ground state, Eq. (5.44) assumes the form of the polarizability tensor
introduced in Eq. (5.43). Details on the discussion of the scattering
tensor can be found in the literature [10].
Within the Born–Oppenheimer approximation (see Section 10.4.2),
the transition moments in the numerator of Eq. (5.44) can be
separated into pure electronic transition moment between states
r and m, and the Franck–Condon overlap integrals between the
vibrational wavefunctions:

(5.45)
Here, states i and j represent the vibrational states of the ground
electronic state; v is a vibrationally excited state of the resonant
excited state. Equation (5.45) thus represents how much the
resonance excited state is displaced along the vibrational coordinate.
For the discussion of resonance enhancement, all states involved are
written as the products of vibrational and electronic wavefunctions,
and the dipole transition moments are evaluated separately for the
purely electronic and vibrational wavefunctions. This allows the
scattering tensor to be written as the sum of two terms, referred to as
the A and B terms:

(5.46)

(5.47)

The equation for the A term describes the resonance enhancement in


totally symmetric modes, whereas the B term dominates when the
vibrational modes mix the two excited electronic states. The
resonance enhancement due to the A and B terms have different
frequency responses, which determine the onset of resonance
enhancement as the laser wavelength approaches, in energy, an
electronic transition. This aspect is particularly important when
discussing the resonance Raman spectra of inherently symmetric
moieties, such as the iron–porphyrin groups in heme proteins.
Resonance Raman studies also focused on the intermediates and the
conversation dynamics in the light‐harvesting proteins such as
bacteriorhodopsin and similar molecules. Here again, the prosthetic
group and its interaction with the protein bundles spanning the
cellular membrane could be studied.

5.5 Selection Rules for IR and Raman


Spectroscopy of Polyatomic Molecules
In Chapter 4, the selection rules for the harmonic and the
anharmonic oscillator were derived to be

(4.37)
and

(4.49)
respectively. A diatomic harmonic oscillator has but one normal
mode of vibration, Q; and the only condition for a transition to be
allowed in absorption is the requirement of the molecule being polar;
otherwise, the expression

(5.48)

For polyatomic molecules, the general rule holds that for a transition
along a normal coordinate Qk to occur in absorption – both for polar
and non‐polar molecules – the condition

(5.49)

must hold. In Eq. (5.49), the subscript α implies that any x, y, or z


component of the dipole transition moment must be non‐zero. This
was mentioned before in Section 5.3.1 for the non‐polar symmetric
CS2 molecule.
In Chapter 3, the general form of the transition moment in
absorption between two states was given as
(3.20)

For a vibrational absorption transition, we accordingly define the


transition moment for the 1 ← 0 transition along normal coordinate
Qj as

(5.50)

However, for polyatomic molecules, it may not be obvious which


vibrations change the dipole moment, and we need to resort to group
theory and symmetry to determine which transitions are allowed in
absorption. This is presented in more detail in Chapter 11.
The condition corresponding to Eq. (5.49) for a transition to be
active in Raman scattering is

(5.51)

where the subscripts αβ denote permutations of the coordinates x, y,


and z. The expression for the transition moment contains two terms,
since Raman scattering is a two‐photon process, see Figure 5.7b.
Thus, terms such as

(5.52)

are involved where ψint is the intermediate state from which the
scattering takes place. Since there are two transitions involved, the
selection rules will depend not just on the x, y, and z components,
but on binary combinations such as x2, xy, xz, and so forth. This also
will be discussed in Chapter 11.

5.6 Relationship between Infrared and


Raman Spectra: Chloroform
In this section, a comparison between the observed IR absorption
and Raman spectra for the same molecule, chloroform, will be
presented. Raman and IR absorption spectroscopy are two
complementary techniques in the sense that for some molecules
(those with a center of inversion symmetry element; see Chapter 11),
transitions that are allowed in Raman scattering are forbidden in IR
absorption, and vice versa. However, even for many other molecules,
certain vibrational modes – those that strongly change the dipole
moment – are strong in absorption and weak in scattering. This is
shown in Figure 5.8 for chloroform, CHCl3.
This highly symmetric, pentatomic molecules should exhibit 3N–6 =
9 vibrational modes of freedom. However, some of these modes are
degenerate (see Section 2.4 and Chapter 11); thus, only six bands are
observed in both Raman and infrared spectra shown in Figure 5.8
and summarized in Table 5.1. The modes observed include the C–H
stretching motion that is observed at 3034 cm−1, typical for CH
stretching modes (cf. Example 4.1), and the C–H bending motion at
1220 cm−1 in which H–C–Cl bond angle changes. This mode is
doubly degenerate and is quite strong in infrared absorption and
weak in Raman scattering. The symmetric ‐CCl3 stretching mode at
680 cm−1 is the strongest band in the Raman spectrum but quite
weak in absorption, whereas the antisymmetric stretching mode at
774 cm−1 is extremely strong in the absorption spectrum but weak in
Raman scattering. The atomic displacement vectors of these two
modes are shown schematically in Figure 5.9.
The two remaining normal modes are the symmetric and the
antisymmetric ‐CCl3 deformations. In the former, all three Cl–C–Cl
angles decrease by the same amount and in phase. This mode is
nicknamed the “umbrella” mode and occurs at 363 cm−1. The
degenerate antisymmetric deformation is observed at 261 cm−1.
Table 5.1 Vibrational modes and assignments for chloroform,
HCCl3.

Designation Observed transition Mode


wavenumber (cm−1) description
ν6 261 CCl3 antisym
deformation
ν5 363 CCl3 sym
deformation
ν4 680 CCl3 sym
stretching
ν3 774 CCl3 antisym
stretching
ν2 1220 CH bending
ν1 3034 CH stretching

As in the case of diatomic molecules, the potential energy function


that determines the molecular vibration of polyatomic molecules is
not quadratic along the normal coordinates. Thus, anharmonicity
corrections need to be included for the detailed understanding of the
vibrational transitions. The anharmonicity, as before, permits the
occurrence of overtone and hot bands. In addition, transitions
between different energy ladders (see Figure 5.3) are now weakly
allowed, which leads to combination and difference bands, such as
ν2 + ν3, that is, the normal coordinates mix in such transitions.
Moreover, an effect known as Fermi resonance can increase the
intensity of such transitions. Detailed discussions of these effects can
be found in more specialized books on vibrational spectroscopy [1].
Figure 5.8 (a) Raman spectrum of chloroform as a neat liquid. (b)
Expanded view of (a) showing the C–Cl stretching and the C‐H
deformation modes. (c) Infrared absorption spectrum in the same
region as spectrum (b) showing a number of overtone and
combination bands.

5.7 Summary: Molecular Vibrations in


Science and Technology
Vibrational spectroscopy is a widely used analytical technique, and it
also is an active field of research to probe vibronic states and very
fast molecular dynamic processes. Also, the field of non‐linear
spectroscopy – where molecular properties that depend on the
square or cube of the electric field strength (see Appendix A3) are
probed – requires a detailed understanding of the principles of
vibrational spectroscopy discussed in this chapter. This field of
nonlinear spectroscopy is a hot topic in modern optical research, and
results from this research have trickled down into consumer
products, such as the green laser pointer available for under $50.
This device uses a frequency‐doubled Nd:YAG laser at 532 nm, and is
the size of a pen. This field of nonlinear optics and non‐linear
spectroscopic effects will be touched upon in Appendix 3.
Molecular vibrations play a major role in many other branches of
science and technology. At standard thermodynamic conditions (1
atm pressure, 25 °C), the energy levels that are mostly populated are
vibrational energy levels. The thermal energy, RT, at room
temperature, is given approximately by

(5.53)
(For conversion of the energy units, see Appendix 1). Thus, at room
temperature, the vibrational energy levels are most populated,
according to the Boltzmann distribution, whereas the electronic
energy level population is very low (electronic transitions in the
visible and ultraviolet spectral regions have energies upward of ca.
15 000 cm−1), therefore, are barely populated. Rotational states, with
much lower energy spacing, are highly populated at room
temperature only for gaseous molecules since molecular rotation
depends on molecules rotating freely in space (see Chapter 6). Thus,
molecular vibrations are the most populated energy states under
standard conditions. According to statistical thermodynamics,
quantities such as the heat capacity of materials are mostly
determined by the population of the vibrational energy levels via the
expressions of the partition function. The heat capacities of gases, a
favorite subject in engineering courses are also highly dependent on
the rotational and vibrational degrees of freedom, since each degree
contributes the amount of ½ R.

Figure 5.9 Atomic displacement vectors for (a) the symmetric –


CCl3 stretching and the two degenerate antisymmetric –CCl3
stretching modes (b, c).

References
1 Diem, M. (2015). Modern Vibrational Spectroscopy and Micro‐
Spectroscopy: Theory, Instrumentation and Biomedical
Applications. Chichester, UK: Wiley.
2 Wilson, E.B., Decius, J.C., and Cross, P.C. (1955). Molecular
Vibrations: The Theory of Infrard and Raman Vibrational
Spectra. New York: McGraw–Hill Co.
3 Diem, M. (1993). Introduction to Modern Vibrational
Spectroscopy. New York: Wiley‐Interscience.

4 Grant, E.R. et al. (1977). The Extent of Energy Randomization in


the Infrared Multiphoton Dissociation of SF6. LBL Report 6201.
5 Born, M. and Wolf, E. (1970). Principles of Optics. New York:
Pergamon Press.
6 Tatum, J. (2019). Natural Broadeninf (Radiation Damping).
Physics Libretexts. https://phys.libretexts.org/.
7 Bradley, M.S. (2015). Lineshapes in IR and Raman spectroscopy:
a primer. Spectroscopy 30 (11): 42–46.
8 Kauzman, W. (1957). Quantum Chemistry. New York: Academic
Press.
9 Spiro, T.G. and Strekas, T.C. (1972). Resonance Raman spectra of
hemoglobin and cytochrome c: inverse polarization and vibronic
scattering. Proceedings of the National Academy of Sciences of
the United States of America 69 (9): 2622–2626.
10 Koningstein, J.A. (1972). Introduction to the Theory of the
Raman Effect. Dordrecht: D.Reidel Publising Co.

Problems
Advise: The interpretation of vibrational spectroscopy of polyatomic
molecules requires an understanding of the symmetry properties of
the molecules and their infrared‐ and Raman‐active vibrations. Thus,
the problems in this chapter are doable only after studying Chapter
11.

1. Acetylene, H–CΞC–H is a linear, centrosymmetric molecule.


This molecule exhibits a strong Raman band at ca. 1975 cm−1,
assigned to the CΞC stretching vibration. Do you expect this
band to be observed in the infrared spectrum as well? Explain.
2. Acetylene exhibits the antisymmetric C–H stretching mode at
3295 cm−1. Is this mode Raman active? Treating this molecule as
a diatomic species H–X, predict the vibrational frequency of the
carbon–deuterium stretching frequency (in cm−1) for D–CΞC–H
3. Some of the early applications of vibrational spectroscopy were
in the determination of molecular shapes from observed spectra.
Such structural information is nowadays obtained from simple
models like the valence shell electron pair repulsion theory
(VSEPR) that predicts that SF4 has a seesaw structure of C2v
symmetry shown on page 92, rather than a tetrahedral
structure. Predict the spectral differences you would expect
between a Td and C2v structure.

4. For the following molecules, determine the symmetry groups,


the infrared and Raman allowed transitions and their symmetry
species. Use the literature to find as many of the observed
vibrational spectra and assignments.
a. Carbon tetrafluoride, CF4
b. Fluoroform, CHF3
c. Difluoromethane, CH2F2
d. Fluoro–chloro–methane, CH2FCl
5. How does the symmetry of a molecule affect the number of
infrared and Raman transitions observed in a spectrum?
Specifically, methane shows only two absorptions in infrared
spectroscopy, whereas another pentatomic molecule, CH2FCl
shows nine absorption fundamentals.
6
Rotation of Molecules and Rotational
Spectroscopy
Whereas vibrational spectroscopy discussed in the previous chapter
is a widely applied technique for qualitative analysis via group
frequencies and extensive search libraries containing tens of
thousands of spectra, rotational spectroscopy is used less frequently
in analytical science, mostly because of sampling issues – the sample
has to be in the gaseous phase – instrumental complexity, and the
fact that a simple qualitative interpretation as in nuclear magnetic
resonance (NMR) or infrared spectroscopy is not possible for
rotational spectroscopy. This is unfortunate since much of today's
detailed chemical structural information was obtained from
rotational spectroscopic data, since this technique reveals structural
information like no other spectroscopic technique. In introductory
chemistry courses, for example, a predictive model is introduced, the
so‐called valence‐shell electron‐pair repulsion (VSEPR) model that
allows qualitative prediction that the H–N–H angle in ammonia is
smaller than the H–C–H angle in methane, since the free electron
pair in ammonia occupies more space than a bonding pair in
methane. Such information was originally deduced from microwave
data, since it allows the measurement of the moment of inertia
extremely accurately, and therefore, bond distances and bond angles
can be calculated to five or six significant figures.
As in the case of vibrations of di‐ and polyatomic molecules, a
classical model allows the calculation of the moment of inertia of a
molecule from structural data. This aspect will be covered first in
Section 6.1. The quantum mechanical treatment of the rotation of a
rigid molecule (Sections 6.2 and 6.3) introduces the angular
momentum operator that leads to two differential equations that give
us the rotational wavefunctions, eigenvalues, and selection rules.
These differential equations also form the basis of solving the
Schrödinger equation for the hydrogen atom – after all, the hydrogen
atom could be thought of as a “diatomic molecule” consisting of a
proton and an electron rotating in space. The concept of the angular
momentum also is required for the understanding of spin
spectroscopy. Thus, the mathematics presented in Section 6.3 is the
basis of several subjects in this book.

6.1 Classical Rotational Energy of Diatomic


and Polyatomic Molecules
Classically, any rotatory motion can be described by the rotational
kinetic energy

(6.1)

where L is the angular momentum, and I is the moment of inertia.


This definition is in complete analogy to the definition of the (linear)
kinetic energy

(1.17)

in which the kinetic energy, in both cases, is proportional to the


square of the corresponding momentum divided by a quantity
measuring the inertia, or “resistance” to this motion. The angular
momentum in Eq. (6.1) is defined as the following:

(6.2)
that is, the vector product of the linear momentum and the radius of
the circular motion. Thus, L is a vector quantity, indicated in bold
face.
Multiplying out the vector components r = ix + jy + kz and p = ipx +
jpy + kpz in Eq. (6.2) reveals that

(6.3)
with

(6.4)
and similar expressions holding for the other components of L.
The (scalar) moment of inertia, as pointed out above, depends very
much on the shape (symmetry) of the molecule. In the simplest case
of a linear, diatomic molecule, the moment of inertia is given by

(6.5)

where m1 and m2 are the masses of the two atoms and r1 and r2 their
distances from the center of mass (see Figure 4.1). The center of
mass condition is given by m1r1 + m2r2 = 0. For a diatomic molecule,
one moment of inertia is zero (the one along the bond axis) because
it is assumed that the atoms are point masses, and rotation about the
bond connecting the two atoms experiences no inertia. The other two
components of the moment of inertia, Ix and Iy, are equal. The same
argument holds true for linear molecules with more than two atoms.
For a nonlinear molecule in an arbitrary coordinate system, the
moment of inertia is a tensor defined as

(6.6)

In Eq. (6.6), the summation is over all atoms in the molecule, mi is


the mass of the individual atoms, and x, y, and z are their Cartesian
coordinates [1]. If the inertial tensor is written in arbitrary
coordinates x, y, and z, its values will depend on the choice of the
origin of the coordinate system; i.e. different numerical values for the
tensor terms Ixx and others are obtained when the origin is shifted.
However, the rotational motion of a molecule, or for that matter any
freely rotating body, can be described such that the moment of
inertia is independent of the choice of the coordinate origin. This is
the case when the overall rotation and translation of the molecule is
uncoupled and was accomplished for diatomic molecules by
introducing the reduced mass, which is based on a coordinate system
whose origin lies at the center of mass.
If the center of mass of the molecule does not translate, the
molecular motion is pure rotational motion. Under these conditions,
the inertial tensor can be written in the diagonal form:

(6.7)

Here, T is a coordinate transform matrix between an arbitrary


coordinate system and the principal axes of inertia; in this latter
coordinate system, the inertial tensor is diagonal, and the diagonal
terms IA, IB, and IC are known as the principal moments of inertia
and are related to the observables in rotational spectroscopy. The
superscript in TT denotes the transpose of the matrix T.
The square of the total angular momentum, L2, required in Eq. (6.1)
can be written as a sum of the individual principal components
according to

(6.8)

since squaring the vector quantity in Eq. (6.3) eliminates all cross‐
terms. Thus, the rotational kinetic energy is

(6.9)

Since the potential energy of free rotational motion is zero, the total
energy of the system is given by the kinetic energy, i.e. E = T. The
kinetic energy is generally expressed in terms of the rotational
constants A, B, and C, rather than IA, IB, and IC. The rotational
constants are defined as

(6.10)

with the convention that IA ≤ IB ≤ IC. Notice that the rotational


constants are expressed in units of 1/s = Hz. To convert to energy
units, these constants need to be multiplied by Planck's constant. In
terms of the rotational constants, the rotational kinetic energy is
written as

(6.11)

The rotational constants are related to the quantum mechanical


energy eigenvalues observed in rotational spectroscopy. They contain
all the structural parameters of a molecule and allow the
determination of molecular structures within picometer and
millidegree accuracy.
Depending on the shape of molecules, they can possess either one,
two, or three distinct moments of inertia and are referred to as
spherical and linear top rotors, symmetric top rotors, or asymmetric
top rotors. These have different energy eigenvalues as we shall see in
Section 6.2 that depend on the rotational constants as follows:

a. If all rotational constants A, B, and C of a molecule are equal, it


is referred to as “spherical top rotor.” A molecule has to belong
to one of the spherical point groups (cf. Chapter 11) of Td or Oh
symmetry. The total rotational energy can be expressed in terms
of L2 only. By convention, the rotational constant B was selected
to represent the rotational energy (cf. Eq. [6.11]) as

(6.12)

A spherical top molecule cannot have a permanent dipole moment


and therefore will not exhibit a rotational absorption (microwave)
spectrum. Its polarizability also will not change during a rotation,
and spherical top molecules will not exhibit a pure rotational Raman
spectrum either. They will, however, exhibit a rotational–vibrational
absorption spectrum in some bands (see Section 6.6) since some
vibrational transitions distort the molecule such that it shows a
temporary dipole moment.

b. For linear molecules (obviously including diatomic molecules),


one rotational constant is zero (the one along the bond axis),
and the other two are equal. This leads to an energy equation for
linear molecules identical to the one for spherical molecules (Eq.
[6.12]). However, a linear molecule must possess a permanent
dipole moment to exhibit a rotational absorption (microwave)
spectrum. Thus, N2 and Cl2 will not show a rotational spectrum,
but they will exhibit a pure rotational Raman spectrum due to
the different selection rules in Raman spectroscopy.
c. For symmetric tops, two moments of inertia are equal: either

IA = IB < IC (oblate symmetric top rotor) or


IA < IB = IC (prolate symmetric top rotor). For the former case,

(6.13)

since

(6.14)

(6.15)

(6.16)
(6.17)

For the prolate top rotor, this corresponding equation is

(6.18)

In both equations, the two operators, L2 and , commutate and


have eigenvalues in the same vector space. Thus, the problem can be
solved.

d. For the asymmetric top rotor, all three moments of inertia are
unequal, and their operators do not commutate. Therefore, the
problem cannot be solved explicitly.

The solutions of the rotational Schrödinger equation will be


presented in Section 6.3, first for the case of linear or spherical top
molecules for which the rotational energy depends on the
eigenvalues of the L2 operator according to Eq. (6.12).

6.2 Quantum Mechanical Description of the


Angular Momentum Operator
In analogy to the discussion of vibrational spectroscopy, a classical
model for the rotational energy was derived above. The classical
models for both the rotation and the vibration of molecules predict
the molecular energies properly in a steady‐state situation, that is,
without interactions with light. In order to predict these interactions
and the possibility of observing a spectrum due to the molecular
system undergoing transitions between stationary state energy
levels, quantum mechanics needs to be invoked.
Since there is no change in the potential energy of a molecule during
molecular rotations, the equation that governs the total energy
during molecular rotations is just Eq. (6.11). To obtain the rotational
Schrödinger equation, the angular momentum components in Eq.
(6.11) need to be replaced by the corresponding quantum mechanical
operators. This is accomplished by writing the angular momentum
components given by Eq. (6.4)

(6.4)
and subsequently substituting the linear momentum components by

(2.2)

The component of the angular momentum operator given in Eq.


(6.4) then becomes

(6.19)

with equivalent expressions holding for the other two Cartesian


components of the angular momentum. Because of the
differentiation required when operating with the , and
operators on a function f(x, y, z), the commutator of these operators
is defined as

(6.20)
Figure 6.1 Definition of spherical polar coordinates.
is nonzero. As discussed in Section 2.1, the consequences of Eq.
(6.20) are quite far‐reaching: for a molecule with three different
moments of inertia along the three principal axes (a so‐called
asymmetric top rotor), it is impossible to determine the three
principal moments of inertia. However, the total angular moment
operator, L2, does commutate with all the components of the angular
momentum operator and thus, it is possible to determine the total
moment of inertia and one of the three components, normally
assumed to be the Lz operator (although any of the three Cartesian
components could be selected).

(6.21)

The problem of solving the quantum mechanical expressions for the


rotational energy of a molecule rotating about its center of mass is
best tackled in spherical polar coordinates, rather than Cartesian
coordinates. Spherical polar coordinates are defined as shown in
Figure 6.1
In spherical polar coordinates, the and operators take the
form

(6.22)

and

(6.23)

The transformation of the operators from Cartesian to spherical


polar coordinates is by no means trivial and described in detail in
[2].
The functions that simultaneously are eigenfunctions of the two
operators in Eqs. (6.22) and (6.23) are the so‐called spherical
harmonic functions Y(θ, φ):

(6.24)

with eigenvalues c and b.


The angular momentum operator components have been defined
above in a space‐fixed coordinate system as x, y, and z, whereas the
classical moment of inertia has been described in a coordinate
system that rotates with the molecule, the principal axes of inertia.
The relation between these two coordinate systems is given by Eq.
(6.7) and is just a simple coordinate system transformation.

6.3 The Rotational Schrödinger Equation,


Eigenfunctions, and Rotational Energy
Eigenvalues
To obtain the eigenfunctions and eigenvalues for a rotating molecule,
the Schrödinger equation for the total angular momentum
(6.1)

with

(6.25)

and one of its Cartesian components, for example, , with

(6.26)

need to be solved simultaneously. As we shall see below, the


solutions of these two operator/eigenvalue problems are the
spherical harmonic functions Y(θ, φ), which can be separable into
two functions in the variables θ and φ according to

(6.27)
To find the eigenvalues and eigenfunctions of the and
operators, we start with the operator:

(6.28)

Since only contains the variable φ and T(φ) depends on φ only,


Eq. (6.28) simplifies to

(6.29)

Integrating Eq. (6.29) yields

(6.30)
where A is a constant (the amplitude). Since T is a function of φ only,
the boundary conditions for T(φ) are such that the function has to
have the same value for multiples of 2π; otherwise, the function
would self‐destruct by interference, as shown in Figure 6.2. This
boundary condition can be written as

(6.31)
from which follows that

(6.32)
Figure 6.2 Graphical representation for the condition T(φ) = T(φ
+ b 2π), where b is an integer.
Defining

(6.33)

Eq. (6.30) can be rewritten as

(6.34)
with K = 0, ± 1, ± 2, ± 3, …, which implies that the eigenvalues of the
z‐component of the angular momentum are quantized, and K is a
rotational quantum number associated with the operator.
Normalization of Eq. (6.34) by integrating over φ( 0 ≤ φ ≤ 2π) results
in the final form of the eigenfunctions of the operator as

(6.35)

Next, the θ‐dependent part of Eq. (6.27) or (6.28) needs to be solved


to give the eigenfunctions S(θ) of the operator. Since this
operator contains derivatives with respect to both θ and φ, we utilize
the second derivative of Eq. (6.35) in Eq. (6.36):

(6.36)

which accounts for the factor K2 in Eq. (6.37):

(6.37)

Equation (6.37) shows why the solutions of S(θ ) contain the


quantum number K, which accounts for, as we shall see later, the
degeneracy of the eigenvalues of S(θ ).
The differential equation, Eq. (6.37), is solved [3], as in the case of
the harmonic oscillator, by a power series expansion in cos(θ). Again,
this power series has to be broken off at some value of the expansion
coefficient; otherwise, the wavefunctions wouldn't be finite. This
condition, along with the recursion formula for the expansion
coefficients, leads to the energy eigenvalues “c” in Eq. (6.37):
(6.38)
where J can assume values J = 0, 1, 2, 3, … and K can be between
−J and +J:

(6.39)
Substituting the energy eigenvalues (Eqs. [6.34] and [6.38]) back
into the power series expansion for S(θ ) gives the final form of the
rotational wavefunctions:

(6.40)

The functions are referred to as the spherical harmonic


functions that are polynomials in the variables θ and φ and contain
two indexes, J and K, that determine the highest power of the
variables θ and φ, respectively, in the polynomials. Notice that the
index K in the form of these polynomials PJ |K| can have values form
−J to +J, but cannot exceed | J|. The K‐values between −J and +J
account for the 2J + 1‐fold degeneracy of the energy levels
determined by J.
The functions in Eq. (6.40) are known as the “associated
Legendre polynomials” in the variable (cosθ) only but parametrically
contain the integer index K from φ‐dependent part (Eq. [6.34]). The
associated Legendre polynomials form an infinite orthogonal vector
space, as did the Hermite polynomials discussed in Chapter 4. The
first few of these polynomials are given below:

(6.41)
with z = (cos θ). Higher‐order polynomials can be computed from the
recursion formula given below.
The first few spherical harmonic functions follow directly from Eqs.
(6.40) to (6.41), as demonstrated in Example 6.1:

Example 6.1 Derivation of the form of the first few spherical


harmonic functions from

and Equations (6.40) and (6.41)


Answer:

(E6.1.1)

(E6.1.2)

(E6.1.3)

Eq. (E6.1.3) is often written as a linear combination that eliminates


the imaginary function as
The remaining functions can be constructed from the recursion
formula for the associated Legendre polynomials and Eq. (6.42).

The spherical harmonic functions Y(θ, φ) are commonly encountered


in physics and describe, among other situations, the tides on a planet
covered by an ocean of uniform depth or the modes of electrons on a
spherical particle (the electron‐on‐a‐sphere). This latter model is
useful to describe effects such as surface plasmon resonances that
lead to several forms of interesting spectroscopic effects [4].
Like the Hermite polynomials, the functions of degree (J + 1) can be
obtained via a recursion formula from polynomials of degree J and
(J − 1):

(6.42)

Thus, the selection rules for rotational transitions (see Section 6.4)
may be derived in analogy to the derivation of the harmonic
oscillator selection rules. Notice that in the recursion formula for
each subsequent polynomial, the index K does not change.
The eigenvalues of the rotational Schrödinger equation are quite
simple expressions:

(6.43)

where J assumes integer values, J = 0, 1, 2, 3,…. Since the rotational


energy of linear and spherical top molecules depend on the
eigenvalues of the operator only, Eq. (6.43) contains all the
information needed for these special cases. However, for symmetric
top molecules, the eigenvalues of the operator are needed as well,
which are given by

(6.44)
Readers familiar with the quantum mechanics of the hydrogen atom
(see Chapter 7) will recognize that the eigenfunctions and
eigenvalues presented here are the same for the nonradial part of the
hydrogen atom, and the limitations of the indexes of the spherical
harmonics determine the shape and the splitting of the s, p, d, f, and
other orbitals into the sublevels determined by the magnetic
quantum number. K can assume integer values from −J to +J.
Equations (6.12) and (6.43) lead to the surprisingly simple equation
for the rotational energy of spherical and linear molecules:

(6.45)
where B was defined above as

(6.10)

and J = 0, 1, 2, 3,…. The consequences of Eq. (6.40) on the


appearance rotational spectra observed for linear molecules will be
discussed in Section 6.5, after the discussion of selection rules for
these species.
The commutation rules of the Cartesian components of the angular
momentum with each other (see Eq. [6.20]) and the fact that the
total angular momentum operator, commutates with lead to
some interesting aspects of what is known as the “spatial
quantization” of the angular momentum. This aspect will be
introduced in Section 7.6 for the spatial quantization of the
electronic angular momentum in the hydrogen atom. The spatial
quantization affects many aspects of spectroscopy dealing with
electronic and spin angular momenta.

Example 6.2 Analysis of the units of the rotational constants


In rotational spectroscopy of heavy molecules, it is customary to
express the rotational constants in units of Hz (or MHz), since
microwave spectroscopy is carried out using instruments with
radiofrequency generators that are typically calibrated in frequency
units.
Answer:
According to Eqs. (6.5) and (6.6), the units of the moment of inertia
are given as [kg m2]. Therefore, the units of B are obtained as follows

(E6.2.1)

Since E = h ν, the expression in Eq. (6.10) is multiplied by h for


energies to be expressed in units of Joule. Thus, the expression for B
would be

(E6.2.2)

For light molecules, such as H‐F, the rotational constant is


conventionally expressed in wavenumber units. This conversion is
obtained by dividing the frequency by the velocity of light.

6.4 Selection Rules for Rotational


Transitions
The selection rules for rotational transitions are derived from the
recursion formula of the associated Legendre polynomials [5]. The
derivation of these selection rules is carried out in analogy to the
derivation of the selection rule for the harmonic oscillator where the
recursion properties of the Hermite polynomials were employed (cf.
Eqs. [4.34]–[4.36]).
The transition moment between two rotational energy levels J and J′
for the z‐component of the electric dipole transition operator is
written as

(6.46)

As in any form of absorption spectroscopy, one requires this


transition moment to be nonzero:

(6.47)
for a transition to occur. In analogy to the arguments for the
transition moment between harmonic oscillator wavefunctions, one
substitutes the expression in Eq. (6.46) by the recursion
formula

(6.48)

This yields an expression that is written in simplified form (omitting


the fractions containing the parameters J and K) as

(6.49)

(6.50)

Since the associated Legendre polynomials are orthogonal, the two


integrals in Eq. (6.50) are nonzero if and only if

(6.51)

(6.52)

in complete analogy with the arguments presented for the harmonic


oscillator transition moment in Eqs. (4.33)–(4.34). This leads to the
selection rules for a diatomic rigid rotor:

(6.53)
Thus, transitions between adjacent energy levels are allowed, both
up and down the energy ladder.
Although this selection rule is the same as that for the harmonic
oscillator, the spectra are vastly different. In the latter case, only one
peak was predicted due to the equidistant energy levels. In rotational
spectroscopy, a set of spectral peaks is observed due to the fact that
the energy levels are spaced quadratically, and the transition energy
depends on the J‐level from which the transition originates. This will
be demonstrated in the next section.

6.5 Rotational Absorption (Microwave)


Spectra
6.5.1 Rigid Diatomic and Linear Molecules
According to Eq. (6.45), the energy levels for a spherical or linear
molecule are given by

(6.45)
which leads to energy levels

(6.54)
etc.
With the selection rule given by Eq. (6.53), this leads to transition
energies of 2B, 4B, 6B, etc. between adjacent energy levels. This is
shown in Figure 6.3a by the vertical arrows.
In general, the spacing between rotational transitions can be
expressed as follows. Let us assume that transitions occur from state
J to a more energetic state J' = J + 1. Such transitions occur at
transition frequencies of

(6.55)
Thus, the rotational spectrum for a linear molecule consists of
equidistant spectral lines spaced by 2B. This is shown schematically
in Figure 6.3b. The intensity of the spectral lines shown in
Figure 6.3b will be discussed next. (Recall, however, that that
spherical top molecules, although they have exactly the same energy
levels, do not exhibit a rotational absorption spectrum because they
are, by definition, devoid of a dipole moment. The same holds for
homonuclear diatomic molecules.)
The transition moments for the rotational transitions differ by the
factors

(6.56)

in Eq. (6.48) that were omitted in the derivation of the dipole


transition moment (Eqs. [6.49]–[6.52]) and are found to differ only
slightly. Thus, one may argue that the intensity of the observed
spectral lines depends mostly on the population of each of the energy
levels, rather than the differences in transition moments between
different J‐levels. Assuming that the rotational energy differences (a
few wavenumbers) are small compared to the thermal energy (208
cm−1 at room temperature), the Boltzmann distribution (Eq. [3.31])
predicts that many rotational energy levels have similar populations.
The probability PJ of finding a molecule in a given rotational level J
is given by
Figure 6.3 (a) Energy level diagram for linear rotors. (b)
Schematic rotational spectrum for linear rotors.

(6.57)

In this equation, the exponential expression is the normal Boltzmann


distribution. The pre‐exponential expression contains the factor
(2J+1), which results from the K‐fold degeneracy of the rotational
energy (see Section 6.3). The factor B/kT results from the total
number of occupied rotational states, given by
(6.58)

Here, one assumes that the energy levels are so closely spaced that J
becomes a continuous variable. The intensity distribution, as a
function of the rotational transition, is shown schematically in
Figure 6.4. Using Eqs. (6.45) and (6.58), the rotational spectrum for
a diatomic molecule can be predicted, as demonstrated in 6.3.

Example 6.3 Calculation of the expected rotational spectrum of


35Cl–F at room temperature, given

Figure 6.4 Simulated rotational spectrum of 35Cl–F at room


temperature, using B = 0.369 cm−1 and Eq. (6.58) to compute the
populations of each of the J‐levels. The spacing between adjacent
spectral bands is ca. 0.74 cm−1; thus, the spectral envelope shown
covers about 37 cm−1.
and
Answer:
For a diatomic molecule, the equation for the moment of inertia is

(6.5)

or

(E6.3.1)

This relationship is the subject of problem (1) at the end of the


chapter.

(E6.3.2)

(E6.3.3)

(6.10)

(E6.3.4)

(E6.3.5)
The rotational spectrum consists of equidistant lines, spaced by 2B =
0.738 [cm−1]. Next, the intensity profile as a function of J given by
Eq. (6.58) needs to be computed. For this, EXCEL software can be
used advantageously. The result of this simulation is shown in
Figure 6.4. Here, it was assumed that the transition moments for all
J‐levels are equal.

The calculations for the 35Cl–F molecules were carried out using the
rigid rotator approximation. The observed spectrum is slightly
different since the molecule rotates faster at higher values of J and
centrifugal effects increase the bond distance. This, in turn, increases
the moment of inertia, which reduces the rotational constant B;
consequently, the rotational line spacing becomes smaller at
increasing J‐levels.

Figure 6.5 Schematic of the center‐of‐mass (COM) position in an


oblate (a) and prolate (b) symmetric top rotor such as chloroform
and methyl chloride, respectively.

6.5.2 Prolate and Oblate Symmetric Top Molecules


It was pointed out earlier (Section 6.1) that for some molecules, two
of the moments of inertia are equal and different from the third one.
These molecules belong to the C3v point group (see Chapter 11) or
higher (but not to spherical point groups) and are referred to as
symmetric top rotors. Examples of such molecules are chloroform
(CHCl3), ammonia (NH3), or methyl chloride (H3CCl). Depending on
the masses of atoms, the center of mass may lie on either side of the
central atom: in methyl chloride, for example, the center of mass lies
along the C–Cl bond, and the molecule is said to be a prolate top (see
Fig. 6.5b). On the other hand, in chloroform, the center of mass lies
on the symmetry axis below the central carbon atom. This species is
referred to as an oblate top; see Figure 6.5.
For oblate top rotors, the classical rotational energy was found to be
(6.17)

With the eigenvalues of the and operators given in Eqs. (6.43)


and (6.44), the quantum mechanical energy of the oblate top rotor is
given by

(6.59)

For the prolate top rotor, the corresponding equations are

(6.18)

and

(6.60)

An energy level diagram for the oblate and prolate symmetric top
rotor is shown in Figure 6.6a and b, respectively. For symmetric top
rotors, the rotational energy depends on both J and K, as shown.
From the aforementioned convention IA ≤ IB ≤ IC, it follows that A ≥
B ≥ C; therefore, (C–B) is always negative, and (A–B) is always
positive. Thus, the energy levels for increasing K‐values are
decreasing for the oblate and increasing for the prolate top. This
increase/decrease occurs with the square of K.
Figure 6.6 Energy level diagram for (a) oblate and (b) prolate top
rotors. See the text for detail.
A comparison between Figures 6.3 and 6.6 demonstrates the loss of
degeneracy of the rotational energy levels for symmetric top rotors.
In Figure 6.3, for linear or spherical top rotors, the energy levels are
(2J + 1)‐fold degenerate, because K can assume values from −J,
−J+1, 0, …J−1, J.
The discussion of the selection rules based on the recursion
properties of the associated Legendre polynomials (Eqs. [6.49] and
[6.52]) indicated that the quantum number K stays constant when
applying the recursion formula. This leads to selection rules for
symmetric top rotors:

(6.61)
These selection rules require that transitions occur within a given K‐
level, as indicated by the gray arrows in Figure 6.6a.
Therefore, the transition energies for the J = 1 to J = 2 transition for
K = −1, K = 0, and K = 1 are still the same, since the energy levels are
reduced by the same amount, namely, (C − B)K2, yet this transition is
composed of three separate components. Only in the presence of an
external electric field (Stark splitting) will these components be
observed separately. Under these conditions, the transitions will
exhibit multiplet splitting. In the absence of an electric field,
symmetric top rotors exhibit equidistantly lines spaced at 2B, just as
linear molecules.

6.5.3 Asymmetric Top Molecules


Asymmetric top rotors, i.e. molecules with symmetry point groups
lower than C3v, as mentioned above, cannot be treated in a simple
model described for the other tops above. The water molecule that
belongs to the C2v point group (cf. Chapter 11) has three different
moments of inertia and, therefore, is an asymmetric top rotor. Since
the components of the angular momentum operator do not
commutate, approximate methods are used to interpret the
microwave spectrum of asymmetric top rotors. This is accomplished
by describing the molecule by its “asymmetry parameter κ,” which is
defined as

(6.62)

This parameter takes the value of −1 for the prolate top (with B = C)
and +1 for the oblate top (A = B). Thus, transition energies E(κ) are
interpolated for an asymmetric top rotor between the oblate and
prolate top limiting cases, and the overall rotational energy is written
as E(J, κ):

(6.63)

The rotational spectra of asymmetric top rotors are quite


complicated and beyond the intent of this book. Interested readers
are referred to specialized books such as Townes and Schawlow [5].

6.6 Rot–Vibrational Transitions


Infrared and Raman vibrational spectra of gaseous molecules exhibit
broad wings on the low‐ and high‐frequency side of the
corresponding vibrational transitions, as shown in Figure 6.7a. These
wings are due to rotational transitions superimposed on vibrational
transitions and show the distinct rotational–vibrational (rot–
vibrational) fine structure at high spectral resolution, see
Figure 6.7b. To describe the origin of these transitions, one writes
the rot–vibrational wavefunctions as products of the pure rotational
and the pure vibrational wavefunctions:

(6.64)
and the corresponding energy levels simply as the sum of rotational
and vibrational energies:

(6.65)
For a diatomic molecule obeying the harmonic oscillator and rigid
rotor approximation, Eq. (6.65) would read as

(6.66)

In Eq. (6.66), both the vibrational and rotational energies are


expressed in units of energy; that is, B is expressed in Hz and the
vibrational energies in terms of the frequency (rather than the
wavenumber) of the transition. The energy level diagram for such a
diatomic molecule is shown in Figure 6.8a.
Figure 6.7 (a) Observed rot–vibrational band envelopes in the
infrared absorption spectrum of gaseous methane, CH4. (b) High‐
resolution spectrum of the region shown in (a).
Figure 6.8 (a) Rot–vibrational energy level diagram for a
harmonic oscillator/rigid rotor. (b) Rot–vibrational transitions with
Δn = ± 1; ΔJ = ± 1 selection rules. (c) Spectral lines corresponding
to these transitions.
The selection rules for rot–vibrational transitions are

(6.67)
This leads to rot–vibrational transitions from the rotational sublevels
of the ground vibrational state to rotational sublevels of the
vibrationally excited state with either lower or higher rotational
quantum numbers, as shown in Figure 6.8b.
Inspection of Figure 6.7b also reveals that the spacing between the
rotational transitions becomes smaller from lower to higher
wavenumber. This is due to the centrifugal distortion effect: as the
molecule rotates faster, centrifugal forces stretch the bond, thereby
increasing the moment of inertia and decreasing the rotational
constants, resulting in more closely spaced lines. This effect, as well
as the contribution of anharmonicity, is contained in Eq. (6.68):

(6.68)

The anharmonicity term in Eq. (6.68) was defined


before (cf. Eqs. [4.45] and [4.46]). The centrifugal distortion
constant D accounts for the change of the moment of inertia at
higher rotational energy levels.
The rot–vibrational transitions for which ΔJ = −1 (the low‐frequency
progression of rot–vibrational bands) are referred to as the P‐
branch, those with ΔJ = 0 are referred to as the Q‐branch, and those
with ΔJ = +1 as the R‐branch. Symmetry rules sometimes prohibit
the occurrence of the Q‐branch, and typical rot–vibrational
absorption spectra, for example, for gaseous H‐Cl, may appear as
shown in Figure 6.8c. Notice that there are many transitions in the
Q‐branch, since the Δn = 1, ΔJ = 0 transitions can originate from J =
0, J = 1, J = 2, etc.
In linear polyatomic molecules, vibrational transitions that do not
change the dipole moment are forbidden in absorption; however,
transitions that change the dipole moment, such as the
antisymmetric stretching vibration of CO2 (cf. Chapter 5), do exhibit
a rot–vibrational spectrum. For this vibration, the transition dipole
lies along the molecular axis; consequently, the band envelope is
referred to as a “parallel” envelope, characterized by the absence of a
Q‐branch, with an appearance similar to the rot–vibrational
spectrum of a heteronuclear diatomic molecule. This is shown in
Figure 6.9b. In the CO2 deformation mode, on the other hand, the
dipole moment change is perpendicular to the molecular axis;
consequently, the resulting band envelopes are referred to as
“perpendicular” bands. Perpendicular bands have a pronounced Q‐
branch, broadened by the overlap of the n = 0 to n = 1 transitions for
different J‐values (cf. Figure 6.9a).

Figure 6.9 Simulated rot–vibrational spectral band profiles for


the deformation mode (a) and the antisymmetric stretching mode of
CO2 (b). [From vpl.astro.washington.edu/spectra/co2/htm].
Although spherical top molecules such as CH4 do not exhibit a pure
rotational absorption spectrum, the triply degenerate modes such as
the antisymmetric stretching mode exhibit a rot–vibrational
spectrum that permits the determination of the rotational constant
B. This vibration exhibits a rot–vibrational spectrum of the
perpendicular type, with a distinct Q‐branch at ca. 3019 cm−1 as
shown in Figure 6.7.
References
1 Kwon, Y.H. Mechanical Basis of Motion Analysis: Inertia Tensor.
http://www.kwon3d.com/theory/moi/iten.html.
2 Levine, I. (1970). Quantum Chemistry, vol. I&II, 83. Boston:
Allyn & Bacon.
3 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon
Chapter 5.
4 Zeng, S. et al. (2014). Nanomaterials enhanced surface plasmon
resonance for biological and chemical sensing applications.
Chemical Society Reviews 43 (10): 3426–3452.
5 Townes, C.H. and Schawlow, A.L. (1975). Microwave
Spectroscopy. New York: Dover Publications.

Problems
1. Demonstrate that [L2, Lz] = 0.
2. The classical definition of the moment of inertia for a diatomic
molecule is

(6.5)

where r1 and r2 are the distances of masses m1 and m2 from the


center of mass. Show that Eq. (6.5) is equivalent to where
mR is the reduced mass of atoms 1 and 2 and r12 is the bond length.

3. In the linear, centrosymmetric molecule acetylene C2H2, the two


carbons are at ±60 pm, and the two hydrogens are at ±67 pm
from the center of mass.
a. Calculate the moment of inertia for acetylene.
b. Calculate the rotational constant.
c. What is the spacing of the rot–vibrational lines observed for
the antisymmetric, infrared active C–H stretching mode of
at 3295 cm−1?
4. The spacing of rotational lines in the spectrum of H–35Cl is
21.173 cm−1. Do you expect the spacing of lines in H–37Cl to be
lower or higher? Justify your answer.
5. The rotational constant for deuterium fluoride, 2D–19F, is
11.007 cm−1. The atomic masses for 19F and 2D are 18.9984032
and 2.0141018 amu, respectively. Calculate the bond length of
deuterium fluoride to the maximum number of significant
figures consistent with this information.
6. Given the selection rule for a rigid rotor, ΔJ = ±1, show that its
rotational spectrum consists of equidistant lines spaced by 2B.
7. Although the rotational Schrödinger equation cannot be solved
exactly for an asymmetric rotor such as water, experimental
observation of its microwave spectrum gives the rotational
constants, using the approximate methods outlined in Section
6.5.3.
a. Use the definition of the inertial tensor (Eq. [6.6]) and the
following coordinates of the atoms in a water molecule to
calculate its rotational constants.
b. According to the results in part (a), where does the water
fall in terms of its κ‐value?
Atomic coordinates (in pm)
X Y Z
O 0.0 6.52 0
H1 −75.8 −54.5 0
H2 75.8 −54.5 0
mO = 16.0 amu mH = 1.00 amu
7
Atomic Structure: The Hydrogen Atom
Although this book was conceived as an introduction to the quantum
mechanical foundations of molecular spectroscopy and the hydrogen
atom hardly qualifies as a molecule, this chapter is included for
several reasons. First, a quantum mechanical textbook just isn't
complete unless the hydrogen atom is given some detailed
discussion. Second, the work presented in Chapter 6 on the solution
of the rotational Schrödinger equation is very applicable to solving
the hydrogen atom Schrödinger equation; in fact, the graphical
representation of the spherical harmonic functions makes more
sense in the context of orbital shapes than for the rotational states of
molecules. This is so at least in the eyes of the author who cannot
imagine a rotating molecule to have rotational energy states shaped
like spherical harmonic functions. Third, the concept of electronic
orbitals – and particularly their shapes – is of importance in the
discussion of chemical bonding that needs to be introduced in order
to discuss molecular electronic spectroscopy. Finally, it was the
hydrogen atom to which quantum mechanics was first applied, and it
cannot be overemphasized that it was the ingenious insight of E.
Schrödinger who realized that by substituting the classical
momentum by its quantum mechanical equivalent (see Eq. [2.2]), a
set of differential equations was obtained that described the
hydrogen atom. Schrödinger, trained originally as a classical wave
mechanicist, knew these differential equations before applying them
to the hydrogen atom and intuitively knew that the form of their
solutions would explain atomic electronic structure. This discovery
forever changed science and brought order into the periodic chart of
elements that, at this point, was a purely empirical correlation of
chemical properties of elements without rhyme or reason.
This latter aspect is in line with the philosophy of this book: rather
than approaching the subject of quantum mechanics as a theoretical
concept, the view taken here is to connect experiment and theory.
The experimental observations underlying the discussion in this
chapter are the previously discussed atomic emission and absorption
spectra (see Section 1.4) that were fit to an empirical, yet
unexplained equation (cf. Eq. [1.19]) that explains the now well‐
established form of the periodic chart that is the basis of all
chemistry. As we shall see in the present chapter, the organization of
the periodic chart is a direct consequence of the solutions of the
angular part of the Schrödinger equation and explains the number of
elements in each row in the periodic chart. Again, for the author, it is
nearly miraculous how the form of the spherical harmonic functions
and the rules that control the indexes in these functions – the
quantum numbers in the parlance of scientists – are responsible for
all aspects of chemistry.

7.1 The Hydrogen Atom Schrödinger


Equation
As indicated in Section 2.2, an electron in the potential field of a
nucleus experiences a potential energy that is given by Eq. (7.1):

(7.1)

Equation (7.1) is the integrated form of Coulomb's law. Here, “e” is


the electronic charge in Coulomb (C) (see Appendix 1), and ε0 is the
dielectric constant (permittivity) of vacuum. (Notice that the
electronic charge “e” is written as a non‐italicized quantity, whereas
the basis of the natural growth function ex is written in italics.) A
quick analysis of the units in Eq. (7.1) is appropriate at this point.
The permittivity has the numerical value of ε0 = 8.854 × 10−12
[F/m = C2/(J m)]. Thus, the right‐hand side of Eq. (7.1) has units of

(7.2)

which are indeed units of energy.


The potential given in Eq. (7.1) has spherical symmetry, that is, for
any constant value of r, the potential energy is equal for all values of
the angles θ and ϕ (see Figure 6.1). Thus, the Schrödinger equation
for the electron in the potential field of the nucleus is best defined in
spherical polar coordinates. In Chapter 2, the Schrödinger equation
for a one‐dimensional case of an electron in a (negative) hyperbolic
potential of a nucleus was written as

(2.16)

With the condition of the potential having spherical symmetry


discussed in the previous paragraph, Eq. (7.1) needs to be substituted
into Eq. (2.16), and the total equation recast into spherical polar
coordinates:

(7.3)

Here, mR is the reduced mass of the proton–electron system, and ∇2


is the Laplace operator (or the Laplacian) given by

(7.4)

in Cartesian or

(7.5)

in spherical polar coordinates [1]. The angle‐dependent part of Eq.


(7.4) is just the operator that was discussed in Chapter 6:
(6.22)

and was obtained there by assuming a constant value of r for the


rigid rotor.
Thus, the ∇2 operator will have eigenfunctions that can be separated
into distance‐ and angle‐dependent functions according to

(7.6)

where the θ‐ and ϕ‐dependent parts are the eigenfunctions of the


operator that are the spherical harmonic functions Y(θ,ϕ).
Furthermore, the reduced mass of the proton–electron system is
nearly equal to the electron mass, since mp + me ≈ mp. Thus, Eq. (7.3)
then can be written as follows:

(7.7)

where we used the eigenvalues ħ2l(l + 1) of the operator. For


historical reasons, the quantum number of the total angular
momentum operator that was referred to as “J” in the discussion of
rotational spectroscopy (cf. Eq. [6.38]) is called “l” in the discussion
of the hydrogen atom. Since this quantum number describes the
component of the angular momentum of the electron orbiting the
nucleus, this quantum number l is usually referred to as the “orbital
angular momentum” quantum number. Similarly, the quantum
number of the operator that was referred to as “K” in rotational
spectroscopy (Eq. [6.34]) is designated “m” or “ml” in the discussion
of the hydrogen atom and referred to as the “magnetic” quantum
number for reasons we shall see in Section 7.6. The fact that the
angular part of the hydrogen atom Schrödinger equation is the same
solutions as the rigid rotor problem is not too surprising: after all, we
may envision the proton–electron pair as just a rigid rotor tumbling
in space.
Defining a new constant “a” (which happens to be the “Bohr radius”;
see Example 7.1)

(7.8)

we may rewrite Eqs. (7.5) and (7.6) as

(7.9)

where R″ and R′ are the second and first derivative of R(r) with
respect to r. Equation (7.9) is known as the LaGuerre differential
equation, and its solutions (to be presented below) are the LaGuerre
polynomials. The fact that the eigenvalues of the operator are
explicitly contained in Eq. (7.8) leads to the appearance of the
quantum numbers l and ml in the hydrogen atom wavefunctions.

7.2 Solutions of the Hydrogen Atom


Schrödinger Equation
The process of solving Eq. (7.9) follows a similar but even more
complicated approach that was introduced for the Hermite and
Legendre differential equations. This approach involves a power
series expansion:

(7.10)
where

(7.11)

and “s” a yet undefined integer value. The expansion coefficients are
given by the recursion formula

(7.12)

where n is an integer (see Eq. [7.13]). Complicated arguments on the


value of R(r) for large values of r in the trial function, Eq. (7.10), lead
to the requirement of terminating the power series expansion after a
finite number of terms. As in the case of the Hermite and Legendre
differential equations, this leads directly to the allowed energy
eigenvalues for the hydrogen atom:

(7.13)

Here, Ry is the Rydberg constant introduced in Eq. (1.19) (see


Example 7.1 in which the numerical value of Ry and a are calculated)
and “n” the main quantum number that is restricted to the values n
≥ l + 1.
The eigenvalues of the angle‐dependent part of the hydrogen atom
Schrödinger equation are exactly the same as those for the rigid
rotor, given by Eqs. (6.43) and (6.44):

(6.43)

(6.44)
where the symbols for the quantum numbers have been changed
from J and K to l and m, as pointed out in Section 7.1.
Like Eq. (6.39) that restricted the values of the J and K, the quantum
numbers determining the eigenvalues, the hydrogen atom
eigenfunctions are restricted as follows:

The solutions of Eq. (7.5)

(7.5)
then can be written as

(7.14)

where

(6.40)

and Rnl(r) are the radial part of the LaGuerre polynomials given by

(7.15)
A plot of the radial parts of the wavefunctions is shown in
Figure 7.1a. Remember that these radial parts have spherical
symmetry. Thus, in any direction from the nucleus, the radial
wavefunction drops off as shown in Figure 7.1, and the wavefunction
has a cusp at the nucleus as shown in Figure 7.1b.
The first few spherical harmonic functions given (see Example 6.1)
by

(7.16)

are shown in Figure 7.2. The hydrogen atom wavefunctions are then
given by the product of Eq. (7.15) and (7.16). Using the convention of
designating l = 0 wavefunctions as “s,” l = 1 wavefunctions as “p,”
and l = 2 as “d” orbitals, the complete wavefunctions for the
hydrogen atom are obtained as follows:
(7.17)

Figure 7.1 (a) Plot of radial part of hydrogen wavefunctions in


units of r/ao. (b) Cusp of the “s” orbital at the origin. Since the radial
part of the wavefunction has spherical symmetry, the decrease in the
wavefunction occurs in all directions from the origin.
The orbital energies and degeneracies due to the rules governing the
allowed quantum numbers n, l, and ml are shown in Figure 7.3.
The hydrogen wavefunctions presented in Eq. (7.17) form an
orthonormal vector space, which is an important consideration for
the later discussion of many‐electron systems such as molecules,
where the molecular orbitals are approximated by linear
combination of the hydrogen‐like wavefunctions (see Postulate 5 in
Section 2.1). The hydrogen wavefunctions in Eq. (7.17) also hold for
all one‐electron ions, such as He+, Li2+, etc., with the atomic number
Z entering Eq. (7.1):

(7.18)

Consequently, every occurrence of the expression in Eq. (7.17)

needs to be changed to for all other one‐electron systems.


Accordingly, the ψ200 = ψ2s wavefunction, for example, reads

(7.19)

for a one‐electron ion with an atomic number Z.


Figure 7.2 Plot of first few spherical harmonic functions. Notice
that the and the functions have the same shape
but differ only in phase. The same holds for the and the
functions. Modified from
https://math.stackexchange.com/.
Source: Modified from Stack Exchange, plot of first few spherical harmonic functions.
Figure 7.3 Orbital energy eigenvalues and degeneracies for the
hydrogen atom.
Equation (7.17) and Figure 7.1 demonstrate that for “s” orbitals, the
hydrogen wavefunctions have a maximum at r = 0, namely, at the
cusp shown in Figure 7.1b. This may (erroneously) be construed as
suggesting that the highest probability of finding the electron in a
hydrogen 1s orbital is at the nucleus. This will be discussed next.
The probability P of finding an electron in the 1s orbital in the
volume element dτ, where is

(7.20)
is expressed as the square of the wavefunction times the volume
element:

(7.21)
Figure 7.4 Radial part of the wavefunctions (dashed lines) and
radial distribution functions (solid lines) for 1s (black) and 2s (gray)
orbitals.
To determine the probability of the electron to be found in a thin
spherical shell of thickness dr, independent of θ and ϕ, one needs to
integrate Eq. (8.21) over θ and ϕ to obtain

(7.22)

(7.23)
since the spherical harmonics are normalized. The expression
r2 (R10)2 is called the radial distribution function and presents the
probability of finding the electron in a shell with thickness r + dr,
normalized with respect to the volume of the shell. At r = 0, the
volume of this shell is zero; thus, the probability of finding the
electron at the nucleus is zero. The argument presented here holds
for all orbitals, not just the “s” orbitals. A plot of the radial
distribution function is given in Figure 7.4, which shows that the 1s
orbital has its maximum at the Bohr radius (see Example 7.2).
The orbital shapes rendered in every chemistry textbook are created
by drawing a (usually a 90 %) contour of the probability of finding
the electron in this region of space. Although these orbital shapes
have been predicted decades ago and have become an integrated part
of many branches of chemistry, it was not until the last two decades
that experimental verification of these shapes was achieved [2],
although some of the interpretation of this work has to be taken with
a grain of salt (see [3]).

Example 7.1 Computation of numerical values and analysis of the


units of the Bohr radius and the Rydberg constant from Eqs. (8.7) to
(8.12)
Answer:
The Bohr radius a, as defined in Eq. (8.7), was just a number of
constants. Here, we'll show that a, indeed, has units of inverse length
and determine its numerical value:

(7.7)

(E7.1.1)
(E7.1.3)

(E7.1.4)
Now, for the Rydberg constant:

(7.12)

(E7.1.5)

(E7.1.6)

(1.19)

Thus, the quantum mechanically derived energy eigenvalues


perfectly agree with experimental data from the hydrogen
absorption/emission spectra.

We shall now turn to another example, namely, the determination of


the maximum of the radial distribution function that should, for the
1s orbital, just be the Bohr radius. This is accomplished as follows in
Example 7.2:
Example 7.2 Calculation of the maximum probability of finding a
1s electron.
Answer:
The probability of finding the electron at any distance r is given by

(E7.2.1)

Since the integration over the angle‐dependent part of the


wavefunction gives a factor of 1 due to the normalization condition of
the spherical harmonics, we may use here just the radial distribution
function of the 1s electron, which is

(E7.2.2)

To determine the maximum of the probability of finding the electron,


we differentiate the expression in Eq. (E7.2.2) with respect to r and
set the derivative to zero:

(E7.2.3)

(E7.2.4)

from which follows that r = a, as expected.

7.3 Dipole Allowed Transitions for the


Hydrogen Atom
Next, the selection rules that determine which transitions are
allowed for the hydrogen atom will be discussed. As introduced
before, an electric dipole transition is allowed if
(3.20)

Equation (3.20) implies for the hydrogen atom:

(7.24)

The selection rules for the angle‐dependent part of the wavefunction


were calculated previously:

(6.61)
and are the same here. There are no restrictions to the changes in the
main quantum number n:

(7.25)
Equation (7.25) can be proven rigorously by calculation of the
transition moment using just the radial part of the wavefunctions.
However, it can also be visualized easily by considering that the
distance between nucleus and electron increases for increasing
values of n. Such an increased distance between the nucleus and
electron is, of course, a manifestation in the change of magnitude of
the dipole moment.
All changes in n are allowed and give rise to the different spectral
series (Lyman, Balmer, Paschen, Brackett) that originate from the
states n1, n2, n3, n4… with Δn = ± 1, ± 2, ± 3…. values, respectively.
However, since the quantum number l has to change according to
Eq. (6.61), the energy level diagram for the hydrogen atom
transitions appears as shown in Figure 7.5

7.4 Discussion of the Hydrogen Atom


Results
At this point, it may be appropriate to pause and summarize what
was learned so far in the discussion of the hydrogen atom.
Figure 7.5 Energy level diagram and allowed electronic
transitions for the hydrogen atom, demonstrating the selection
rules Δl = ± 1.
The task of solving the hydrogen atom problem was attacked using
the same approach that was described earlier for the particle in a
box, the harmonic oscillator, and the rigid rotor: first, the
Schrödinger equation was set up with the appropriate potential
energy function. The differential equation was solved easily in the
first scenario, but from the harmonic oscillator on, a power series
expansion was required to solve the Hermite and Legendre
differential equation. In both these cases, rather involved arguments
were required to ensure that the resulting eigenfunctions were finite,
which led to the requirement of truncating of the power series
expansions. The recursion equation for the expansion coefficients of
the power series led to the rather simple equations of the energy
eigenvalues in both the harmonic oscillator and the rigid rotor. In the
hydrogen atom, the power series expansion was required for both the
radial and the angular part of the wavefunction, the latter of which
was the same for the proton–electron system and the rigid rotor.
Finally, substituting the energy eigenvalues into the trial solution of
the differential equation resulted in the final form of the
wavefunctions.
The order of the progression presented in this book follows the
complexity of the Schrödinger equation for each problem. When one
excludes the pure model case – the particle in a box – one finds the
validation of the approach by comparison of theory with the
experimental results. The harmonic oscillator presents a single peak
in the vibrational spectra, the rigid rotor a progression of peaks in
the rotational spectra, and the hydrogen atom the hydrogen
emission/absorption spectra, as predicted from the energy
eigenvalues and selection rules. Small deviations between observed
and the idealized models could readily be explained and corrected
for, by next‐level approximations such as the anharmonic oscillator
and the nonrigid rotor models.
In the hydrogen atom, the quantum mechanical solutions match
perfectly with the experimental results and support the quotation by
Feynman about the necessity of agreement between experiment and
theory. Finally, the argument raised in Section 7.1 that the quantum
mechanical results for the hydrogen atom brought order into
chemistry (that is, the periodic system) is exemplified by Figure 7.3.
Although not quite in its final form for many‐electron systems, this
figure predicts the number of elements in each row of the periodic
chart and, as we shall see later, contains the concept that certain
elements will exhibit similar chemical traits. It is awe‐inspiring that
the mathematical conditions that relate the quantum numbers,
namely, the allowed values of l and m for each n‐level, dictate the
appearance of and the order within the periodic chart and thereby
the entire field of chemistry. Perhaps, the true genius of
Schrödinger's work was the realization that the formalism presented
in Sections 7.1–7.3 determines the structure of matter like no other
discovery.

7.5 Electron Spin


The discovery of electron spin is another chapter in the history of
science that is fascinating. It emphasizes not only the connection
between theory and experiment but also how the opinion of older,
well‐established scientists can suppress new and ultimately correct
ideas of young scientists proposing new concepts. This aspect is
nicely demonstrated in two review articles on the discovery and
theoretical formulation of electron spin [4, 5].
It was stated above that the quantum mechanical description of the
hydrogen atom brought order into the field of chemistry in the mid‐
1920, when this work was performed and published. However, it was
again an unexplainable experimental observation that required
refinements of this model. This experimental work was based on the
observation of atomic emission spectra in the presence of a magnetic
field that needs a short introduction.
When the hydrogen atom emission spectrum was carried out in the
presence of a static magnetic field, it was found that some of the
spectral lines split into multiple lines or “multiplets.” This is known
as the Zeeman effect that had been well‐established by the mid‐1920,
and at first, the splitting seemed to be in perfect agreement with the
new quantum mechanical interpretation of the hydrogen atom, as
can be seen from the following argument. The presence of a magnetic
field lifts the degeneracy of the 2p, 3p, 3d, and other sublevels that
have different “magnetic” quantum numbers ml, as shown in
Figure 7.6a: p orbitals split into three and d orbitals into 5 levels,
according to their ml value. The energy splitting of these sublevels is
given by

(7.26)

where B is the magnetic field (in Tesla, T), “e” the charge, and me the
mass of the electron. Furthermore, in the presence of a magnetic
field, the selection rule stated above (Eq. [6.61]) needs to be
augmented to read

(7.27)

Figure 7.6 (a) Energy level diagram of the hydrogen atom orbitals
in the presence of a magnetic field. (b) Energy level diagram for the
sodium 589 nm doublet. See text for detail.
The energy level diagram shown in Figure 7.6a is obtained, and the
Balmer series transition from 3d into the 2p orbitals are split into
triplet with transition frequencies ν + Δν, ν, and ν − Δν, where ν is the
frequency of the transition in the absence of a magnetic field and Δν
is the splitting between the energy levels of different ml values (see
Eq. [7.26]). This is shown by the vertical lines in Figure 7.6a. The
splitting into three components is due to the selection rule Δl = ± 1
for the electron's angular momentum quantum number and
Δml = 0, ± 1 for the magnetic quantum number. Thus, there are only
three observed spectral lines although there are nine different
transitions.
However, in other atoms, the splitting pattern of spectral lines
cannot be explained by the discussion above. A prime example for
this was found in the so‐called sodium D line (at 589 nm) of the
sodium emission spectrum, which results from a transition between
the excited sodium state (1s2 2s2 2p6 3p1) to its ground state (1s2 2s2
2p6 3s1). The exact meaning of this notation – the atomic electron
configuration – will be explained in Section 9.3. Since this transition
occurs from a 3p state with threefold degeneracy to a 3s state with no
degeneracy and because of the selection rule Δl = ± 1, two transition
are observed, namely, the famous sodium doublet at c. 589 nm (see
Figure 7.6b). This will be discussed in more detail using atomic term
symbols in Chapter 9 (see Example 9.1).
However, in the presence of a magnetic field, one of the sodium
doublets splits into four and the other into six components. The only
explanation for this “anomalous Zeeman” effect was that the electron
undergoing this transition itself had two energy states. At the point
in time when the anomalous Zeeman effect was first reported, the
results of the Stern–Gerlach experiment were known, which also had
suggested that electrons could have two inherent energy states. Since
these energy states seemed to be caused by two different orientations
of the electron's magnetic moment and since a rotating charge causes
a magnetic moment, these two energy states were (somewhat
misleadingly) referred to as “spin states” of the electron, implying
that the electron was actually spinning about an axis. Thus, one may
be tempted to visualize the two possible energy states as an electron
spinning either clockwise or counterclockwise to create a magnetic
moment that aligns parallel or antiparallel to an external magnetic
field. However, this classical view is strictly a visualization, and the
electronic spin should be viewed as an inherent property of
fundamental particles. Other approaches to quantum mechanics,
noticeably those by Heisenberg and Dirac, directly predict the spin
quantum number as a fourth required parameter to specify an
electron in a hydrogen atom [5].
Since the electron exhibits two distinct energy states, there must be
an operator that has these two eigenstates as solutions. This operator
is referred to as the spin angular momentum operator. In analogy to
the orbital angular momentum operator and its Cartesian
components , one defines a spin angular moment
operator and its components with the same
commutation properties as the orbital angular momentum operator
(see Eqs. [6.20] and [6.21]).1
For electrons, the eigenvalues of the operator are

(7.28)

With the the allowed spin quantum number , the eigenvalues


become

(7.29)

For the operator, the eigenvalues are

(7.30)

in analogy to m that can have values from −l to +l. The


corresponding electron spin eigenfunctions are referred to as α and β
such that
(7.31)

Because of the commutation rules between the and the


operator, the eigenfunctions described by Eq. (7.31) are
eigenfunctions of the operator as well, with eigenvalues (see
Eq. [7.29]).
The consequences of the discussion in the last section are that the
electronic wavefunctions for the hydrogen atom need to be
augmented by a spin function, that is,

(7.32)
The electron spin results presented in this section do not change the
outcome of the quantum mechanical treatment of the hydrogen atom
that can be adequately described by three quantum numbers, but it
directly affects any system containing more than one electron, both
in atoms such as the He atom and up, as well as any molecules
containing more than one electron, such as H2 and up. However, it is
interesting to note that the relativistic quantum mechanical
approach by Dirac includes the electron spin quantum number
explicitly. The electron's spin states lead to the splitting of atomic
spectra in a magnetic field. Again, experimental results led to the
refinement of the earlier theories.
Figure 7.7 Spatial, or orientational quantization of the orbital
angular momentum for (a) l = 1 and (b) for l = 2.
Source: Goudsmit [4]; Commins [5].

As will be shown in Chapter 8 on nuclear magnetic resonance


spectroscopy, nuclei themselves can have an intrinsic spin angular
momentum, which however is much smaller than that of the
electron. Thus, the interaction between nuclear and electronic spin
can be ignored for the hydrogen atom.

7.6 Spatial Quantization of Angular


Momentum
Since the eigenvalues of the operators cannot be
determined simultaneously, the exact location the eigenvalue of the
operator, cannot be determined either,
although its length is know exactly. This can be seen from the
following argument. Consider a Cartesian coordinate system
oriented along the axes of the operators, as shown in
Figure 7.7a. Since the eigenvalue of , namely, ±mħ, is known and
lies along the z‐axis, the |l| vector must lie on a cone of loci with
radius l(l + 1) − m2 (in units of ħ). This follows directly from
Pythagorean's theorem.
In Figure 7.7, the possible values of m (−1, 0, +1) for l = 1 and (−2, −1,
0, 1, 2) for l = 2 are shown along the z‐axis. In the former case, the
length of |l| is and in the latter . This
spatial quantization of the angular momentum will be encountered
again in spin spectroscopy, such as electron paramagnetic or nuclear
magnetic resonance (NMR) spectroscopy. The net magnetization
encountered in NMR is a sum of the spin angular momenta of each
nuclear spin “s” (see Eq. [7.26]), which can be depicted to precess
along a conical pattern about the z‐direction that is the direction of
the applied external magnetic field.

References
1 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.

2 Wang, S.G. and Eugen Schwarz, W.H. (2000). On closed‐shell


interactions, polar covalences, d‐shell holes, and direct images of
orbitals: the case of cuprite. Angewandte Chemie International
Edition 39 (10): 1757–1762.
3 Scerri, E.R. (2000). Have orbitals really been observed. Journal
of Chemical Education 77 (11): 1492–1495.
4 Goudsmit, S.A. (1971). De ontdekking van de electronenrotatie.
Nederlands Tijdschrift voor Natuurkunde 37: 386–394.
5 Commins, E.D. (2012). Electron spin and its history. Annual
Review of Nuclear and Particle Science 62: 133–157.

Problems
Qualitative Questions
1. What are the four quantum numbers of an electron in a
hydrogen‐like orbital called?
2. With which variable of the hydrogen atom Schrödinger equation
is each of the quantum numbers associated?
3. What are the allowed values for the four quantum numbers for
an electron in a hydrogen‐like orbital, and what physical
properties are designated by each of the four quantum
numbers?
4. What is the degeneracy of the hydrogen orbitals with the
quantum number l?
5. Which two parameters determine the size of a hydrogen‐like
orbital in one‐electron systems?
6. The wavefunctions of all s orbitals all have peak amplitudes at
the nucleus. What mathematical procedure was invoked to
arrive at a spherical electron distribution that confirms the
experimental observation of the electron found most likely at the
Bohr radius?
7. Describe briefly the integral that had to be solved to answer
question 6.

Quantitative Questions. For the following problems, use


.

1. Normalize the H‐atom 2s orbital, for which ψ200 = N (1−r/2ao)


e−r/2ao.
2. In Example 7.2, the maximum of the radial distribution function
for the 1s orbital was found to occur at the Bohr radius. Perform
the same calculations for radial distribution function of the 2s
orbital. This will result in a quartic equation that can be solved
graphically. Discuss the results.
3. For the H‐atom 1s orbital, the maximum of the radial
distribution function occurs at ao (see problem 2). Compare this
result with another measure of the electron's most probable
location, the expectation value 〈r〉 of the electron's position.
4. Show that the H 1s and 2s wavefunctions, ψ100(r) and ψ200(r),
are orthogonal.

Note
1 The qualitative discussion of the spin operators and and the
corresponding spin functions is sufficient at this point. However,
a more detailed background of the spin operators and functions is
given in Appendix 5.
8
Nuclear Magnetic Resonance (NMR)
Spectroscopy

8.1 General Remarks


There are two major spectroscopic techniques that are based on the
observation of transitions due to reorientation of spins in a magnetic
field when exposed to electromagnetic radiation. These are electron
paramagnetic resonance (EPR) and nuclear magnetic resonance
(NMR) spectroscopies. In the former, the transitions are between
energy states that an unpaired electron spin experiences in an
external magnetic field. Some principles of the electron spin were
introduced in the previous chapter.
NMR spectroscopy, on the other hand, is due to nuclear spin energy
states that have not yet been introduced in previous discussions.
Whereas EPR requires the presence of a radical molecule or ion (i.e.
one with an unpaired electron), NMR is observed for any atom with a
nonzero nuclear spin and has been developed mostly for 1H and 13C
atoms. One can safely say that no spectroscopic method used in
chemical and biochemical research has seen as explosive a growth
over the past five decades as has NMR spectroscopy that has become
a cornerstone for researchers for identification of molecules and to
establish molecular structure in solution and in the solid state. A
major reason for this growth was the introduction of pulse Fourier
transform (FT) methodology in the 1970s (to be discussed in
Appendix 4) over continuous excitation that was common before that
time. The advent of FT methodology itself was the result of the
availability of low cost and dedicated instrument control computers
and software to perform the FT rapidly. One other advantage of pulse
FT methodology is the ability to program pulse sequences that excite
specific aspects of the magnetization of the sample that lead to
several new forms of spin interactions that reveal structural details
not available in continuous excitation. One further reason for the
growth of NMR spectroscopy was the available of high (magnetic)
field superconducting magnets that enormously improved spectral
resolution and sensitivity of NMR methods.
In this chapter, the principles of spin spectroscopy will be introduced
for 1H and 13C NMR spectroscopy. Many of the principles of 1H NMR
can be applied to EPR spectroscopy that will not be discussed any
further. Even 1H NMR will be treated at an introductory level only
since entire volumes are necessary to treat all aspects of NMR
spectroscopy.

8.2 Review of Electron Angular Momentum


and Spin Angular Momentum
Spin resonance spectroscopy deals with the “spin angular
momentum” of an atomic nucleus or an electron and, therefore, with
the “spin Hamiltonians.” In classical physics, the angular momentum
L is used to describe the rotational kinetic energy of a spinning body:

(6.1)

where I, the moment of inertia, is defined as

(6.5)

Equation (6.1) relates angular momentum and moment of inertia


just as the linear momentum and mass are related for linear motion:

(1.17)

The angular momentum is defined in classical physics as

(6.2)
which is shown in Figure 8.1a. Equation (6.2) leads to the classical
definition of the Cartesian components of L according to

(6.3)

with

(6.4)
and similar expressions holding for the other components of L. The
quantum mechanical expression for the component of the
angular momentum operator given in Eq. (6.4) then becomes

(6.19)

Figure 8.1 (a) Definition of the angular momentum in terms of


radius r and linear momentum p. (b, c) Depiction of spin angular
momentum eigenvalues of the operator (solid arrows) and
operator (dashed arrows) for spin wavefunctions α (panel B) and β
(panel C).

As discussed before, the Cartesian components , , and do


not commutate, and, therefore, their eigenvalues cannot be
determined simultaneously. They do commutate, however, with the
operator for the total angular momentum, which has eigenvalues

(8.1)
The eigenvalues of one of the Cartesian components (usually ) are
given by

(8.2)
with allowed values of K between –J and +J.
As we have seen, the quantization of the orbital angular momentum,
or “spatial quantization” (see Section 7.6), resulting from Eq. (8.1)
has a major influence on the quantum mechanical treatment of the H
atom. Furthermore, we saw that an electron has its own angular
momentum, referred to as its spin angular momentum. Incidentally,
the interaction of orbital and spin angular momentum determines
the energetics and spectral behavior of multi‐electron systems, as we
shall see in Chapter 9.
We now return to the discussion of the intrinsic spin angular
momentum (see also Appendix 5). For an electron we defined the
spin angular momentum operator and its components and
with the same commutation properties as the orbital angular
momentum operator and its components , , and , that is,

(8.3)

In analogy to Eq. (6.43)

(6.43)

for the total angular momentum Schrödinger equation, one writes


the eigenfunctions of the operator as

(8.4)

Here, α and β are the spin wavefunctions (see below), and the
eigenvalues are
(8.5)
or

(8.6)

For electrons, the spin quantum number only can assume the value
of

(8.7)

Thus, the eigenvalues of the are operator for an electron are

(8.8)

For the operator, the eigenfunctions are

(8.9)

with eigenvalues

(8.10)

since the quantum number m can have values from −s to +s, in


analogy to Eq. (8.2). The corresponding electron spin eigenfunctions
are referred to as β and α, respectively. Because of the commutation
rules between the operator and the operator, the
eigenfunctions described by Eq. (8.9) are eigenfunctions of the
operator as well. Often, the “spin” is described in terms of the
electron spinning around an axis, similar to the earth spinning
around its axis. However, this view is a coarse simplification and
does not imply that the electron or nucleus really “spins.” The spin is
an inherent quantity of an electron that implies that it has an angular
momentum.
However, the idea of the electron actually “spinning” produces the
correct consequence, namely, that the electron (or a nucleus) has a
magnetic moment, that is, it may behave like a small permanent
magnet. In the absence of a magnetic field, the energy states of the
spin functions given in Eq. (8.9) are indistinguishable (degenerate).
However, if an electron is placed in a magnetic field – either an
externally applied magnetic field or the magnetic field of another
electron or nucleus – the energy states of the spin momentum split
into “parallel” and “antiparallel” spin states. When exposed to
electromagnetic radiation of the appropriate energy, transitions
between these spin states may occur. This gives rise to a
spectroscopic technique referred to as “electron paramagnetic
resonance,” also sometimes known as “electron spin resonance”
(ESR). EPR signals can only be observed for molecular or atomic
systems that have unpaired electronic spin, that is, for “radical”
atomic or molecular systems. Although EPR is a mature and very
useful technique, many of its principles are transferable from proton
magnetic resonance spectroscopy, and it will not be covered here in
favor of “nuclear” magnetic resonance spectroscopy, which is one of
the most widely applied spectroscopic methods in chemical research.

8.3 Nuclear Spin


What was elaborated upon in Eqs. (8.3)–(8.10) for electrons holds
equally well for any nucleus that has a nuclear spin quantum number
, such as 1H and 13C. The concept of nuclear spins arises from
the fact that nucleons themselves have an – albeit much smaller –
intrinsic spin moment. In atomic nuclei that have even numbers of
protons and neutrons, the spin moment is zero. Examples for such
nuclei are 12C and 16O. If the number of protons and neutrons in a
nucleus are both odd, the spin quantum number can be either of the
following:

(8.11)
For example, 14N has a nuclear spin quantum number of 1, whereas
for 10B, this value is 3.
Finally, if the sum of protons and neutrons is odd, such as in 1H, 13C,
and 19F, the spin quantum number can be any odd multiple of :

(8.12)

1H, 13C, and 19F all have nuclear spin quantum number of , whereas
17O has a spin of

All nuclei with a spin quantum number s > 0 have a magnetic


moment or create a magnetic moment along an axis. The magnetic
moment is proportional to the total nuclear angular momentum S
(see Eq. [8.6]) and is given by

(8.13)

In Eq. (8.11), e is the electronic charge unit, mP is the mass of the


proton (see Appendix 1), and gN is the unitless “nuclear g‐factor” (see
below). The expression (see Appendix 1) also is referred to as the
“nuclear magneton” and has the value

(8.14)

(see Example 8.2 for units and unit conversions). Here, the magnetic
field strength is given in units of T (tesla). Thus, Eq. (8.13) can be
rewritten as

(8.15)
In Eq. (8.15), γ is the magnetogyric ratio defined by

(8.16)
The magnetogyric ratio is the ratio of a nucleus' magnetic moment to
its angular momentum, cf. Eq. (8.16). Its units (and its name)
suggest that it is an indication of how fast a nucleus spins (in rad/s)
for a given magnetic field.
Equation (8.15) shows that the nuclear g‐factor is just a nucleus‐
specific proportionality constant between the observed magnetic
moment and the nuclear angular momentum S. The value of gN for
the proton is 5.5854. Values of other nuclei (with spin of ½)
commonly used in NMR spectroscopy are listed in Table 8.1, along
with the magnetogyric ratio and magnetic moments, expressed in
units of nuclear magnetons. The quantities in this table are related to
each other by Eqs. (8.15) and (8.16).
When a magnetic moment μ interacts with a static magnetic field B0
(also known as the magnetic flux density) along the z‐direction, the
energy of interaction is given by

(8.17)

Table 8.1 Nuclear g‐factors, magnetogyric ratios, and spin


moments for some spin ½ nuclei.
gN γ [rad/(T s)] μ [units of βN]
1H 5.5854 26.752 × 107 2.7927
13C 1.4048 6.728 × 107 0.7024
19F 5.2537 25.162 × 107 2.6289
Combining Eqs. (8.15) and (8.17) yields

(8.18)

For a proton the value for s is given by (see Eq. [8.10]). Thus,
the energy difference between the α and β states of a proton in a
magnetic field along the z‐axis is given by
(8.19)

Example 8.1 Calculation of the proton resonant frequencies and


energy differences for the field strength achievable a) with a room
temperature electromagnet (B0 = 2.34 T) and b) a very high field
superconducting magnet (B0 = 21.1 T)
Answer:

(E8.1.1)

(E8.1.2)

(E8.1.3)

(E8.1.4)

(E8.1.5)
Case a) corresponds to a “100 MHz NMR” spectrometer that can be
operated with a room temperature electromagnet. Case b)
corresponds to a high field, commercially available “900 MHz”
machine that require a liquid helium cooled, superconducting
magnet

Example 8.2 Analysis of the units of the magnetic moment:

(8.13)

where and gN and S are unitless.


Thus, the magnetic moment μ has units of

(E8.2.1)

The unit conversion in Eq. (E8.2.1) results from the


Lorentz force law that states that a particle that carries a charge of 1
coulomb and passes through a magnetic field of 1 T at a speed of 1m
per second experiences a force of 1 N.

Equation (8.18) leads to a very plausible explanation of NMR


spectroscopy. The two nuclear spin states α or β of a proton in a
magnetic field lead to two energy states that we can visualize as the
two spins magnetic moments being either parallel or antiparallel to
the external field (more on this spin alignment below). If
electromagnetic radiation with energy equal to the energy difference
between the two spin states impinges on the sample, the spin
function changes from α to β, photons of this frequency are
absorbed, and an NMR signal is generated. This is shown
schematically in Figure 8.2. This view treats the NMR effect strictly
as an electric dipole‐mediated transition, and, therefore, electric
dipole selection rules apply (see below). However, we shall see that a
different interpretation for NMR is possible (Section 8.4.3). For a
more detailed quantum mechanical representation of nuclear spin,
the reader is referred to the literature [1].
Figure 8.2 Energy of the α and β proton nuclear spin states as a
function of the external magnetic field. The vertical lines indicate the
transition frequencies (100, 400, and 900 MHz) for common
commercial magnets.
8.4 Selection Rules, Transition Energies,
Magnetization, and Spin State Population
8.4.1 Electric Dipole Selection Rules for a One‐Spin
Nuclear System
In keeping with the approach followed so far in this book, the next
step in discussing allowed interactions between nuclear spin systems
and electromagnetic radiation is to consider the selection rules for a
single proton spin to undergo a transition between spin states.
Interestingly, in many descriptions of NMR spectroscopy, this aspect
is not discussed, and it is tacitly assumed that these transitions are
always electric dipole allowed.
As we know from the discussion of molecular rotation (see Chapter
6), the selection rules for the total angular momentum operator
and its component along the z‐axis, , respectively, are

(6.53)
and

using the nomenclature of atomic systems, rather than that of


molecular rotations.
The corresponding operators for nuclear spins are and , with a
spin quantum number of

(8.7)

The two spin states α or β can have values between –s and +s or


With Δs = ± 1 for a one‐spin system, the transition
between the two spin states α or β is allowed. For spin systems
consisting of more than one spin, different selection rules will hold
[2].

8.4.2 Transition Energies


From Eq. (8.19), it follows that the allowed transition between states
α and β requires a photon of frequency

(8.20)

This equation is novel with respect to the fact that the molecular
energy levels are no longer molecule‐specific quantities (such as a
rotational, vibrational, or electronic energy levels discussed in
previous chapters) but depend on the strength of the external
magnetic field. Therefore, NMR can, in principle, be carried out at a
fixed magnetic field by scanning the photon energies or at fixed
photon energies by scanning the magnetic field until the resonance
condition is reached. Before the advent of pulse FT NMR
spectroscopy, nearly all NMR experiments were performed at a fixed
photon frequency by fine‐tuning the magnetic field.
The reasons for that will become clear from the following
consideration. In NMR spectroscopy, one of the desired prime pieces
of experimental information is a quantity known as the “chemical
shift” (see Section 8.5) that expresses how the electron density
around a nucleus shields the nuclear spin from the external magnetic
field. This shielding effect is quite small; consequently, the
absorption frequencies of differently shielded protons in a molecule
may differ only by a few parts per million. For an NMR experiment
carried out at a magnetic field of 2.34 T (see Example 8.1), the
resonance frequencies for a typical organic molecule may differ by
about 1000 Hz due to the different shielding the proton nuclei
experience. Technically, it is very difficult to build a radio frequency
(RF) generator that can be tuned from 100 000 000 to 100 001 000
Hz (for a 100 MHz NMR spectrometer) with sufficiently low
bandwidth and high‐frequency accuracy. Thus, NMR spectrometers,
in the past, used a fixed RF signal and varied the magnetic field
slightly to achieve the resonance conditions for differently shielded
nuclear spins.
With pulse FT NMR spectrometers, the observation of the signal
proceeds quite differently as will be discussed in Section 8.7.

8.4.3 Magnetization
Similar to the electric field of electromagnetic radiation exerting a
force on electrically charged particles (see Eq. [3.3]), the magnetic
field will exert a torque on the individual spin magnetic moments.
The effect of this torque will be an alignment of the magnetic
moments to assume a minimum energy configuration. The resulting
alignment of the magnetic moments will cause a macroscopic
magnetization. This magnetization can be viewed from the principles
of nuclear spins introduced earlier. Since only one Cartesian
component of the spin angular momentum, , and the square of
the total spin angular momentum, can be determined
simultaneously, the direction of the eigenvalues of the in the
external magnetic field cannot be determined. Therefore, a situation
similar to the spatial quantization discussed in Section 7.6 will occur,
namely, that the total spin angular moment lies on a cone about the
z‐axis and “precesses” around the central axis of the cone, as shown
in Figure 8.1b and c, at a frequency known as the “Larmor”
frequency. The sum of the magnetic moments of all individual
nuclear spins is referred to as the magnetization. When
electromagnetic radiation (in form of a broadband radio‐frequency
pulse) is applied to a system of aligned spins, the direction of this
magnetization changes, as will be discussed in Section 8.7. After the
perturbation due to the electromagnetic field ends, the spins will
relax into the original state, and energy will be released in this
process. This “free induction decay (FID)” is the signal picked up in
pulse FT NMR spectroscopy.
It is interesting to note that NMR spectroscopy can be described in
two different ways, namely, at a molecular levels where
electromagnetic radiation provides the energy to cause a transition
between different energy states of individual spins or, in terms of a
macroscopic quantity, the net magnetization. In this respect, NMR
spectroscopy is similar to some of the nonlinear optical effects
discussed in Appendix 3. The hyper Raman effect, for example, can
be described either by the hyperpolarizability (on the microscopic or
molecular level) or the second‐order dielectric susceptibility (on the
“bulk” level). The corresponding macroscopic magnetization will be
used later in more detail in the discussion of pulse FT NMR
spectroscopy (Section 8.7.2).

8.4.4 Spin State Population Analysis


As we have seen in Example 8.1, the energy difference between the α
and β spin states is much smaller than the energy differences
between vibrational or rotational energy states discussed before and
only amounts to about 10−25 J, or about 0.01 cm−1. In Example 8.3,
the population ratios for energy differences in two‐state systems with
energy differences of 1000, 1, and 0.01 cm−1 is calculated at room
temperature. The results represent population ratios typical for
vibrational, rotational, and NMR spectroscopies (see also Example
E3.1).

Example 8.3 Calculation of population differences for typical


vibrational (ΔE = 1000 cm−1), rotational (ΔE = 1 cm−1), and nuclear
spin energy states (ΔE = 0.01 cm−1) at room temperature and at
−100 °C. Use R = 0.698 [cm−1 K−1 molecule−1]:
Answer:
For the vibrational energy difference, the population of the excited
state is only 0.82 % or 0.025 % at room temperature or −100 °C,
respectively; thus, there is no danger of ever saturating the excited
state, particularly, since the lifetime of the excited state is very short
(see Section 5.3.2). In rotational spectroscopy, the population of the
two states is much closer to unity, particularly at room temperature.
In NMR spectroscopy, the two spin states have nearly the same
population, leading to the possibility of saturation (equalization) of
the two energy states. This is particular so since the spin states have
a much longer deactivation, or relaxation, times (0.1 s and higher)
than the other states discussed above. Working with magnetic fields
as high as possible and at temperatures below room temperature
reduces the saturation effect.

8.5 Chemical Shift


In Eqs. (8.19) and (8.20) and Example 8.1, the energy states of an
isolated proton spin in a magnetic field B0 were defined as

(8.19)

However, in chemistry and biochemistry, isolated protons are rather


uncommon, and hydrogen atoms either are attached to a molecule by
a covalent bond or, even in the case of a H+ cation in an aqueous
environment, are surrounded by solvent molecules and, therefore,
electrons. In response to an external magnetic field, the electrons
surrounding all nuclei create a secondary magnetic field that opposes
the external magnetic field, which is known as the diamagnetic
response. This induced magnetic field, Bind, is proportional to the
external magnetic field

(8.21)
where σ is a shielding constant. Thus, the local field at a nucleus
under the shielding effect of the electronic distribution is given by

(8.22)
Accordingly, Eq. (8.20) needs to be modified:

(8.23)

The frequency shift Δν exhibited by a nuclear spin then can be


expressed as

(8.24)

The resulting frequency shift is generally reported as a “chemical


shift” δ:

(8.25)

where νref is a reference signal due to an internal standard added to


the sample (see below).
Chemical shifts depend on the local electron density and present an
exquisitely sensitive probe of the chemical environment (i.e. the
electronic distribution) around a nucleus. Groups with high
electronegativity adjacent to a given proton will reduce the shielding,
thereby increasing the chemical shift. In contrast, electron‐donating
groups will increase the electron density around a nucleus, thereby
decreasing the chemical shift. In proton NMR spectroscopy, the
internal standard used most often is tetramethylsilane, Si(CH3)4,
abbreviated as TMS, and chemical shifts are generally reported in
ppm downfield from the TMS proton signal.
The protons in methyl groups in organic compounds have chemical
shifts very similar to that of TMS and, therefore, appear between 0.5
and 1 ppm from TMS. Protons adjacent to carbon–carbon double
and triple bonds show higher chemical shifts, and protons adjacent
to ketones exhibit chemical shift around 10 ppm.
Aside from the different signals created by the shielding/deshielding
effect of adjacent groups, another effect provides even more
information on the local surrounding of a probe nucleus. This effect
is due to the local magnetic fields created by the nuclear spin of
adjacent nuclei. This effect will be discussed next.

8.6 Multispin Systems


8.6.1 Noninteracting Spins
Let us consider two individual nuclei 1 and 2 in a molecule that
experience a different chemical environment and, therefore, have
different chemical shifts. Such nuclei are said to be nonequivalent.
The magnetic field perturbation due to other nuclei shows up in the
NMR spectrum when the nuclei are nonequivalent and if the
distance between nonequivalent nuclei is less than or equal to three
bond lengths.

From Eqs. (8.19), and (Eq. [8.22]), Blocal = (1


− σ) B0, we can write the Hamiltonian for a system of two
distinguishable, noninteracting spins, labeled 1 and 2, as follows [3]:

(8.26)

The eigenfunctions for the two‐spin system are the products of the
eigenfunctions of the individual spin operators and . There
are four possibilities to write these product functions:
(8.27)
With these eigenfunctions, the following energy eigenvalues are
obtained (see Example 8.4):

(8.28)

Example 8.4 Find eigenvalues of the operator


for the wavefunctions
ψ1 = α(1)α(2) and ψ2 = β(1)α(2):
Answer:

a. For ψ1 one obtains

(E8.4.1)

For the two‐spin wavefunction ψ1 = α(1)α(2) (see Eq. [8.9]),

and
. Thus,
(E8.4.2)

b. For ψ2 one obtains

(E8.4.3)

For the two‐spin wavefunction ψ2 = β(1)α(2) (see Eq. [8.9]),

and
. Thus,

(E8.4.4)
Figure 8.3 (a) Energy level diagram for two noninteracting spins
with shielding constants σ1 and σ2. Energies are given in units of
ħ γ B0 (cf. Eq. [8.28]). (b) Simulated NMR spectrum for two
noninteracting spins.
These energy levels are depicted in Figure 8.3a. The selection rule
(see Section 8.4.1) Δs = ± 1 implies that only one of the two spins can
change in a transition. Thus, four transitions can occur, with the
following frequencies:

(8.29)

and

(8.30)

These transitions are indicated by the gray up arrows in Figure 8.3a.


Since two each of the transition frequencies are the same, only two
spectral peaks are observed, as shown in Figure 8.3b. These peaks
are split by (σ1 − σ2).

8.6.2 Interacting Spins: Spin–Spin Coupling


For two noninteracting spins, as we have seen in the previous
section, the NMR spectrum would show two peaks at (1 − σ1) and (1
− σ2), in accordance with Eq. (8.24). This rather unexciting result
would seriously undermine the importance of NMR spectroscopy, if
the two spins would just produce two signals at their individual
chemical shift. Fortunately, for the structural sensitivity of NMR
spectroscopy, spins in close proximity to each other interact via
spin–spin coupling. There are two mechanisms for such coupling to
occur: “through‐space” and “through‐bond” (also referred to as
scalar) coupling. Of those, the former is influencing solid‐state NMR,
whereas the latter mechanism is of major importance for the
assignment of peaks in NMR studies in liquids and for structural
analysis. This effect is also called J‐coupling and is described as
follows.
Since each spin acts like a magnet, it will affect neighboring spins by
an interaction such that the Hamiltonian given in Eq. (8.26) needs to
be modified to take into account this interaction:

(8.31)

In Eq. (8.31), J12 is known as the coupling constant that indicates the
strength of interaction between the spins. This interaction acts over
relatively short distances and drops off rapidly with distance. Thus,
spin–spin coupling is observed mostly between spins separated by
less than 3 or 4 bonds.
This perturbed Hamiltonian is solved, using the perturbation
methodology described in Appendix 2, using the eigenfunctions of
the unperturbed Hamiltonian (cf. Eqs. [8.26]–[8.28]) to compute
the perturbation energies ΔE (to the unperturbed wavefunctions) as
follows:

(8.32)
where m1 and m2 have the values for the spin states α and
β, respectively. The transition frequencies listed above for the
noninteracting spins

(8.29)

and

(8.30)

are modified to give

(8.33)

This results in a quartet of peaks as shown in Figure 8.4 for two cases
of external magnetic fields. As can be seen from Figure 8.4 and Eq.
(8.33), the splitting between the nuclei with different chemical shift
depends on the external magnetic field, whereas the splitting to the
spin–spin interaction depends only on J12.
Spin–spin coupling produces distinct spectral patterns in NMR
spectra if more than two spins interact. This splitting pattern aids in
the assignment of peaks and is taught in course in organic chemistry
and spectral identifications. These splitting patterns can be
summarized as follows.

8.6.3 Interaction of Multiple Spins


For systems containing multiple interacting spins, a generalized form
of Eqs. (8.30) and (8.32) can be written for protons:

(8.34)

Figure 8.4 Spectral pattern observed for two interacting spins at


lower (a) and higher (b) external magnetic field. The centers of the
doublets in each case are split by (σ1 − σ2).Notice that the J‐coupling
does not depend on the external magnetic field.
In this equation, σA is the shift for nucleus A; mA and mX, as before,
are the spin quantum numbers that have the values for
the spin states α and β; and JAX is the spin–spin coupling constant
between spins A and X, where X denotes the number of additional
spins that interact with spin A. Equation (8.3) collapses to Eq. (8.33)
in the case of two spins, and the splitting pattern of four peaks with a
1:1:1:1 intensity ratio is obtained (Figure 8.4).
For the case of two identical spins X interacting with spin A (an AXX
system), the first spin interacting with A creates two states with a
splitting JAX. Further interaction of the two resulting states creates
four states, two of which are degenerate. This leads to a triplet of
peaks, all spaced at JAX, with a 1:2:1 intensity pattern. This is shown
in Figure 8.5a. By the same token, we can predict that an AXXX spin
system will create a quartet of peaks, all spaced at JAX, with an
intensity pattern of 1:3:3:1, as shown in Figure 8.5b. This explains,
for example, the observed spectrum of an ethyl (–CH2CH3) group.
Here, the methyl protons exhibit a chemical shift of about 1 ppm
from TMS, and the methylene protons a higher shift, between 2 and
4 ppm, depending on the exact chemical environment. The two
methylene protons split the three equivalent methyl protons into a
triplet with 1:2:1 intensity pattern according to the AXX coupling
scheme, whereas the three methyl protons split the two equivalent
methylene protons into a quartet with 1:3:3:1 intensity pattern,
according to the AXXX coupling scheme [2].

Figure 8.5 Spin–spin coupling patterns for (a) JAXX and (b) JAXXX
spin systems. See text for details.
8.7 Pulse FT NMR Spectroscopy
8.7.1 General Comments
In this text book, experimental approaches generally are neglected in
favor of the theory of the spectroscopic method. However, here in the
case of NMR spectroscopy, a few experimental details will be
presented at a rather introductory level in order to point out how the
experimental method development has spawned entirely new
methods in NMR spectroscopy that were either impossible or
impractical to carry out before the methodology known as “pulse FT
NMR spectroscopy” was developed.
As indicated toward the end of Section 8.4.2, most modern NMR
experiments are being carried out in a manner quite differently from
the “conventional NMR” methods that were prevalent before the
1980s. In these conventional methods, the sample was illuminating
by monochromatic electromagnetic radiation in the RF range
(typically 100 MHz) while the magnetic field was varied by small
amounts to achieve the resonance condition. Conversely, one could
have kept the magnetic field constant and scanned the radio
frequencies over an appropriate range, but the former method
proved to be more feasible. The disadvantage of the “conventional
NMR” described above is that only one spectral element is measured
at a given time, where a spectral element can be visualized at a small
band of chemical shift frequencies. In order to collect a spectrum
over a band of 10 ppm from the TMS signal at a spectral resolution of
0.1 ppm, more than 100 spectral elements would need to be
collected. However, in FT methods, all the spectral elements are
collected at the same time via a relaxation phenomenon known as
the free induction decay (see below). The FID represents the same
signal as observed in conventional NMR experiments but is collected
in a different data domain. By performing an operation known as a
Fourier transform on the FID, the same spectrum is obtained as in
the conventional NMR experiment. However, the FT methodology
affords enormous time savings in spectral data acquisition or in a
substantial increase of the data's signal‐to‐noise (S/N) ratio.
In this respect, pulse FT NMR spectroscopy is similar to FT‐IR
spectroscopy, where an interference pattern (an “interferogram”) of
all infrared frequency bands is collected simultaneously. This
interference pattern again represents a different data domain, and
the infrared spectrum is obtained by FT of the interferogram. Both
FT techniques require that a computer is interfaced to the
spectrometers to render the data in the form researchers are
accustomed to see. Details of the actual mathematical procedure
known as Fourier transform are presented in Appendix 4.

8.7.2 Description of NMR Event in Terms of the “Net


Magnetization”
In the presence of an external magnetic field, the nuclear spins
experience a torque and align such that the z component of the spin
aligns with the external field (remember that the x and y component
cannot be determined and lie on a cone as shown in Figure 8.1b and
c). Thus, there is no magnetization in these two directions, and all
magnetization is along the z‐direction. This magnetization is referred
to as MZ.
When such a system of spins is exposed to RF electromagnetic
radiation along the y‐direction of the appropriate energy, spins can
flip from the lower energy α state into the β state. In this process it is
possible to saturate the spin system, i.e. equalize the population of
the α and β states, and reduce the magnetization along the z‐
direction to zero (MZ = 0). Remember that the excess of the α state
over the β state is small to begin with, as was demonstrated in
Example 8.3. An RF field that achieves the net magnetization to
become zero is known as a 90° pulse.
After the perturbation ceases, the relaxation back into the lowest
energy state occurs with the magnetic vector leaving the x–y plane
and gyrating back into the z‐direction. This relaxation process
follows Eq. (8.35):
(8.35)

In Eq. (8.35), M0 is the equilibrium net magnetization, and T1 is


referred to as the spin–lattice relaxation. In this process, energy is
released since the system is relaxing back into its lowest energy state.
In order to describe this energy emission, one must consider that the
magnetic moment, after absorption of the RF pulse, still precesses in
the x–y plane. Thus, when the magnetization vector flips back into
the z‐direction, it does so by its tip describing a spiral pathway as
shown in Figure 8.6a.
This process of reorientation of the magnetization can be viewed in
two ways. If one views this situation in a framework of laboratory
coordinates, the magnetization will describe a spiral pathway.
However, if one describes this movement of the magnetization vector
in a coordinate system that spins about the z‐axis at the Larmor
frequency, the movement of the magnetization vector can be
described by a simple arc in the y–z plane from the x–y plane into
the direction of the z‐axis, as shown in Figure 8.6b.

A detector mounted along the x‐direction detects a periodic signal


due to the energy loss of the system, known as the free induction
decay. This periodic signal contains all the dephasing information of
all the spins of the system. Therefore, it is a superposition of all the
spin frequencies, as shown in Figure 8.7. Consequently, an FT needs
to be carried out to extract individual frequency components. The
principles of Fourier transformation and the fast FT will be
elaborated upon in Appendix 4.
Figure 8.6 Reorientation of magnetization vector following a 90°
pulse, viewed in a laboratory coordinate system (A) and in a
coordinate system rotating about the z‐axis.

Figure 8.7 (a) Simulated “free induction decay” (FID) and Fourier
transformed signal (b).

References
1 Cavanagh, J. (2007). Theoretical description of NMR
spectroscopy. In: Protein NMR Spectroscopy (eds. J. Cavanagh et
al.). New York: Academic Press.
2 Keeler, J. (2010). Understanding NMR Spectroscopy, 2e.
Chichester, UK: Wiley.
3 Engel, T. and Reid, P. (2010). Physical Chemistry, 2e. Upper
Saddle River, NJ: Pearson Prentice Hall.

Problems
1. Why is the J‐coupling in an NMR experiment independent of the
overall field strength?
2. In NMR spectroscopy, two different effects are observed:
chemical shift and spin–spin coupling.
a. Discuss the physical phenomena responsible for these
effects.
b. Which of these effects depends on the magnitude of the
applied magnetic field?
c. Why is it advantageous to perform NMR spectroscopy at
high magnetic fields and low temperature?
3. Why is the NMR signal reported with respect to an internal
standard and not in absolute photon energies as in most other
spectroscopic techniques?
4. In a 500 MHz proton NMR instrument, what is the shift, in Hz,
experienced by a proton whose signal is observed at 8.0 ppm
from TMS?
5. Repeat the calculations performed in Example 8.4, namely, to
find the eigenvalues of the two‐spin Hamiltonian:

for the remaining two


wavefunctions
ψ3 = α(1)β(2) ψ4 = β(1)β(2). NO PEEKING in Example 8.4!
6. Propionic acid has a pKa value of about 4.9. In a 0.1 M aqueous
solution, one mostly will find the propionate anion. Predict the
1H NMR spectrum of the propionate anion.
9
Atomic Structure: Multi‐electron Systems

9.1 The Two‐electron Hamiltonian,


Shielding, and Effective Nuclear Charge
In analogy to the hydrogen atom, the Schrödinger equation for the
first atom containing more than one electron, helium, is written as:

(9.1)

or

(9.2)

In Eq. (9.1), the first two terms in the bracket are the kinetic energy
expressions for electrons 1 and 2; the third and fourth terms are the
attractive forces between the helium nucleus with nuclear charge +2
and electrons 1 and 2, and the fifth term is the electron–electron
repulsion. This equation cannot be solved analytically because each
electron's wavefunction depends simultaneously on r1, r2, and r12.
Thus, one has to resort to an approximate method, known as the
orbital approximation, in which the total electronic wavefunction of a
multi‐electron atom is written as the product of independent one‐
electron wavefunctions. For helium, the orbital approximation can
be formulated as

(9.3)
That is, one assumes that the total wavefunction of multi‐electron
atom can be written just as the product of each electron's individual
wavefunction. This approach ignores the electron correlation, which
is basically an attempt of the electrons to stay out of each other's way
by coordinating their motion. Instead, one assumes that each
electron experiences the presence of other electrons just by an
“effective nuclear charge ζ"due to the time‐averaged position of all
other electrons. The effective nuclear charge the 2nd electron
experiences in the He atom is 1.688, indicating that the first electron
shields the nucleus to such a degree that the second electron does not
perceive a nuclear charge of 2.0, but 1.688 charge units. For the Li
atom, the effective nuclear charges are 2.691 for the 1s and 1.279 for
the 2s electrons [1].
Thus, the wavefunction of the 1s orbital, in the presence of another
electron, is no longer expressed as

(7.17)

but as

(9.4)

where ζ is the shielded nuclear potential (see below). Similarly, the


orbital energies for a multi‐electron atom are no longer given by Eq.
7.12

(7.12)

but by an equation that includes the shielded nuclear potential:

(9.5)
In order to describe a multi‐electron atom, the so‐called Hartree–
Fock self‐consistent method is used. In it, each electron is described
by the orbital approximation (Eq. (9.3)), in which the electron
correlation is ignored and approximated by an averaged position of
all other electrons expressed via the effective potential, Veff that is
calculated by averaging the inner electron wavefunctions over all
angular coordinates.
Each of the one‐electron wavefunctions in Eq. (9.3) has the form

(9.6)

where the effective nuclear charge experienced by each electron


depends on the attraction between the nucleus and electron, and the
shielding by the other electrons. Hartree–Fock calculations are
carried out by minimizing the energy of the multi‐electron system by
numerically varying ζ for the hydrogen‐like wavefunctions.
Furthermore, the spin properties of all electrons in the atom need to
be considered in a multi‐electron atom, according to the Pauli
principle. Thus, the total Hamiltonian is written in the form of an
antisymmetrized wavefunctions φj(rj), given by the Slater
determinant (see next Section).

9.2 The Pauli Principle


When writing the total wavefunction for multi‐electron atom, the
electron spin wavefunctions need to be included. For the He atom,
for example, the assumption was made that the second electron is in
the same 1s orbital as the first electron, but has an opposite spin
function, as compared to the first one.
This is achieved by rewriting the total electronic wavefunction (Eq.
(9.3) of the He atom by including the spin functions, as follows:
(9.7)
which often is abbreviated to

(9.8)
In the description so far, it appeared that we can put labels on the
electrons and call one of them “electron1” and the other “electron2.”
However, since electrons are indistinguishable, Eq. (9.8), as
presented above, is incorrect since it would imply that “electron2”
has the spin function β. Instead, we need to present Eq. (9.8) in such
a way that both electrons can have either spin functions. This is
accomplished by writing the total wavefunction as a superposition of
both possibilities

(9.9)
Equation (9.9) often is further abbreviated, for simplicity of notation,
as

(9.10)
The Pauli principle (see Postulate 7 in Chapter 2) states that the
antisymmetric combination

(9.11)
is the correct way to write the superposition of the two states, and
implies that the wavefunction changes sign when electrons 1 and 2
are exchanged. Equation (9.11) can be represented as a normalized
determinant known as the Slater determinant:

(9.12)

For the Li atom, for example, the Slater determinant assumes the
form
(9.13)

Here, we have written all possibilities to have two electrons in the 1s


and 1 electron in the 2s orbital, with all different permutations of the
spin functions. Since a determinant is zero if two rows or two
columns are equal, this formulation accounts for the Pauli principle:
the wavefunction will be zero (forbidden) if any two electrons have
identically the same set of four quantum numbers n, l, m, and ms.
This formalism is included in the Hartree–Fock method in that Eq.
(9.6) needs to be reformulated to include the normalized,
antisymmetric Slater determinant of the spin wavefunctions. This
method then produces the best one‐electron approximation for the
orbital energies in multi‐electron atoms and yields orbital energies
that are summarized in Figure 9.1

9.3 The Aufbau Principle


The energy eigenvalues of the orbitals resulting from the reduced
nuclear charges expressed in Eq. (9.6), combined with the results
from the discussion of the Pauli principle, lead to a revised energy
level diagram for multi‐electron atom, as compared to Figure 7.3.
This new energy level diagram is shown in Figure 9.1. It is
responsible for the shape of the periodic chart, in particular, the
existence of main group elements, the transition metals, and the
lanthanides and actinides.
Figure 9.1 Energy level diagram of multi‐electron atoms,
explaining the Aufbau principle, and the form of the periodic chart.
In the periodic chart, elements in a given row are listed in order of
increasing atomic number, Z. The number of electrons in a neutral
atom also equals Z. These electrons for each atom are filled in the
orbitals according to the orbital energy scheme shown in Figure 9.1,
starting with the lowest energy orbitals, and two electrons with
opposite spin quantum numbers to each orbital (according to the
Pauli principle discussed above). The occupation of electrons in
orbitals is abbreviated by the following example for the element
carbon: 1s2 2s2 2p2, indicating that there are two electrons in the 1s
orbital, two electrons in the 2s orbital, and 2 unpaired electrons in
two 2p orbitals (Hund's rule, see below).
In Figure 9.1, the energy of the 4s orbital is lower than that of the 3d
orbitals. Consequently, the order of filling the orbitals, as Z increases
from element to element, is 1s, 2s, 2p, 3s, 3p, 4s, 3d,…, and not 1s, 2s,
2p, 3s, 3p, 3d, 4s,…. This accounts for the form of the periodic chart
with the transition metal elements (Sc through Zn) inserted into the
fourth row between Ca and Ga. This is repeatesd for the 14
lanthanide elements that are inserted after the 6s2 5d1 configuration
is reached.
If orbitals are degenerate, such as the 2p orbital, electrons are filled
with parallel spins into l = 1 orbitals with different m‐values. Thus,
the C atom in its ground state has two, and the N atom has three
unpaired and parallel spins. This further rule within the Aufbau
principle is known as “Hund's rule” that states that degenerate
orbitals are filled in such a way as to maximize spin multiplicity or
maximum number of unpaired spins. This is due to the fact that
pairing the spins of two electrons requires a certain amount of
energy, known as the spin‐pairing energy. The number of unpaired
spins directly affects the magnetic properties of atoms (as well as of
molecules). Species with two unpaired, parallel spins, such as the C
atom in its lowest energy state, are paramagnetic. If its spins were
paired, the atom would be diamagnetic. Thus, one finds that the
shape of the periodic chart of elements, and many properties of the
atoms, are determined by the energy level diagram shown in
Figure 9.1 that can be obtained from Hartree–Fock calculations.
The spin‐pairing energy mentioned above is also responsible for a
few minor exceptions in the Aufbau order, for example, in the case of
the chromium atom. According to the energy level diagram shown in
Figure 9.1, one should expect the Cr atom to have a [Ar]4s23d4
electronic configuration, but the actual configuration is [Ar]4s13d5,
that is, an arrangement devoid of paired electrons in the outer shells.
(Here, [Ar] denotes filled energy shells up to the 18th electron: [Ar] ≡
1s2 2s2 2p6 3s2 3d6). This demonstrates that the energy difference
between the 4s and 3d orbitals is so small that it can be overridden
by the spin‐pairing energy.

9.4 Periodic Properties of Elements


Many experimentally and theoretically obtained periodic properties
of elements can be explained by the energy level diagram shown in
Figure 9.1 and are discussed in detail in many introductory texts on
general chemistry and introductory physical chemistry. Here, just
two of these properties will be discussed further to illustrate the
information available from this energy level diagram. One of these is
the ionization energy (also known as the ionization potential) that
expresses the energy required to ionize an element M to give a
positively charged cation M+ and an electron, e− :

(9.14)
All alkali metals, for example, have a single electron in the “s”
orbitals (Li: 1s2 2s1, Na: 1s2 2s2 2p6 3s1, etc.). Since these electrons
experience a nucleus that is highly shielded by the inner shell
electrons (see Eq. (9.6)), it is relatively easy to remove these
electrons, either by subjecting gaseous atoms to an external potential
(hence the term ionization potential) or to light (the previously
discussed photoelectric effect, see Section 1.3). When plotted against
the atomic number, a graph of the ionization energies, as shown in
Figure 9.2, is obtained. This figure demonstrates that the ionization
energy is low for all alkali metals and the alkali earth metals but
increases toward the right (with increasing atomic number) for each
row in the periodic table.
The graph even shows that for elements filling the “p” orbitals, there
is a dip in the ionization energy when a half‐filled shell is reached, in
agreement with the discussion of the spin‐pairing energy above. In
this plot, the transition metals were omitted, since for these
elements, inner electrons (for example, the 3d orbitals in the
elements from Sc to Zn) are filled that occupy space within the radius
of the 4s orbital and for which the ionization energy varies only a
little.
Figure 9.2 Ionization energies (a) and atomic radii (b) for main
group elements.
Similarly, atomic radii can be predicted from an analysis of
Figure 9.2. Since the nuclear charge increases for elements within a
row of the periodic chart, the attraction an electron experiences
(despite the shielding effect) increases, and the atomic radius
decreases within a row of elements in the periodic chart. However,
when the next element in a new row is reached, a dramatic increase
in the atomic radius is observed; see Figure 9.2.

9.5 Atomic Energy Levels


We used the hydrogen atom emission/absorption spectra as a vehicle
to introduce the stationary‐state energy levels in the simplest atom,
and the quantized energy packets required for transitions between
these states. As it turns out, atoms with more than one electron also
exhibit rich spectral features that have been used for centuries (think
Chinese fireworks!) due to the bright color emissions exhibited by
certain elements. Also, in modern analytical science, the detection
and quantification of many elements by atomic absorption
spectroscopy (down to part‐per‐billion concentrations) are
accomplished by the absorption of a specific wavelength of light by
atomic species.
Atomic spectroscopy will be briefly discussed here to introduce some
principles of electronic transitions in atoms that contain more than
one electron. In particular, the formalism of combining total angular
and spin angular momenta into a new quantity that determines the
selection rules for atomic electronic transitions will be introduced,
since this formalism also allows the classification of electronic
transitions in small (diatomic) molecules.

9.5.1 Good and Bad Quantum Numbers and Term


Symbols
For the hydrogen atom, all operators required to solve the
Schrödinger equation, i.e., , , and , commutated with the
total Hamiltonian ; thus, the eigenvalues l, ml, and s, as well as the
total energy, can be determined simultaneously. Since the
Schrödinger equation for a multi‐electron atom, given by Eq. (9.1):

Figure 9.3 Vector addition schemes for (a) the total orbital
angular momenta and (b) the total spin angular momenta.

(9.1)
cannot be solved explicitly (even for the case of two electrons), one
cannot obtain good quantum numbers any longer. Good quantum
numbers are defined as those whose operators commutate with the
total Hamiltonian. Although the eigenvalues of the individual orbital
angular momenta do not commutate with the total Hamiltonian,
their sum (i.e. the total orbital angular momentum of all electrons)
does commutate with the total Hamiltonian. This total orbital
angular momentum is obtained by the vector sum of the orbital
angular momenta li of all i electrons:

(9.15)

In Eq. (9.15), “s” electrons never contribute to the vector sum, since
their angular momentum is zero due to the spherical nature of “s”
orbitals. Furthermore, paired electrons in the same orbital do not
contribute either since their spin moments add to zero.
Allowed values (for a two‐electron system) of the vector addition in
Eq. (9.15) are

(9.16)
Equation (9.16) is also referred to as Clebsch–Gordan expansion that
will be used here without proof. For example, in an excited atom or
ion that has one electron in a “p” orbital (l = 1) and another in a “d”
orbital (l = 2), the vector addition leads to three different total
angular momenta, namely 3, 2, or 1 (see Figure 9.3a). In accordance
with the nomenclature of individual electronic orbital angular
momenta, the sum of the angular momenta is denoted by the capital
letters S, P, D, F… as shown in Table 9.1:
Table 9.1 Symbols of states for different l and L values.
Similarly, one defines a total spin angular momentum S as

(9.17)

where S commutates with the Hamiltonian. Again, for the case of


two spins with s = ½, there are two possibilities, S = 1 if the spins are
parallel, or S = 0 if they are antiparallel. The vector addition schemes
described by Eqs. (9.16) and (9.17) are depicted for a two‐electron
system in Figure 9.3b.
The spin multiplicity is defined as 2S + 1. Thus, for S = 0, the
multiplicity is one (a “singlet” state) and for S = 1, the multiplicity is
3 (a “triplet” state). The multiplicity is written as a left superscript of
the total orbital angular momentum, e.g. 1D or 3F. These symbols are
called “term symbols” and specify the spin states of a multi‐electron
system (cf. Example 9.1).
Finally, the total orbital angular momentum and the total spin
momenta can interact via spin‐orbit coupling to produce a total
angular momentum that is specified by

(9.18)
This spin‐orbit quantum number J is not to be confused with the
rotational quantum number discussed in Chapter 6.
The allowed values of J are again given by the Clebsch–Gordan
expansion rule

(9.19)

This total spin‐orbit angular momentum quantum number J is


written as a right subscript to the term symbol, such as 3P0. Such an
expression is known as a “level.”

Example 9.1 What are the term and level symbols of the ground
and excited states of the Na atom referred to in Section 7.5 and
Figure 7.6b?
Answer:
The ground‐state Na atom has a 1s2 2p6 3s1 electron configuration.
The filled shells do not contribute to the term symbol. The lone
electron in the 3s orbital has an orbital angular momentum of zero
(since it is in an s orbital) and a spin angular momentum of .
The spin multiplicity is 2; therefore, the term is 2S. The total angular
quantum number Thus, the ground state can be
described as 2S1/2.

The excited state is a 1s2 2p6 3p1 configuration. The electron in the p‐
orbital has L = 1, and the term is 2P. The total angular quantum
number is or ; thus, the excited states can
be described as 2P3/2 or 2P1/2. These two states differ slightly in
energy; thus, the transition (the 589 nm sodium line) is split into a
doublet.

Atomic spectra are very rich in spectral lines since the atomic species
exhibit a large number of excited states due to the existence of many
unoccupied atomic orbitals into which electrons can be promoted.
Transitions into or from these states are either dipole allowed (for
the discussion of selection rules, see next section) or are caused by
radiationless processes, such as collisions with other highly excited
species. The former case of radiation‐induced transitions was
discussed before for the hydrogen atom emission and absorption
spectra (see Section 7.3). The latter case was mentioned for the
excitation process in the He–Ne laser where the Ne atom excited
states are populated via collisions with electrons and/or He ions (cf.
Section 3.4). The number of excited states is further increased by the
interaction of angular and spin momenta, as indicated by Eqs.
(9.15)–(9.19). Just how many states can be produced from one
electronic configuration is shown in Example 9.2. Population or
depopulation of some of these states may not be allowed by
radiation‐induced transitions; nevertheless, these states are real and
can contribute to the observed atomic spectra.
Example 9.2 What are the possible term and levels of an excited C
atom with an electronic configuration of 1s2 2s2 2p1 3d1?
Answer:
As we have seen in Example 9.1, any electrons in filled shells do not
contribute to the total orbital angular momentum. The electron in
the p orbital has an angular momentum l = 1, and the electron in the
d orbital has l = 2. According to Eq. (9.16), the allowed L values are:
L = 3, 2 or 1, corresponding to D, P and S symbols.
The total spin angular momentum S can be 1 or 0; thus, the spin
multiplicity will be 3 or 1. Consequently, the possible term symbols
are 3D, 3P, 3S, 1D, 1P, 1S.
According to Eq. (9.19),

(9.19)
the total angular momentum J can have values from 4, 3, 2, 1; thus,
each of the term symbols can have the subscripts from 4 to 1, e.g.,
3D , 3D , 3D and 3D
4 3 2 1

9.5.2 Selection Rules for Transitions in Atomic Species


The selection rule for atomic species, in general, follows closely the
selection rule for the H‐atom in that it is determined by the condition
Δl = ± 1. With the coupling of individual angular momenta into a
total angular momentum

(9.15)

and the spin‐orbit coupling

(9.18)
Figure 9.4 Simplified energy level diagram of the Li atom and
transitions indicated in Table 9.1.
additional selection rule arises, given by

(9.20)
and

(9.21)
There is an additional condition, that the total spin angular
momentum cannot change, that is,

(9.22)
This latter selection rule prohibits, for example, singlet to triplet
transitions, and will be particularly important in electronic
spectroscopy of di‐ and polyatomic molecules.

9.6 Atomic Spectroscopy


The large number of atomic energy levels, discussed in Section 9.5.1,
makes for rich atomic spectra in both absorption and emission. We
shall investigate in some detail the atomic spectrum of one of the
simplest atoms, Lithium, which has a 1s22s1 electronic configuration.
A tabulation of the transitions of neutral Li can be found in the
literature [2] and lists 21 transitions between 200 and 2700 nm
wavelengths. The most relevant of these are listed in Table 9.2 and
shown in Figure 9.4.
Similar to the situation in sodium, the major transition, marked “C”
in Table 2, is between the ground state and an excited state in which
the 2s electron transitions into the 2p orbital. This configuration
results in two levels, 2P3/2 and 2P1/2, that differ in energy by about a
third of a wavenumber. Consequently, all transitions involving these
two energy levels are split into doublets. The splitting is only about
0.02 nm for the transition marked “C.” We saw before (see Example
9.1) that the same two levels involved, namely 2P3/2 and 2P1/2 were
responsible for the splitting observed in the sodium 589 nm line.
Incidentally, the bright red emission at 670.8 nm gives Li its deep
red color in firework displays. It is also easily visible when Li
compounds are aspirated into a flame (see below).
Table 9.2 Transition, energies, term symbols, and wavelengths of
the prominent Li atomic lines.
Source: Adapted from National Institute of Standards and Technology [2], Transition,
energies, term symbols and wavelengths of the prominent Li atomic lines.

Transition Term Energy [cm−1] Wave‐ Relative


excited ‐ ground length intensity
[nm]
1s24d1 ← 2D
3/2 36 623.30 – 14 460.28 15 A
1s22p1 2 903.62
← P1/2

1s24d1 ← 2D
5/2 36 623.31 – 14 460.29 30 A
1s22p1 ← 2P1/2 903.62

1s23d1 ← 2D
3/2 31 283.02 – 14 610.35 300 B
1s22p1 ← 2P1/2 903.62

1s23d1 ← 2D
5/2 31 283.05 – 14 610.36 400 B
1s22p1 ← 2P1/2 903.62

1s22p1 ← 2P
3/2 14 903.96 ‐ 0.00 670.77 500 C
1s22s1 ← 2S1/2
1s22p1 ← 2P
1/2 14 903.62 ‐ 0.00 670.79 1000 C
1s22s1 ← 2S1/2
1s23s1 ← 2S
1/2 27 206.07 – 14 812.62 150 D
1s22p1 ← 2P1/2 903.62

1s23s1 ← 2S
1/2 27 206.07 – 14 812.64 300 D
1s22p1 ← 2P3/2 903.96

1s23p1 ← 2P
3/2 30 925.61 – 27 2687.76 10 E
1s23s1 ← 2S1/2 206.07

1s23p1 ← 2P
1/2 30 925.51 – 27 2687.78 5 E
1s23s1 ← 2S1/2 206.07
A Li atom energy level diagram is shown in Figure 9.4 with the
transitions listed in Table 9.1 indicated by the capital letters. In this
energy level diagram, the splitting by the spin‐orbit coupling is not
indicated since it contributes only fractions of a wavenumber.

9.7 Atomic Spectroscopy in Analytical


Chemistry
The abundance of atomic transitions in metal makes atomic
spectroscopy, specifically, atomic absorption spectroscopy, one of the
most widely and cost‐effective methods to analyze for the presence of
metal ions in analytes down to the ppm‐level. In atomic absorption
spectroscopy, the analyte is aspirated into the flame of a burner,
typically a very hot flame from an acetylene–oxygen mix. Under
these conditions, all compounds are broken down into atoms and
ions that – due to the high temperature in the flame, exhibit
absorption spectra with broadened lines. The flame is illuminated by
light from a special source, known as a hollow cathode lamp (HLC).
This lamp is specific for the element to be analyzed: to determine the
arsenic concentration in a sample, an arsenic hollow cathode lamp
must be used. In an HCL, the analytes emission spectrum (in this
case, the arsenic emission spectrum) is produced at low temperature
by sputtering atoms off the As cathode in a carrier gas. These atoms
collide with electrons in the electric discharge, get excited and emit
narrow atomic emission lines. The absorption of the HCL radiation
by the atoms/ions in the flame is observed and converted to an
analyte concentration via the Beer–Lambert law (see Eq. [3.25]).
The narrow lines emitted by the HCL makes this method particularly
useful to eliminate interference from other elements. Furthermore,
the sheer number of strong and weak transitions in an atomic
emission allow the user to select lines that are most suitable for the
concentration range to be analyzed.

References
1 Clementi, E. and Raimondi, D.L. (1963). Atomic screening
constants from SCF functions. The Journal of Chemical Physics
38 (11): 2686–2689.
2 National Institute of Standards and Technology. Basic Atomic
Spectroscopic Data.
https://physics.nist.gov/PhysRefData/Handbook/Tables/lithiumt
able2.htm.

Problems
1. Write electronic configurations for Sc, Sc+2, and Sc+3
2. State Pauli's principle in at least two different ways.
3. a. State Hunds' rule, and give a physical explanation for it.
b. What are the implications of Hund's rule given the
magnetism of transition metals and rare earth metals?
4. Discuss the trend in the 1st ionization energies of the alkali
metals
5. Discuss the trend in atomic radii of the alkali metals
6. Why is the 3rd ionization energy of Ca much larger than the 1st
and 2nd?
7. Explain why shielding is more effective by electrons in a shell of
lower principal quantum number than by electrons having the
same principal quantum number (i.e., why does a 2s electron
shield a 3p electron more effectively than a 3s electron would?)
8. Write the normalized Slater determinant for the ground‐state
configuration of Be.
9. In analogy to the Li and Na emission spectra, discuss the
observation of a doublet at 393.37 and 396.85 nm in the
emission spectrum of potassium.
10
Electronic States and Spectroscopy of
Polyatomic Molecules
In electronic spectroscopy, electrons are promoted into more highly
excited electronic or vibronic (a concatenation of vibrational and
electronic; see Section 10.4) states that can either be centered at
metal atoms in coordination compounds or unoccupied molecular
orbitals (MOs) such as antibonding σ* or π* molecular orbitals. In
general, the electrons originate from the highest occupied molecular
orbitals (HOMOs). There are other techniques in which core
electrons are excited or even ejected from the molecular systems, but
these techniques usually require photon energies beyond the
wavelength limit discussed here, namely, below about 190 nm.
In this chapter, spectroscopy in the more classical ultraviolet–visible
(UV‐vis) spectral range will be discussed, again from the viewpoint
introduced earlier, namely the interplay between quantum
mechanical theory and the experimental results that can be and have
been used for decades to refine the theory. Furthermore, chemical
structural information can be obtained from UV‐vis spectra,
although this information is not as direct as that obtained from
rotational spectroscopy (Chapter 6) or NMR spectroscopy (Chapter
8).
In order to discuss electronic or UV‐vis spectroscopy, we need to
understand the electronic structure and bonding of the molecular
samples. Whereas classical models are available for rotational and
vibrational energy states of molecules (see Sections 6.1 and 5.1), such
simple models do not exist for electronic structures and energy states
of molecules; thus, a quantum mechanical description is necessary.
This will be presented in Section 10.1 in a somewhat cursory fashion,
because a detailed treatment of molecular orbital theory is far
beyond the scope of this book, in particular the computational
methods that have been developed in this field. The discussion
follows the development presented in a textbook Physical Chemistry
by Engel and Reid [1].

10.1 Molecular Orbitals and Chemical


Bonding in the H2+ Molecular Ion
The first step in understanding electronic spectroscopy is a
description of the chemical bond in molecules, in particular the
formation and energetics of molecular orbitals. This discussion starts
with the simplest of all molecules, the H2+ molecular ion that
consists of two protons and one electron. In analogy to the
Hamiltonian that was introduced for a system consisting of a nucleus
and two electrons (Eq. [9.1])

(9.1)

one defines the Hamiltonian for the H2+ molecular ion with two
nuclei and one electron as

(10.1)

This equation will need to be augmented for electron–electron


repulsion terms in a H2 molecule with two electrons. Equation (10.1)
cannot be solved exactly due to the interdependence of the spatial
variables R, ra, and rb. However, it can be solved explicitly after
invoking the so‐called Born–Oppenheimer approximation (see
Section 10.4): since the nuclei are about 2000 times heavier than
electrons, they move much more slowly than the electrons. Thus, one
may treat the nuclear coordinates as fixed in their equilibrium
position while the electron carries out its motion independently of
the nuclear position. This removes the dependence of the total
energy on R. Next, the electronic wavefunction of the single electron
is written to be delocalized over the two nuclei. One may view the
molecular system in the following discussion as a hydrogen atom
with a nucleus referred to as “a,” to which a proton, without an
electron (nucleus “b”), is added. Then, the wavefunction ψ of the one
electron is written as a linear combination of two hydrogen atom
wavefunctions ϕ located at nuclei a and b as follows:

(10.2)
Because of symmetry arguments and the fact that the electron
density must be invariant to interchange of the two nuclear
positions, the condition

(10.3)
holds. This leads to the formation of two molecular orbitals that are,
due to the symmetry (see Chapter 11) of the species, referred to as the
gerade (g, or even) and ungerade (u, or odd) states:

(10.4a)

(10.4b)

The values of the expansion coefficients cg and cu can be determined


by normalizing the wavefunctions. For ψg,this yields
(10.5)

Since the hydrogen 1s orbital is a normalized wavefunction, the first


two integrals in Eq. (10.5) have a value of one. Thus,

(10.6)

We define the overlap integral

(10.7)

that is, the area that lies under both the 1sa and 1sb atomic orbitals
(cf. Figure 10.1a). With this definition, the normalized coefficient cg
becomes

(10.8)

and

(10.9)

Next, the energy expectation values for the “u” and “g” states will be
discussed. As usual, the energy expectation value is given by (see
Postulate 4, Section 2.1)

(10.10)
Equation (10.10) is one of two energy eigenvalues of the Hamiltonian
that was defined in Eq. (10.1) but subsequently was simplified to

(10.11)

after invoking the Born–Oppenheimer approximation.

Figure 10.1 (a) Overlap of the two 1s orbitals on nuclei a and b.


The volume obtained by rotating the shaded area about the r‐axis
represents the overlap integral. (b) Energy level diagram for bonding
and antibonding orbitals in the H2+ molecular ion.
Taking into account the normalization condition for the denominator
of Eq. (10.10), the energy of the “g” state becomes

(10.12)

or
(10.13)

where the abbreviations Haa and Hba denote

(10.14)

(10.15)

Thus, we may write

(10.16)

and

(10.17)

Next, the expression for Haa is evaluated:

(10.18)

The first integral in Eq. (10.16) can be solved exactly, since it denotes
the energy ground state, E1s, of the hydrogen atom. The second term
is a constant for a fixed value of R and evaluates to just e2/4πεoR,
since the atomic orbital wavefunctions are normalized. The third
term is referred to as the Coulomb integral and expresses the
interaction energy of a positively charged nucleus “b” with the
electron in the 1s orbital of nucleus “a.” Thus, Eq. (10.18) is rewritten
as

(10.19)

Haa is the energy of a hydrogen atom perturbed, in the absence of a


chemical bond, by a naked proton at a distance R. Thus, one may
consider Haa the energy of a nonbonded state. The strength of the
bond formed between the nuclei by delocalization of the electron (see
Eq. [10.4a]) will be given by Eg − Haa.
Similar to the integral expression in Eq. (10.18)), the terms
Hba = Hab are evaluated:

(10.20)

(10.21)

With the definitions of Haa and Hba, the energy changes for the
lowered “bonding” state ΔEg and the raised “antibonding” state ΔEu
are [1]

(10.22)

(10.23)

The results in Eqs. (10.16, 10.17, 10.22, 10.23) can be summarized as


follows for the H2+ molecular ion (and similarly for more
complicated molecular systems, see Section 10.2): two atomic
orbitals, , are combined to form two molecular orbitals,
the bonding σ molecular orbital and the antibonding σ* molecular
orbital. The energy for the bonding orbital is lowered by the amount,
ΔEg, given in Eq. (10.22) that accounts for the formation of the
chemical bond. The absolute value of ΔEg is less than that of ΔEu
since (1 + Sab) > (1 − Sab); thus, the energy lowering of the bonding
orbital is less than the energy increase of the antibonding orbital (see
Figure 10.1b). The wavefunctions of the resulting bonding and
antibonding orbital are similar to those shown in Figure 10.2 for the
H2 molecule.
Figure 10.2 Wavefunctions for the (a) bonding and (b)
antibonding molecular orbitals in H2. Notice the larger overlap of the
1s orbitals in H2, as compared with H2+ (Figure 10.1a), due to the
shorter bond length in H2 (74 pm vs. 105 pm).

10.2 Molecular Orbital Theory for


Homonuclear Diatomic Molecules
For a molecule with more than one electron, one proceeds in analogy
with the discussion presented in Section 10.1 for the H2+ molecular
ion. For the H2 molecule, for example, one could form molecular
orbitals σi by combining the atomic orbitals ϕi of each of the
hydrogen atoms according to

(10.24)
assuming that a set of interacting electrons in a molecule may be
represented by a sum of single‐electron Hamiltonians. Next, one
minimizes the energy of the system by varying the expansion
coefficients that form the molecular orbitals. This is illustrated for
the case of two atomic orbitals in a H2 molecule. Again, one starts by
writing the energy expectation value (see Eq. (10.9))
(10.25)

(10.26)

(10.27)

Next, the values of c1 and c2 corresponding to the minimum value of


E need to be found. This is accomplished by calculating ∂E/∂c1 and
∂E/∂c2 and setting the derivatives to zero. This leads to two
equations:

(10.28)

These equations have nontrivial solutions when

(10.29)

The solutions of this determinant are

(10.30)

in analogy to the solutions for the H2+ molecular ion (Eqs. [10.16]
and [10.17]). Details of the intermediate steps can be found in [1].
The resulting molecular orbital wavefunctions are shown in
Figure 10.2.
The one‐electron orbitals used in these calculations are of the form

(9.4)

that is, spherical hydrogen‐like orbitals with a parameter ζ that


accounts for the size of the orbital. Notice that the mathematical
form of these orbitals was introduced earlier (Eq. [9.4]). The
expressions for H11, H12, and S12 contain the overlap, Coulomb, and
exchange integrals defined earlier (see Eqs. [10.7], [10.18], and
[10.20]):

(10.7)

(10.18)

(10.20)

written here in a more general form of atomic orbitals. The integrals


in Eqs. (10.7), (10.18), and (10.20) are generally solved by
approximating the orbital function by a sum of 3 or 5 Gaussian
functions. Gaussian functions are used to facilitate the integration
necessary in the computation of the Coulomb, exchange, and overlap
integrals, although Gaussian functions do not exhibit the cusp at r =
0, as the true wavefunctions do (cf. Figures 7.1 and 7.4). This
introduces a small error, but the orbital interactions described in
Eqs. (10.7), (10.18), and (10.20) are most significant at values of r
closer to the bond distance, rather than at very small values of r.
The formalism introduced so far neglected electron correlation that
was discussed in Section and restricted the molecular orbital
expansion (Eq. (10.24)) to just two atomic orbitals. In typical
molecular orbital calculations, the expansion given in Eq. (10.24)
includes a much larger number of orbitals even if they are not
occupied:

(10.31)
This approach is referred to as the linear combination of atomic
orbital (LCAO) method. In addition, there exist more sophisticated
molecular orbital calculations that explicitly include electron
correlation, which was neglected in this approach described here.
Second‐row diatomic molecules are treated in a similar fashion by
adjusting the expansion coefficients in Eq. (10.31) to minimize the
energy of the molecule. Since these calculations seek to establish the
minimum energy of the molecular ground state, the variation
method (see Appendix 2) can be employed. These computations
result in molecular orbital energy schemes similar to the one shown
in Figure 10.3 for the O2 molecule. The “inner” (1s) electronic
orbitals form bonding σ1s and antibonding molecular orbitals
that are completely filled with two electron pairs and resemble the
2s, σ2s, and orbitals shown at the bottom of Figure 10.3a. These
electrons do not contribute to the bonding since the energy increase
of the antibonding orbitals is larger than the energy reduction of the
bonding orbitals; see Eqs. (10.22) and (10.23) and Figure 10.1b. In
Li2, the 2s electrons form molecular orbitals that are described as
bonding σ2s and antibonding molecular orbitals.
Figure 10.3 (a) Energy level diagram of the MOs formed from the
overlap of 2s and 2p orbitals in diatomic homonuclear molecules
such as N2, O2, and F2. (b) Visualization of the “head‐on” overlap of
the 2pz orbitals to form the σ2p and the lateral overlap of the 2px and
2py orbitals to form the two π2p orbitals.
The elements that have electrons in 2p orbitals form two types of
molecular orbitals: if we define the z‐axis as the direction of the
chemical bond, the overlap of the 2pz orbitals forms bonding and
antibonding molecular orbitals that are referred to as σ2p and ,
respectively. The lateral overlap of the 2px and 2py atomic orbitals of
the two atoms is responsible for the double‐ or triple‐bond molecular
orbitals that are referred to as the π2p and orbitals, as shown in
Figure 10.3b. Notice that in these orbitals, the highest electron
density is not along the direction of the chemical bond.
The MO energy diagram for a homonuclear diatomic molecule is
given in Figure 10.3a. The order of the orbital energies shown holds
for N2, O2, and F2. Since the molecular orbitals are degenerate,
the oxygen molecule, with 12 valence electrons, has two unpaired
electrons in these orbitals. Notice that the (inner) 1s atomic orbitals,
and the resulting σ1s and molecular orbitals, are not shown in
Figure 10.3a, since they are much lower in energy than the orbitals
formed by the valence electrons.
The energy level diagram for the oxygen molecule shown in
Figure 10.3a and elaborated upon in Problem 10.3 explains very
impressively two facts about this species: first, the molecule has a
bond order of two1, and second, it is paramagnetic. The bond order
of two is expected from a naïve attempt to write a Lewis structure for
O2 and from the fact that the bond stretching vibrational frequency is
typical for a double bond. However, such a Lewis structure would
have no unpaired electrons, which contradicts the fact that molecular
oxygen is paramagnetic and, therefore, must possess unpaired
electrons. These two electrons occupy the two orbitals with
parallel spins. These aspects will be discussed in more detail in later
sections, see also Problems 2–4 of Chapter 10.

10.3 Term Symbols and Selection Rules for


Homonuclear Diatomic Molecules
Based on the energy level diagram in Figure 10.3, O2 in its ground
state has an electronic configuration of

(10.32)

This ground‐state configuration obeys Hund's rule (see Section 9.3)


in that the degenerate orbitals are filled with two electrons with
parallel spins. When writing the term symbols of molecular species
such as the O2 molecule, one proceeds the same way as one would in
the assignment of the term symbols of multi‐electron elements (see
Eq. [9.15]) by summing orbital angular momenta for electrons in
unfilled subshells according to

(10.33)

The resulting Λ values are designated as given in Table 10.1.


Here, electrons in a σ orbital have l values of zero, whereas electrons
in π orbitals have l = ± 1. Thus, in oxygen, the Λ value is either 0 (for
the ground state) or ±2,leading to either Σ or Δ terms for the oxygen
molecule; see Figure 10.4.
The total spin angular momentum is obtained as before (see Eq.
[9.17]) according to

(10.34)

Since there are two unpaired electrons in the oxygen ground state,
the total spin moment S = ± 1. The spin multiplicity 2S + 1 is added
as a left superscript to the term, resulting in a 3Σ designation for the
oxygen ground state (a triplet state). In order to uniquely define the
electronic configuration of a homonuclear diatomic molecules, two
further indices are necessary. A right subscript g or u indicates
whether the orbitals in which the two electrons are
found have even or odd symmetry with respect to the center of
inversion (see Chapter 11). The two MOs formed from the lateral
overlap of 2p orbitals with the same phase have even parity; thus, the
term for the ground‐state oxygen molecule is written as 3Σg. The final
symbol, written as a right superscript + or –, indicates whether MOs
change sign when reflected by a plane that contains the molecular
bond direction. The orbitals do so, and, therefore, the final
designation of the molecular oxygen ground state is . The
electrons in two MOs are shown schematically in panel (A) of
Figure 10.4.
Table 10.1 Symbols of states for different l and L values.
L 0 1 2 3 (For multi‐electron atoms)
S P DF
Λ 0 1 2 3 (For molecules)
ΣΠΔ Φ
Figure 10.4 Electron and spin populations in the two MOs of
the lowest‐energy configurations of the oxygen molecule. See text for
details.
There are several excited‐state species of the oxygen molecule that
are either singlet or triplet states. The next lowest energy state is a
1Δ species in which the two antibonding electrons are, with opposite
g
spins, in one of the MOs; see Figure 10.4b. Since the spins are
paired, it is obviously a singlet state. Another singlet state is due to a
configuration with both orbitals occupied by one electron, but
the electrons have opposite spins (or do not obey Hund's rule). The
total spin angular moment is zero, but the total orbital angular
momentum is 2, as in the ground state. Thus, the term symbol for
this state is . This state is shown in Figure 10.4c.

The three states described so far have the same bond order, since
they differ only in the way the two highest orbitals are populated, as
shown in Figure 10.4a–c. The next state in energy is one where one
electron from the doubly occupied π2p orbital (see Eq. (10.32)) is
promoted into one of the MOs, leading to two electronic
configurations:

(10.35)

one of which is shown schematically in Figure 10.4d. These two


states represent a quite different situation, since the bond order in
this species is no longer 2. Some of the consequences of the changes
in bond order will be discussed in more detail in Section 10.4. The
term symbol for these states are and .
The selection rules for electronic transitions are as follows:

(10.36)
Thus, Σ↔ Σtransitions will be allowed, as will be the Π↔ Σ
transitions. However, singlet to triplet transitions are forbidden. For
homonuclear diatomic molecules, additional rules hold:

(10.37)

With these rules in place, we can start to interpret observed


electronic spectra of diatomic species like O2, for which the only
dipole‐allowed transition is from the ground state to the state,
that is, the transition, for which all selection rules listed
above are fulfilled. The energy difference between these two states is
about 6.2 eV. The resulting UV absorption spectrum is presented in
the next section.

10.4 Electronic Spectra of Diatomic


Molecules
10.4.1 The Vibronic Absorption Spectrum of Oxygen
Based on the discussion in the previous section, one may expect a
single transition peak in the UV spectrum of molecular oxygen for
the transition at just above 200 nm. Instead, one
observes a number of transitions, as shown schematically in
Figure 10.5 [2, 3]. The appearance of the spectral results immediately
suggests that, in addition to the electronic transition, one observes
excitation into vibrational and even rotational states.
The major sawtooth‐like features shown in Figure 10.5a are due to
vibronic transitions from the ground vibrational state of the
electronic ground state into vibrationally excited state of the
electronic state. As can be seen from the band heads in Figure 10.5b,
the energy spacing between the vibrationally excited states of the
decreases, as expected from a strongly anharmonic oscillator.
The spacing of the band heads, however, is much less than what one
would expect from the spacing of the vibrational energy levels of the
electronic ground state. The reason for this can readily be explained
as follows.
O2 in its ground state has a bond order of two. Therefore, the O–O
stretching frequency is 1580 cm−1, typical for double‐bonded species.
This transition is not allowed in absorption but detected in Raman
scattering. However, the spectral progression shown in Figure 10.5a
corresponds to a much lower vibrational energy level spacing of c.
700 cm−1. This is due to the fact the state that is accessed by the
electronic transition at c. 200 nm has a bond order of one. This
follows from an inspection of the MO energy diagram shown in
Figure 10.4d. Since oxygen in the state has a single bond, one
expects the much lower vibrational frequency.
The rotational constant of the excited‐state oxygen species is 0.819
cm−1, resulting in very closely spaced (c. 0.01 nm) transitions for low
J values (“J” here refers to the rotational quantum number; see
Section 6.3 and Eq. [6.38]), but up to 20 cm−1 for high J values when
considering centrifugal interactions. Inspection of Figure 10.5b
shows the vibrational bands [2] with decreasing energy splitting,
whereas the splitting of the rotational energy levels increases for
increasing J values.
Figure 10.5 Observed (a) and simulated (b) vibronic
transition of molecular oxygen. See text for details.
Source: Trentmann et al.[2]; Stamnes [3].

The observation of the progression of vibrational sub‐bands in the


observed UV absorption spectrum of the oxygen molecule needs
further discussion. As pointed out earlier (cf. Sections 4.5 and 6.5),
oxygen does not exhibit neither a vibrational nor rotational
absorption spectrum, being a nonpolar molecule. Obviously, the
electronic selection rules discussed above (Eqs. (10.25, 10.26))
overwrite the vibrational and rotational selection rules. This is
discussed in Section 10.4.2 and an approximate energy level diagram
for all these transitions is given in Example 10.1.

Example 10.1 Schematic energy level diagram for O2,


approximately drawn to scale. The energy difference between the two
electronic states is ca. 50 000 cm−1; the vibrational energy levels are
split by ca. 1500 cm−1 in the ground electronic state and by about
700 cm−1 in the electronically excited state. They are represented
here without the anharmonic contributions. The rotational energy
levels are spaced so closely that they only would be visible at this
scale for J levels larger than ca. 25. Remember that at room
temperature, all O2 species are in the electronic ground state, and
nearly all of them are in the n = 0 vibrational state. However, the
rotational states of the vibrational ground state are highly populated.
Figure 10.E1 See Example 10.1 for details.
Furthermore, the appearance of a broad absorption continuum
below ca. 175 nm (see Figure 10.5a) suggests that the excited oxygen
molecule has reached its dissociation limit after about 15 vibrational
energy levels. Any photons of energy corresponding to 175 nm or less
no longer must have the exact energy of a molecular transition;
rather, photons with energies higher than the highest bound state
will cause photodissociation of the excited species, with the excess
energy being converted into kinetic energy of the two oxygen atoms.
The intensity pattern observed for the procession of vibrational
transitions needs further explanation, and a further discussion of the
consequences of the Born–Oppenheimer approximation is
introduced earlier. This leads to the so‐called Franck–Condon rules
of vibronic spectroscopy that we shall encounter in the next section
and in the discussion of fluorescence spectroscopy as well.

10.4.2 Vibronic Transitions and the Franck–Condon


Principle
The Hamiltonian of the Schrödinger equation for a multinuclear,
multi‐electron molecule or molecular ion contains the following
terms:

(10.38)

This Hamiltonian, in the words of I. N. Levine, is “formidable enough


to strike terror in the heart of any quantum chemist” [4] (Chapter
13). Benzene, for example, has 12 atomic nuclei and 42 electrons. The
time‐independent Schrödinger equation, which must be solved to
obtain the energy and wavefunction of this molecule, is a partial
differential eigenvalue equation in 162 variables, the spatial
coordinates of the electrons and the nuclei. Invoking the Born–
Oppenheimer approximation that uncouples the motion of nuclei
and electrons reduces the problem to 126 dimensions, but it still is a
very complex calculation. We shall use the Born–Oppenheimer to
further explain the vibronic spectral pattern observed for oxygen.
The Born–Oppenheimer approximation can mathematically be
formulated as follows:

(10.39)

In Eq. (10.39), the expression ψ(ri, Rα) denotes the wavefunctions in


all variables defined by the Hamiltonian in Eq. (10.38), namely, the
coordinates Rα of all nuclei α, as well as the coordinates ri of all
electrons i. Because of the fact that the nuclei are much heavier than
the electrons, one argues that their motion is much slower than that
of the electrons; consequently, the wavefunctions are approximated
in Eq. (10.39) as a product of the electronic wavefunctions at fixed
nuclear positions and the purely vibrational wavefunctions that have
been described in Chapters 4 and 5.
The dipole transition moment for the observed spectrum shown in
Figure 10.5 is given by

(3.15)

where n and m denote the states between which the transition


occurs. Thus, the transition moment for the (the lowest‐
energy allowed electronic transition in oxygen) must be nonzero:

(10.40)

Since the position of the nuclei is fixed for the electronic


wavefunctions , Eq. (10.40) can be separated into the
product of an electronic transition moment for fixed nuclear
positions and the overlap of the vibrational wavefunctions in any of
the vibronic states as follows:

(10.41)

The transition moment 〈μ〉 therefore can be written as a product of


the electronic transition moment

(10.42)

and the overlap S of the vibrational wavefunctions (also known as the


Franck–Condon factor)

(10.43)

Within the Born–Oppenheimer approximation, the transition


moment therefore is given by the product of the electronic dipole
operator and the vibrational overlap (the Franck–Condon factor):

(10.44)
Equation (10.44) determines whether or not transitions between
vibronic levels are allowed, as alluded before in the statement that
the electronic selection rule modifies both the rotational and
vibrational selection rules. However, it is the overlap of the
vibrational wavefunctions that determine the intensities of the
particular vibronic transitions, as shown in Figure 10.6.
This figure introduces a scheme commonly used in depicting
vibronic processes. In this picture, the lower part of the two
anharmonic potential curves represents the electronic ground‐state
potential energy in a diatomic molecule, such as O2, or any typical
bond stretching coordinate in a polyatomic molecule. The squares of
the vibrational wavefunctions for the four lowest energy states of this
vibrational coordinate are also shown. The upper potential curve
represents the excited electronic state. In the case of the O2 molecule,
it was pointed out that this electronically excited state has a lower
bond order than the ground state. Therefore, the vibrational force
constant will be lower and the bond is weaker. The weaker bond is
indicated by the fact that the potential energy curve is broader and
shallower and the vibrational energy levels are more closely spaced.
Furthermore, the excited potential function is shifted along the
internuclear axis (the x‐axis) toward a slightly longer bond distance.
Therefore, the maxima of the ground‐state vibrational wavefunctions
in the ground and excited electronic states no longer line up. The
Franck–Condon factor predicts, for the example shown, that the
most intense transition would occur between the ground vibrational
state of the electronic ground state and the fourth vibrational state of
the electronically excited state. A transition into the third vibrational
state of the electronically excited state would have nearly the same
intensity. In these considerations, it is assumed that only the ground
vibrational state of the electronic ground state is populated (see
Example 3.1). The vibronic spectrum, therefore, appears as a series
of nearly equidistant bands, as shown in Figure 10.5a where the
highest intensity band corresponds to the maximum overlap of the
vibrational wavefunctions of the ground and excited electronic
states. This discussion of vibronic transitions following the Franck–
Condon rules will be picked up again in Section 10.6.
Figure 10.6 Vibronic transition between the ground vibrational
state of the electronic ground state and the third vibrational state of
the electronically excited state, according to the Franck–Condon
principle.

10.5 Qualitative Description of Electronic


Spectra of Polyatomic Molecules
In the previous section, the electric transitions in a simple diatomic
molecule were discussed. For larger molecules, a different, simplified
approach is generally used, since the treatment of polyatomic
molecules by the same approach would become very cumbersome,
for several reasons. First, the reduced symmetry of larger molecules
makes the selection rules less stringent (see Chapter 11). Second,
larger molecules often will be in condensed phases, and rotational
transitions cannot be observed. Furthermore, the density of
vibrational states (remember, there are 3N‐6 normal modes) often
prevents progressions like the one shown in Figure 10.5b from being
observed.
This simplified and more qualitative approach pursued here utilizes
the concept of localized electronic transitions in “chromophores”
(literally “color carriers”), that is, groups in which the transition
occurs. This is based on the observation that molecules that contain a
carbonyl group, for example, exhibit similar UV spectral features
regardless of the exact environment the carbonyl group; however,
this concept contradicts the fact the electronic states between which
the transitions occur are delocalized molecular orbitals which vary
from molecule to molecule. This is somewhat analogous to the
approach using group frequencies in vibrational spectroscopy
although we know that a “group vibration” always extends over the
entire molecule.

10.5.1 Selection Rules for Electronic Transitions


For diatomic molecules, we had the selection rule

(10.36)
For polyatomic chromophores in larger molecules, it is not possible
to define good quantum numbers; therefore, the first condition in
Eq. (10.25) no longer applies, and transitions need to be evaluated on
an individual basis, using group theory (Chapter 11) and the
symmetry of the ground and excited states. What remains is the
condition

which implies that singlet to triplet transitions remain forbidden.


This particular aspect will determine many aspects of fluorescence
and phosphorescence and will be discussed in Section 10.6.

10.5.2 Common Electronic Chromophores


The term “chromophores” applies to any part of the molecules,
generally double‐bonded atoms or metal atom that exhibit a
somewhat “localized” electronic transition somewhere in the
ultraviolet to visible part (190–800 nm) of the electromagnetic
spectrum. Such chromophores include carbonyl groups, the peptide
linkage, aromatic groups, or metal ions in coordination compounds
such as myoglobin.

10.5.2.1 Carbonyl Chromophore


A carbonyl “chromophore,” for example, may be visualized by
inspection of an approximate MO energy scheme of the orbitals
involved in a carbonyl group in a molecule such as formaldehyde,
H2CO, see Figure 10.7a. These orbitals can be described, in order of
increasing energy, as the σC − O bond, the πC − O bonding orbitals, the
nonbonding orbitals of the oxygen nO, and the antibonding
and orbitals, as shown in Figure 10.7a. The
transition occurs at about 195 nm and is dipole allowed with a molar
extinction coefficient of about 103 [L/mol cm]. The
transition, on the other hand, is electric dipole forbidden and is
observed as a weak band at about 275 nm with a molar extinction
coefficient of c. 20 [L/mol cm], resulting in an absorption spectrum
shown schematically in Figure 10.7b. The values of the transition
wavelength and dipole strengths are reported here as approximate
numbers, since they depend on the exact nature of the molecule and
its physical state, the chemical environment such as solvation, etc.
Yet, most ketones larger than formaldehyde exhibit quite similar UV
absorption spectra.

Figure 10.7 (a) Approximate MO energy level diagram and UV


transitions for a carbonyl chromophore. no denotes a nonbonding
orbital on the oxygen atom. (b) Simulated UV absorption spectrum
of a carbonyl chromophore with extinction coefficients ε of 1000 and
20 for the π*←π (at 195 nm) and π*← nO (at 275 nm) transitions,
respectively.
The peptide linkage shown below incidentally has a very similar UV
absorption spectrum with a high‐energy (c. 195 nm)
transition and a lower‐energy transition that contribute
prominently to the protein circular dichroism (CD) spectra discussed
in Section 10.7.
10.5.2.2 Olefins
The C=C double bond exhibits the π* ← π transition around 180 nm,
depending on the molecular species and chemical environment. This
transition is dipole allowed with a molar extinction coefficient of
>1000 [L/mol cm]. As discussed in Section 2.5, molecules containing
conjugated double bonds have lower‐energy π* ← π transition but
with higher extinction coefficients. In 1,3‐butadiene, for example, the
transition occurs at c. 220 nm with a molar extinction coefficient of
∼20,000 [L/mol cm]. With increasing number of double bonds, the
transition energy is lowered even more. In carotene, where the
conjugated chain incorporates nine double and ten single bonds, the
transition has been shifted all the way into the visible range, with two
prominent peaks at c. 450 and 500 nm and a molar extinction
coefficient of 140 000 [L/mol cm].
All four DNA bases adenine (A), guanine (G), thymine (T), and
cytosine (C) have highly conjugated C=C, C=O, and C=N
chromophores that absorb in the 240–280 nm region. Double‐
stranded DNA has a broad, featureless band at 260 nm that exhibits
sensitivity to base‐stacking interaction: upon undergoing a transition
from double‐ to single‐strand forms, the absorption intensity
increases by about 20%. The reduction of the absorption upon base
stacking is known as hypochromicity, which is an interaction that
can be explained by the exciton theory discussed in Section 10.7.4.

10.5.2.3 Benzene
Another common chromophore in UV spectroscopy are aromatic
benzene derivatives such as phenyl groups. Benzene itself has an
orbital energy diagram for the six most energetic electrons as shown
in Figure 10.8a. The lowest energy transition from the HOMO to the
lowest unoccupied molecular orbital (LUMO) occurs at 255 nm with
a molar extinction of 180 [L/mol cm]. As described before for the
electronic absorption spectrum of molecular oxygen, the spectrum of
benzene shows a progression of bands as shown in Figure 10.8b. This
progression is due to vibronic transitions into the symmetric ring
breathing mode in benzene. This transition is symmetry forbidden in
infrared (IR) absorption but produces an extremely strong Raman
peak at c. 992 cm−1. The transition into vibrationally excited states of
the excited electronic state, however, is dipole allowed and produces
the progression of peaks that are spaced by vibrational quanta.

10.5.2.4 Other Aromatic Molecules


Toluene and phenylalanine exhibit UV spectra quite similar in
wavelength and intensity to benzene. In both cases, vibronic effects
are observed, although not as well resolved as in benzene. Aromatic
amino acids (tyrosine and tryptophan) exhibit similar UV features,
but these are shifted toward longer wavelength, presumably due to
more delocalized aromatic systems, particularly in the case of
tryptophan.
Figure 10.8 (a) Approximate molecular orbital energy level
diagram of the highest occupied and lowest unoccupied molecular π
orbitals of benzene. The gray vertical arrow corresponds to the
electronic transition shown in panel (b). (b) Vibronic structure of the
transition shown in panel (a). The vibrational structure is due to
Franck–Condon transitions into the symmetric ring breathing
vibration of benzene.

10.5.2.5 Transition Metals in the Electrostatic Field of Ligands


Metal atoms or ions in tetrahedral or octahedral ligand
environments also constitute chromophores that are generally
described in terms of ligand field theory or in terms of crystal field
theory for solids. Since many transition metal complexes have either
four or six ligands, the symmetry of the resulting complexes either
belongs to the Td or Oh point groups (see Chapter 11). In both cases,
the ligands provide a perturbation in space that effects the energetics
of the 3d orbitals.
In octahedral ligand fields, the energy of the dxy, dxz, and dyz orbitals
that transform as the t2g irreducible representation is lowered by the
interaction with the ligand orbitals. The energy of the and
orbitals that transform as the eg irreducible representation is raised,
and the degeneracy of the five original 3d orbitals is split into a triply
degenerate and doubly degenerate set of orbitals. The colors of
transition metal complexes are primarily due to electronic transition
between these two sets of orbitals and depend of course on the
energy difference Δo between these orbitals. Here, the subscript “o”
denotes an octahedral ligand filed. Typical values of Δo correspond to
visible or near‐IR transition energies and depend, in turn, on the
electron density of the ligand. The energy splitting Δo also
determines whether the complex is “high spin” or “low spin”: in Fe2+
complexes, for example, the six iron valence electrons can either all
occupy the three t2g orbitals with completely paired spins (a “low‐
spin” situation) or distribute in all five orbitals (“high‐spin” case). In
the latter situation, the two electrons in the eg orbital would occupy
two different orbitals with parallel spin, and the other four electrons
are in the three t2g orbitals. The total number of unpaired spins is
four in this case.
In tetrahedral ligand fields, the situation is opposite in that the
energy of the doubly degenerate orbitals that transform as the e
irreducible representation is lowered, whereas the triply degenerate
orbitals of the t2 representation are raised in energy. Due to the
different interaction geometry of the 3d orbitals with a tetrahedral
ligand field, the energy splitting Δt is only about one half of Δo. For
more details on the electronic states and transitions in transition
metal complexes, the reader is referred to textbooks on inorganic
chemistry, for example, Cotton and Wilkinson [5].

10.6 Fluorescence Spectroscopy


Fluorescence is a direct consequence of the Franck–Condon
principle discussed in Section 10.4.2 and involves vibronic
transitions to be discussed in the next section. Fluorescence
spectroscopy is a highly useful spectroscopic method that has found
broad applications in cellular biology in the form of fluorescence
microscopic imaging, fluorescence resonance energy transfer, and
elucidation of very fast chemical reaction mechanisms. While
fluorescence is a process that occurs to some extend in many
molecular systems (intrinsic fluorescence), its main applications
involve the use of fluorescent probes or dyes that have been
chemically modified to bind to particular sites in a sample such that
the location and abundance of the binding site can be determined
with high accuracy. A particularly useful fluorescent probe is one
referred to as “green fluorescent protein” (GFP). The gene to produce
this protein is known and available. Thus, this gene can be spliced
into the DNA of a host, for example, into a gene that is responsible
for the formation of a specific organ, say, the liver, in a mammal.
When such a tagged animal grows, it produces an organ that
incorporates GFP, and one literally has an animal that has a green
fluorescent liver. One can only imagine how important such tools are
in developmental biology and cancer research.

10.6.1 Fluorescence Energy Level (Jablonski) Diagram


The discussion of the fluorescence process starts with the same
energy level diagram that was introduced in the discussion of the
Franck–Condon principle in Section 10.4.2 and Figure 10.6 but
includes other pathways that give rise to fluorescence. In
Figure 10.9a, the lower anharmonic potential curve, marked S0,
represents the (singlet) electronic ground‐state potential energy of a
vibrational coordinate in a polyatomic molecule. The upper potential
curve, marked S1, represents the excited electronic state with a
broader potential function shifted along the internuclear axis toward
a slightly longer bond distance.
The fluorescence process consists of a very fast electronic absorption
processes with ΔS = 0 (see Eq. (10.25)) into the excited singlet state
S1 indicated by the black up arrows that are governed by the Franck–
Condon principle. In the case shown, three vibronic states in the S1
manifold are excited. These states may deactivate back to their
original state by spontaneous emission or deactivate by collisional
energy loss, mostly to solvent molecules, into the vibrational ground
state of the S1 state. The collisional deactivation processes are
indicated by the dashed lines in Figure 10.9a. Emission of redshifted
photons from this state into various vibrational states of the
electronic ground state returns the molecule back into the electronic
ground state, indicated by the gray down arrows. This last step again
follows the Franck–Condon principle in that the transitions occur
between states of highest overlap of the vibrational wavefunctions.
The collisional deactivation does not necessarily follow dipole
selection rules; thus, electronic states can be populated that cannot
be excited directly by absorption. Whereas the original excitation
process is very fast, the fluorescent emission occurs at much slower
timescale, about 10−7–10−9 [s] after the excitation. Measurement of
the fluorescent lifetime gives information on molecular dynamics
and solvent exposure of the excited vibronic state.

Figure 10.9 (a) Energy level (Jablonski) diagram for fluorescence.


(b) Energy level diagram for intersystem crossing and
phosphorescence.

10.6.2 Intersystem Crossing and Phosphorescence


Another deactivation process may occur that is shown schematically
in Figure 10.9b. The excitation process into the S1 state is the same as
in Figure 10.9a. What is different is the existence of a triplet state,
marked T1 in Figure 10.9b and indicated by the gray excited‐state
potential function. The direct transition into the triplet state from the
S0 ground state is forbidden by the ΔS = 0 selection rule. If one of the
vibrational energy levels of the T1 and the S1 states are near
degenerate (in Figure 10.9b, the third vibrationally excited state of T1
and the fourth of S1), the excitation energy can be transferred from
the singlet state to the triplet state. This state deactivates through
collisional energy transfer to the ground vibrational state of T1 and is
trapped there. Transitions back into the S1 state are unlikely because
the triplet state is lower in energy than the S1 state and those to the
ground state of the S0 manifold are forbidden. The consequence is
that the excitation energy is trapped in the T1 state and may persist
there for milliseconds or even longer. Weak and slow emission will
occur eventually, which is observed as phosphorescence.

10.6.3 Two‐Photon Fluorescence


Another application of fluorescent methods is that where the
excitation source utilized is an extremely powerful laser with photon
energies half of the amount required to excite a vibronic state in the
S1 manifold. This so‐called two‐photon fluorescence (TPF) is a
nonlinear optical process in which two (near‐IR) photons, typically
from a Ti : sapphire laser, are combined in the medium by the first‐
order non‐linear susceptibility into one photon of half the
wavelength or twice the energy (for discussion of non‐linear
spectroscopic effects, see Appendix 3). This process is represented in
Figure 10.10b. In this case, the fluorescent emission is blueshifted
with respect to the excitation wavelength.
The advantage of TPF spectroscopy is that the near‐IR excitation
wavelength has a much better penetration depth into turbid or
heterogeneous media, for example, into biological tissue. Thus, TPF
can be used to detect fluorescence from depth of many millimeters
inside a sample of human or animal tissue. This technique, combined
with second‐harmonic generation (SHG) (sensitive to collagen) and
coherent anti‐Stokes Raman (CARS) (sensitive to lipids) imaging
modalities in a one microscope setup, has created images of
biological specimens of unprecedented compositional details [6]. The
two nonlinear processes, SHG and CARS, will be discussed in
Appendix 3.
10.6.4 Summary of Mechanisms for Raman, Resonance
Raman, and Fluorescence Spectroscopies
Fluorescence and Raman spectroscopies have similar vibronic
origins and may occur together in an experiment, albeit at quite
different time scales. The following discussion is presented for a
simple diatomic molecule with one degree of vibrational freedom,
although one could consider Figure 10.10 to hold for any vibrational
coordinate, for example, a stretching coordinate in a polyatomic
molecule.
As discussed above in Figure 10.9a, the lower row of potential
functions in Figure 10.10 represents the electronic ground state with
its associated anharmonic vibrational wavefunctions. The upper row
depicts the potential surface for an electronically excited state. A
typical fluorescence process is depicted in panel (A) of Figure 10.10.
Here, the system is promoted into excited vibronic states, as
indicated by the gray arrows. The specific excited vibrational level of
the electronically excited state is determined by the Franck–Condon
principle. Thus, the most likely transitions shown in the example
here are into the n = 2 and n = 3 vibrational states of the
electronically excited state. After slow collisional deactivation (in the
nanosecond regime, shown by dashed lines in panel A) into the
vibrational ground state of the electronically excited state, transitions
occur rapidly into the ground electronic state manifold, again
determined by the Franck–Condon factor. The fluorescence emission
is redshifted with respect to the exciting wavelength and often
exhibits vibronic fine structure, i.e. the fluorescent bands are split by
the vibrational quantum (see Problem 10.12).
Figure 10.10 Schematic diagrams representing fluorescence (a),
two‐photon fluorescence (b), spontaneous Raman (c), and resonance
Raman (d) processes. See text for details.
Panel (B) of Figure 10.10 shows the TPF process. Two photons with
half the energy required to reach the excited vibronic state combine
in a nonlinear process (see Appendix 3) to create one photon with
sufficient energy to reach the excited vibronic state. The process
proceeds via a virtual state (dotted energy level) that is created by the
second‐order nonlinear susceptibility. High laser fields are required
to make this process practical. Collisional deactivation leads to
population of the same state as in standard fluorescence
spectroscopy. The fluorescence emission is blueshifted with respect
to the exciting wavelength. This technique, as pointed out above,
offers the advantage of using low‐energy photons with better
penetration into turbid media to excite the fluorescence.
In contrast to fluorescent processes, the spontaneous, nonresonant
Raman process (far from resonance) is shown in Figure 10.10c. Here,
the dashed horizontal line represents the virtual state that is created
by the polarizability of the molecule. In the resonance Raman
process, the virtual state coincides or is energetically very close to a
real electronic state, as shown in Figure 10.10d. In this case, the
fluorescent photons and the Raman‐scattered photons, indeed, have
the same energy and may overlap in the observed spectra.
Furthermore, the Raman‐scattered intensities may be amplified
thousandfold or more due to the small energy difference in the
denominator of the expression for the polarizability tensor elements
(see Eq. [5.43]). Interestingly, fluorescence and Raman spectra can
be separated by the different timescales at which they occur. The
emission of photons from the virtual state in Raman spectroscopy
takes place in femtoseconds, whereas the fluorescent photons are
emitted nanoseconds or even microseconds after the excitation.
Using pulsed lasers and time‐gated detectors, it is therefore possible
to eliminate fluorescent contributions from the Raman spectra.

10.7 Optical Activity: Electronic Circular


Dichroism and Optical Rotation
10.7.1 Circularly Polarized Light and Chirality
There are molecules that possess a property that allows them to
interact differently with left and right circularly polarized light.
These molecules are said to be chiral or possess “handedness,” and
such molecules exist in two forms known as enantiomers. Left and
right circularly polarized light itself is chiral and can be viewed as a
form of light in which the electric (or magnetic) vector proceeds
along a left‐ or right‐handed helix, rather than oscillates in a plane.
The two forms of circularly polarized light are shown in Figure 10.11
and can be created from two orthogonal, linearly polarized waves
with wavelength λ that are shifted by ±λ/4 with respect to each other
(see Figure 10.11a). The different interaction of a medium with left
and right circularly polarized light is referred to as optical activity.
Figure 10.11 (a) Left (top) and right (bottom) circularly polarized
light. (b) Examples of the two enantiomeric forms of an asymmetric
molecule, fluorochlorobromomethane (top), and a dissymmetric
chiral molecule, 2,3‐pentadiene. The planes in which the methyl
groups lie and the symmetry axis are shown.
Handedness is a property found in macroscopic entities as well as in
molecules. Think about a pair of shoes or a regular bold and nut you
can purchase in a hardware store, which both normally are right‐
handed. A left‐handed nut will never fit on a right‐handed bolt, even
after turning it around, since the handedness is invariant to all
symmetry operations (see Chapter 11) except reflection by a mirror
plane. Thus, your left shoe will be superimposable on the mirror
image of your right shoe. This last sentence actually defines chirality,
in general: an item is chiral if it is nonsuperimposable on its own
mirror image (or superimposable on the mirror image of its
enantiomer).
Chiral molecules fall into two generally classes: asymmetric
molecules that belong to a symmetry group that has no symmetry
element except the identity element (see Chapter 11) and
dissymmetric molecules that belong to purely rotational point
groups. An example of an asymmetric molecule would be any
tetrahedral carbon atom surrounded by four different groups, such
as glyceraldehyde or fluorochlorobromomethane, shown in
Figure 10.11b, whereas an example of a dissymmetric molecule would
be 2,3‐pentadiene (also known as 1,3‐dimethylallene), which has a
C2‐axis through the center C atom and at 45o with respect to the two
planes that contain the methyl groups and the hydrogen atoms.
Helical molecules also belong to the class of dissymmetric molecules.
For a chiral molecule to exhibit optical activity, one of the
enantiomers must be in excess over the other. A 50:50 mixture of
enantiomeric molecules is known as the racemate, and a racemic
mixture does not exhibit optical activity.
Optical activity was discovered in the mid‐nineteenth century by L.
Pasteur who separated mirror image crystals of ammonium sodium
tartrate, a salt of tartaric acid, under a microscope into “left‐ and
right‐handed” forms. When these mirror image crystals were
separately dissolved in water, the resulting solutions rotated the
plane of incident linearly polarized light in opposite directions. The
“optical rotation” is one of several manifestations of electronic
optical activity, that is, an unequal response of the medium toward
left or right circularly polarized light.

10.7.2 Manifestation of Optical Activity: Optical


Rotation, Optical Rotatory Dispersion and Circular
Dichroism
The rotation of plane polarized light described in the previous
paragraph can be explained as follows. Linearly polarized light can
be described as being composed of two co‐propagating circularly
polarized light waves, one of which is left circularly polarized and the
other is right circularly polarized. If the refractive index of the
sample toward right circularly polarized light, nR, differs from that
for left circularly polarized light, nL, the two circularly polarized
waves propagate with different velocity through the medium;
consequently, the polarization plane of linearly polarized incident
light will be rotated upon exiting the sample solution. In the
following discussion, the subscripts L and R represent left and right
circular polarizations. Thus, optical rotation can be described by the
difference in refractive indices toward circular polarization as

(10.45)
Optical rotation was first observed in the visible part of the spectrum
and was found to generally increase in magnitude with decreasing
wavelength of light. Furthermore, it was found that the optical
rotation underwent anomalous dispersion at the wavelength of an
absorption maximum as shown by the top trace in Figure 10.12. The
wavelength‐dependent variation of the optical rotation of a medium
is known as optical rotatory dispersion (ORD). In addition, it was
found that within an absorption peak, there exists a nonzero
difference between extinction coefficients toward left and right
circularly polarized lights:

(10.46)

Figure 10.12 Relationship between ORD and CD. Notice that the
differential refractive index changes sign within a region of
maximum CD signal.
or

(10.47)
This effect is known as circular dichroism. The relationship between
ORD and CD is shown in Figure 10.12 that is analogous to Figure 5.6
with a minor modification, namely, that the differential refractive
index actually undergoes a sign change, whereas the refractive index
shown in Figure 5.6 always stays positive. ORD and CD are often
introduced during the discussion of optical activity and
stereochemistry in introductory courses in organic chemistry.
In the following two sections, the optical activity of asymmetric
molecules and dissymmetric molecules will be presented separately,
although this conceptual separation is somewhat arbitrary but serves
as a pedagogical tool to point out different origins of natural optical
activity.

10.7.3 Optical Activity of Asymmetric Molecules: The


Magnetic Transition Moment
The transitions discussed so far in this book, giving rise to rotational,
vibrational, atomic, and molecular electronic spectroscopies, have
one feature in common: they are caused by the electric field of the
incident light, mediated by the electric dipole operator, according to

(3.20)

The exceptions to this statement are the transitions observed in


NMR that can be attributed either to the electric field and electric
dipole moment or in terms the magnetic dipole operator acting on
net magnetization, as described in Section.
One defines the dipole strength D01 for a transition from the ground
state |0〉 to an excited state 〈1| as

(10.48)

D01 represents the integrated area under an absorption band. Up to


now, we have assumed that only the electronic dipole moment
contributes to the transition, and the effect of the magnetic field of
electromagnetic radiation has been ignored.
There is a class of molecules, however, in which the magnetic
transition moment also contributes to observed transitions, namely,
optically active molecules. For these molecules, one defines a
quantity known as the rotational strength R01 in analogy to Eq.
(10.48):

(10.49)

This quantity corresponds to the integrated area under a CD curve


for a pure enantiomer for a given transition. In Eq. (10.49), μ is the
electric dipole operator defined earlier (Eq. [3.4]) as

(3.4)

and m is the magnetic dipole transition operator defined as

(10.50)

where e, m, r, and p are the charges, masses, positions, and


momenta of particle i and the summation is over all particles. The
vector cross product in Eq. (10.50) is the same that was encountered
before for the definition of the angular momentum operator; see Eq.
(6.2). Thus, the magnetic dipole operator expresses a torque exerted
by the magnetic field on the motion of electrons.
In most manifestations of natural optical activity, the desired
observables – i.e. Δε, ΔA, or R01 – arise through the interference of
electric and magnetic dipole transition moments. The components of
the electric dipole operator transform like translation along the X‐,
Y‐, and Z‐directions, whereas the magnetic moment transforms like
rotation about the X‐, Y‐, and Z‐axis (such as RX, RY, and RZ; see
Chapter 11). Only in molecules that belong to the purely rotational
point groups do the magnetic and electric dipole operators transform
in the same irreducible representation and thereby allow for optical
activity. Molecules in these molecules possess “chirality” or
handedness as introduced above in Section 10.7.1.
The sensitivity of chiroptical techniques toward the handedness of
the molecule results directly from the form of the magnetic
momentum operator: it contains the vector product of momentum
and position vectors, the result of which is another vector. The sign
of this vector is determined by the handedness of the coordinate
system and will change sign upon converting from a left‐handed to a
right‐handed coordinate system. By changing the configuration of
the molecule from one to the other enantiomer, the sign of the
magnetic transition moment changes. Similarly, by keeping the
configuration of the molecule fixed and changing from left to right
circularly polarized light, the sign of the magnetic transition moment
is reversed.
Figure 10.13 (a) CD (top) and UV absorption (bottom) spectra of
an asymmetric molecule, camphor. (b) CD (top) and UV absorption
spectra (bottom) of two achiral but dissymmetrically arranged
chromophores (oval structures) to produce an “exciton couplet”.
Source: Redrawn from Woody [10].

The consequences of the last paragraph are that the enantiomeric


forms of a chiral molecule give exactly equal and opposite CD
spectra, as shown in Figure 10.13a for (R) and (S) camphor. This
result shows the standard form of electronic CD spectra in which an
electronic transition – in this case, the carbonyl π* ← n transition at
ca. 290 nm – is perturbed by a nearby chiral center and exhibits
equal and opposite CD for the two enantiomeric forms. CD is a
relatively weak effect, as measured by the ratio of Δɛ/ɛ, which is
about 10−3 for many chiral molecules.
In the absence of an electronic transition, both CD and optical
rotation of a molecule will be very small since the magnitude of the
magnetic transition moment alone is much smaller than the electric
transition moment. Thus, a chiral hydrocarbon such as 3‐
methylheptane or 2‐deuterobutane does not exhibit any observable
CD above 200 nm and shows minimal optical rotation at 589 nm (the
standard wavelength at which optical rotation is reported).
However, in the presence of a chromophore, the chiral perturbation
of its electronic spectrum has been extremely useful in assigning the
stereochemistry of many compounds. In chiral molecules whose
structures are derived from the androstane skeleton and contain a
carbonyl chromophore (such as testosterone and estrogen), the
observed CD patterns correlate with the location of the chiral center
within the reference framework of the C=O group. These resulting
“quadrant” or “octant” rules were used to establish the
stereochemistry of many of these substituted androstane
structures [7].

10.7.4 Optical Activity of Dissymmetric Molecules:


Transition Coupling and the Exciton Model
The discussion of CD, so far, has emphasized asymmetric species
(belonging to point group C1 that has only the identity symmetry
element) whose optical activity is attributed to the interaction of
electric and magnetic dipole transition moments; see Eq. (10.49).
However, optical activity is also exhibited by dissymmetric species of
pure rotational point groups. These molecules are exemplified by
twisted or helical structures, as shown in Figure 10.11b. Inherently
achiral chromophores or groups can produce optical activity by a
coupling mechanism of the electric dipole moments that is described
by an “exciton” model of optical activity (see Tinoco [8] and Bayley et
al. [9]).
In the exciton model, one considers a chiral polymer, such as peptide
α‐helix composed of n identical peptide units in a fixed chiral
geometry, for example, in a helix. The peptide electronic transitions,
notably the electronically allowed π* ← π transitions, have near‐
degenerate energy eigenvalues En and interact with each other via
dipolar coupling to form a delocalized excited state. A one‐photon
excitation into the manifold of coupled peptide groups results in an
“excitonic” state in which the excitation is highly delocalized between
the coupled excited states. The energies of these excited states are
defined by the eigenvalues of the interaction energies:

(10.51)

where the diagonal values are the degenerate or near‐degenerate


energy eigenvalues of unperturbed transitions and the off‐diagonal
elements Vij are the interaction or coupling energies between the
transition moments, given to a first approximation by dipole–dipole
interactions:

10.52

In Eq. (10.52), μi and μj are the transition dipole moments of groups


i and j, and Tij is the distance vector between them. For the case of n
coupled transitions, Eq. (10.51) is diagonalized numerically to yield
the energy eigenvalues νk of the k'th exciton component and the
eigenvector matrix C, from which the transition dipole strength Dk
for each of the k coupled transitions can be computed according to

10.53

where cik are the appropriate eigenvector matrix elements.


Such molecular systems containing chirally arranged transitions
exhibit rotational strengths for the k'th exciton component that are
given by
10.54

where c is the velocity of light.


For the dimeric case (n = 2), Eq. (10.54) simplifies to the well‐known
“coupled transition” equation [8]

10.55

in which R± denotes the rotational strengths of the symmetric ∣ +〉


and antisymmetric ∣ −〉 combination states of the two transitions. The
resulting CD features are often referred to as an “exciton couplet,”
shown in Figure 10.13b, in which the absorption spectrum shows two
(often overlapping) peaks split by the “exciton” energy. The CD of the
symmetric and antisymmetric states shows a couplet of peaks with
opposite signs, as described by Eq. (10.55). The sign pattern can be
correlated with a right‐ or left‐handed twist.
The exciton formalism conveys a simple model for the optical activity
produced by the dissymmetric interaction of achiral transitions.
Notice that in the presentation here, the magnetic transition moment
was omitted, although it is nonzero as well, but the electronic
coupling contributions are assumed to dominate the exciton
rotational strength.
The most prominent application of the exciton formalism is for the
interpretation of the large CD signals observed for different
secondary structural motifs of peptide chains in proteins, depicted in
Figure 10.14. Here, the large CD signal with a zero crossing at just
over 200 nm for the α‐helix is assigned as the π* ← π exciton
component. Although model calculations are hampered by the
uncertainty in the direction of the transition moment for this
transition, good agreement between observed and computed CD
spectra was achieved [10]. It is interesting that the β‐sheet
conformation exhibits a CD couplet of the same sign pattern as the α‐
helix that leads to the conclusion that the sheet conformation has an
overall twist and is not planar. What is often referred to as the
“random coil” conformation should be rather called the poly‐L‐Pro
(II) structure that appears to have an overall twist opposite to that of
the sheet structure. These peptide secondary structures and models
to reproduce the observed CD patterns have been the subject of
numerous research papers (see [10] and references therein).

10.7.5 Vibrational Optical Activity


In a chiral molecule such as camphor (see Figure 10.13a), one
observes the perturbation of the chromophore – in this case, a
carbonyl group – by the chiral environment. There are, however,
other methods that probe chirality directly at the chiral center. These
methods are (IR) vibrational circular dichroism (VCD) and Raman
optical activity (ROA). In these forms of vibrational optical activity
(VOA), the effects of the chirality on vibrational, and not electronic,
transitions are observed. VCD is an analogue to electronic CD in that
Eq. (10.49) describes VCD, if vibrational wavefunctions are
substituted. ROA, like Raman spectroscopy, is a scattering, rather
than an absorption phenomenon where the scattered intensity
differential contains cross terms between the polarizability tensor
elements defined earlier
Figure 10.14 CD (top) and UV absorption (bottom) spectra of (a)
α‐helical, (b) β‐sheet, and (c) polyproline II (also sometimes referred
to as “random coil”) peptide conformations.

(5.42)

and the magnetic polarizability

(10.56)

Thus, both ROA and VCD are truly forms of vibrational spectroscopy
but take into account the magnetic transition dipole moment that
was omitted in the discussion in Chapter 5. As indicated earlier, both
techniques offer the advantage that they can sample the chirality
right at the chiral center and do not require an electronic transition
in a chromophore. The theory and applications of these techniques
are beyond the scope of this book, and the reader is referred to
abundant literature in the field [11].

References
1 Engel, T. and Reid, P. (2010). Physical Chemistry, 2e. Upper
Saddle River, NJ: Pearson Prentice Hall.
2 Trentmann, J. et al. (2003). Impact of accurate photolysis
calculations on the simulation of stratospheric chemistry. Journal
of Atmospheric Chemistry 44 (1): 225–240.
3 Stamnes, K. (2015). Radiation Transfer in the Athmosphere:
Ultraviolet Radiation, in Encyclopedia of Athmosphereic Scince
(eds. G.R. North, J. Pyle and F. Zhang). Academic Press.
4 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.
5 Cotton, F.A. et al. (1999). Advanced Inorganic Chemistry, 6e.
Wiley.

6 Pope, I. et al. (2013). Simultaneous hyperspectral differential‐


CARS, TPF and SHG microscopy with a single 5 fs Ti:Sa laser.
Optics Express 21 (6): 7096–7106.
7 Crabbe, P. (1972). ORD and CD in Chemistry and Biochemistry.
New York: Academic Press.
8 Tinoco, I. (1962). The exciton contribution to the optical rotation
of polymers. Radiation Research 20: 133–139.
9 Bayley, P.M., Nielsen, E.B., and Schellman, J.A. (1969). The
rotatory properties of molecules containing two peptides groups:
theory. The Journal of Physical Chemistry 73 (1): 228–243.
10 Woody, R.W. (2005). The exciton model and the circular
dichroism of polypeptides. Monatshefte für Chemie 136: 347–366.
11 Nafie, L.A. (2011). Vibrational optical activity: principles and
applications. Chichester, UK: Wiley.
12 Pescitelli, G. Electronic Circular Dichroism.
https://encyclopedia.pub/212

Problems
1. Figure P.1 shows the molecular orbital energy level diagram for
first row (H, He) diatomic molecules. Which of the following
molecules are expected to exist, and what would be the bond
order?
a. H2+
b. H2
c. H22−
d. He2+
Figure P.1
2. Figure P.2 shows the molecular orbital energy level diagram for
the valence electrons of N2. Based on this diagram, answer the
following questions.
a. What is the bond order in N2?
b. Does this bond order agree with the Lewis structure?
c. Is the molecule paramagnetic or diamagnetic?
Figure P.2
3. Figure P.2 also holds for the MO scheme for homonuclear
diatomic molecules O2 and F2. For each of the four species listed
below, draw the MO diagram. From these diagrams, indicate
orbital occupancy, bond order, and magnetic properties
(paramagnetic vs. diamagnetic). Predict the vibrational
stretching frequency, assuming that a single bond contributes a
stretching force constant of 570 [N/m]:
a) O2+
b) O2 ( )

c) O2 (
d) F2
4. In the second‐row diatomic molecules B2 and C2, the MO
energies of the σ2p and π2p orbitals are exchanged as shown in
Figure 3. Predict the electron configuration, bond order,
vibrational stretching frequency (see previous problem),
magnetic properties, and term symbol for the lowest energy
state of B2.
Figure P.3
5. List the two electronic transitions observed in the 180–300 nm
UV spectral region of ketones.
6. Describe a spectroscopic method discussed in the chapter to
monitor the “melting” of DNA from double‐ to single‐stranded
structures.
7. What conditions must be fulfilled for a medium to exhibit
circular dichroism (CD)?
8. a. Name everyday items that possess chirality.
b. What is circularly polarized light?
9. A sample molecule dissolved in a transparent solvent and
contained in a cuvette with 1 cm path exhibits an absorbance of
1.22 in a transition that has a molar extinction coefficient of
1490 [L/(mol cm)] at 280 nm.
a. What is the concentration of the sample?
b. What percentage of the incident light is transmitted at 280
nm?
10. For polyelectronic atoms or ions, write the orbital
approximation, and explain the symbols, notation, and
implications of this approximation.
11. What transitions are responsible for protein absorptions at the
following?
a. About 200 nm
b. About 270 nm
c. Between 270 and 290 nm
d. About 480–600 nm
e. About 1650 cm−1
12. Draw a vibronic energy level diagram and explain the
relationship between the vibronic absorption spectrum and the
corresponding fluorescence spectrum shown in the Figure P.4;
in particular, discuss the origin of the near‐equidistant bands in
both spectra.
Figure P.4
13. Consider the hypothetical MO energy scheme shown in Figure
P.5. Fill in the resulting MOs, and discuss the bond order and
the spin multiplicity of the resulting MO scheme, if each atom
contributes
a. 2 electrons
b. 3 electrons
Figure P.5

Note
1 Bond order is defined as (number of electrons in bonding orbital –
number of electrons in antibonding orbital)/2.
11
Group Theory and Symmetry
In earlier chapters, selection rules for infrared (IR), Raman, and
electronic spectra were discussed, and general rules were developed
to determine whether or not a given transition is allowed or not. For
IR spectroscopy, we found that for dipole‐allowed transitions, the
integral

(5.45)

must be nonzero. This led to the conclusion that a normal mode of


vibration must change the dipole moment of a molecule:

(5.44)

In analogy, for a mode to be Raman active, the polarizability

(5.46)

must change during a normal mode. For a transition between


electronic states to occur, the expression

(10.29)

which holds within the Born–Oppenheimer approximation, must be


nonzero. In the following discussion, the principles of electric dipole‐
allowed transitions will be developed for vibrational spectroscopy
but hold as well for electronic transitions, except that electronic (or
vibronic) rather than vibrational wavefunctions need to be utilized.
In small molecules, such as CO2, simple arguments help determine
which modes are IR and/or Raman active. In more complex
molecules, particularly those of high symmetry, one needs to
establish how the molecular symmetry influences their spectra. This
is the subject of group theory, introduced in this chapter. We shall
see that each normal mode of vibration can be described by a
symmetry representation and that the transformation properties of
these representations will determine whether or not a given normal
mode of vibration is symmetry allowed.
Group theory is a topic that is much too broad to be treated in full
detail in just one chapter; thus, the reader is referred to specialized
textbooks such as the classic Chemical Applications of Group Theory
[1] for derivations and more examples. However, the basic concepts
of group theory as applied to vibrational and electronic
spectroscopies will be presented in this chapter, and a connection
between this of mathematics to multidimensional vector spaces and
linear algebra will be made. The aim of this chapter is to present
sufficient information for the discussion of spectroscopy and the
derivation of symmetry‐based selection rules.

11.1 Symmetry Operations and Symmetry


Groups
Molecules are classified into symmetry‐related categories, called
“symmetry groups,” according to the number and nature of
symmetry operations that can be carried out on a molecule. A
symmetry operation is a procedure applied to a molecule that leaves
it in an arrangement that is indistinguishable from the arrangement
before the operation was carried out. An example of such a symmetry
operation is shown in Figure 11.1: when the water molecule is rotated
by 180° about the axis shown, it appears identical to and
indistinguishable from its original arrangement, since the two H
atoms are indistinguishable. This particular symmetry operation is
called a proper twofold rotation about a symmetry axis and referred
to as a C2 symmetry operation. The symbol “C” will be used for all
proper rotations, which will be discussed below in more detail.
Investigating the symmetry properties of a number of arbitrary
molecules, or even macroscopic objects such as a snow crystal, a
cube, certain letters, and many other items which one intuitively
associates as being “symmetric,” one finds that there exist only a few
distinct and independent symmetry operations. These can be
combined and repeated to create other symmetry operations. For the
discussion of the symmetry properties of small molecules, there are
five symmetry operations of interest: the identity element, proper
axes of rotation, the reflection by a mirror plane, the center of
inversion, and rotation–reflection axes, also known as improper axes
of rotation. These operations will be discussed next.

1. The Identity Element, designated E. This operation leaves the


molecule unchanged, and the necessity for such an operation
appears superfluous at first. However, the identity operation is
necessary for two reasons. The first of these is purely
mathematical and will become clear later during the discussion
of symmetry groups. The other reason can be viewed intuitively
as follows: if one defines a symmetry operation as a procedure
that will return the molecule into a state that is indistinguishable
from the state before the operation, a formalism is needed to
include the possibility that the molecule was, indeed, left
unchanged, that is, the molecule was not operated on at all.
Figure 11.1 Example of a symmetry operation (C2).

Figure 11.2 (a) Example of one of three σν mirror planes in


the molecule chloroform. Notice that the major symmetry axis,
marked C3, is contained in the symmetry plane. (b) Example of a
σh plane that is perpendicular to the C3 axis in the molecule
boron trifluoride.
2. Proper Axes of Rotation (Cn). This operation was introduced
above for a twofold rotation. In general, one designates Cn
operations as rotations about an axis where “n” describes how
often the operation has to be repeated until a rotation by 360°
has occurred. Thus, Cn indicates a rotation about an axis by
360/n degrees (see Figure 11.1).
3. Reflection by a Mirror Plane (σ). Reflection operations can be
visualized as a reflection by a mirror plane that can be parallel or
perpendicular to a symmetry axis. The former is called σν and
the latter σh. Such mirror planes are shown in Figure 11.2.
4. Center of Inversion (i). When pairs of atoms are at positions for
which the coordinates are identically the same except that all
their signs are reversed, then these atoms are related by a center
of inversion. Planar ethene (C2H4) is a molecule that
incorporates a center of inversion (see Figure 11.3), located
halfway between the two C atoms. The hydrogen atoms
numbered 1 and 4 in Figure 11.3 have equal and opposite
coordinates, as do hydrogen atoms 2 and 3 and the two C atoms.
5. Rotation–Reflection Axes or Improper Axes of Rotation (Sn).
This operation consists of a rotation about an axis by 360°/n,
followed by a reflection by a plane perpendiculars to the axis.
The hydrogen atoms in ethane, for example, are related by an S6
operation (Figure 11.4): rotation of the molecule about the
direction of the C–C bond by 60° brings the top three hydrogen
atoms into a position opposite to the original position of the
lower three hydrogen atoms. Reflection of the three upper
hydrogen atoms by a plane bisecting the C–C bond brings the
upper three hydrogen atoms into the positions originally
occupied by the lower three hydrogen atoms.
Figure 11.3 Definition of a center of inversion, located at the
coordinate origin, in planar ethene (C2H4). The coordinates of H1
and H4, for example, are equal in value but opposite in sign.
Figure 11.4 Description of an improper rotation operation (S6)
for ethane.
It is a matter of practice to identify all the symmetry elements of a
molecule. Benzene, with a perfect hexagonal structure, belongs to the
D6h point group and has the following symmetry elements: two C6
axis, two C3 axes coincident with the C6 axis, one C2 axis coincident
with the C6 axis, three C2 axes that bisect opposing sides, three C2
axes that contain opposing atoms, two S6 and two S3 axes, a
horizontal symmetry plane σh, three σν planes that contain opposing
atoms, and three σd planes that bisect opposing bonds. When one
hydrogen atom is substituted by deuterium, or one carbon by a 13C
atom, the symmetry is lowered enormously to just contain one each
of the following elements: E, C2, σν, and σν′.

Certain molecules, such as water, benzene‐d1, and formaldehyde,


H2CO, possess the same symmetry elements, namely, E, C2, σν, and
σν′. (The prime denotes that there are two mutually perpendicular
planes both containing the major axis. These symmetry elements are
also denoted as σ[yz] and σ[xz], respectively.) The occurrence of the
same symmetry elements in different molecules suggests that the
grouping of the four symmetry elements above is, indeed, special,
and it can be shown that the four symmetry elements form a group.
A group in the mathematical sense is a collection of elements that are
connected according to certain rules:

1. Successive applications of any two or more operations of the


group produce another symmetry operation of the group. For
example, successive applications of the C2 operation bring the
molecule back to its original orientation. Thus,

(11.1)
with E being another operation of the group. Similarly,

(11.2)
2. There exists an identity element E such that for any operation A:

(11.3)
As pointed out above, the E operation leaves the molecule
unchanged; thus, any operation A followed by the E operation is
equivalent to performing the operation A only.
3. The associative law holds. Let A, B, and C be symmetry
operations in a group. Then, the order in which they are applied
to a molecule does not matter:

(11.4)
4. Every operation must have an inverse operation. The inverse
operation A−1 for a symmetry operation A is defined such that
AA−1 = E, i.e. the inverse operation negates the original
operation A and leaves the molecule unchanged (or operated on
by the E operator).

For the discussion of molecular symmetry relevant to molecular


vibrations, one needs to be concerned with about 40 different
symmetry groups, of which only about a dozen or so are common in
molecules. The groups are given special designations depending on
the symmetry elements they contain, and the particular group with
the symmetry elements discussed above (E, C2, σv, and σv′) is known
as the C2v point group. In this symmetry, or point group, there are
four symmetry elements, each occurring only once. Thus, one defines
the “order h” of this point group to be 4.
Ammonia, NH3, belongs to a symmetry group known as C3v, which
contains the following elements:

The two C3 axes denote clockwise and counterclockwise rotation by


120° (see Eq. (11.19)), which are two different operations since they
produce different orientation of the hydrogen atoms. The three
symmetry planes are all parallel to the C3 axes and contain one of the
N–H bonds each. Thus, in C3v, there are six symmetry elements in
three classes, and some elements occur more than once in this group.
The order of the group is h = 6. The order of a group and the number
of classes will be used in the next section for the derivation of the
irreducible representations. A few of the common symmetry groups
are listed at the end of this chapter.

11.2 Group Representations


Next, a scheme will be presented to describe the effect of a symmetry
operation mathematically. This is accomplished most easily by
considering a Cartesian coordinate system and following the effect of
a symmetry operation on this coordinate system. This is shown for
an identity (E) and a reflection by a mirror plane, (the yz‐plane) in
Figure 11.5. To avoid confusion about the definitions of the symmetry
planes σν and σν′, these symmetry planes will henceforth be referred
to as σxz and σyz.
For the discussion in this chapter, the following conventions will be
adopted. When a symmetry operation is performed, the original
coordinates are referred to as the x, y, and z coordinates, whereas the
coordinates after the transformation are designated as x′, y′, and z′
coordinates. Then, each symmetry operation can be represented by a
transformation matrix between the old and the new coordinate
system. For example, the E operation in Figure 11.5 can then be
described by

(11.5)

whereas the σyz reflection is described by

(11.6)

These 3 × 3 transformation matrices can be verified by multiplying


out, for example, Eq. (11.6), which yields

(11.7)
This is what one expects by inspection of Figure 11.5.
Thus, the transformation matrices of a Cartesian coordinate system
under the four symmetry operations of the point group C2ν can be
summarized as

(11.8)

The matrix representation introduced in Eq. (11.8) can also be used


to demonstrate, by simple matrix multiplication, that the successive
application of two symmetry operations of a group will produce
another operation of the group. For example, the successive
application of σyz and σxz in the C2ν point group can be represented
as follows:

Figure 11.5 Effects of symmetry operations E and σyz on a


Cartesian coordinate system.

(11.9)

This result implies that the product of the transformation matrices


for two consecutive operations must equal the matrix representing
the product operation, which is another member of the group,
namely, C2.
The traces of the transformation matrices within a group for given
objects can be used to construct what is known as “representations”
of a group. For example, the traces of the transformation matrices for
a Cartesian coordinate system for the E, C2, σxz, and σyz operations in
the C2v point group are 3, −1, 1, 1, respectively (see Eq. [11.8]). The
“vector” 3, −1, 1, 1, therefore, is a “representation” for the
transformation properties of a Cartesian coordinate system for the
four symmetry operations that define the point group C2ν:

(11.10)
where the symbol Γ designates a representation and the numeric
values in the “representation vector” are given the symbols X. Thus,
the notation X(Γi) implies any of the numeric values listed in Eq.
(11.10) for a given representation.
Next, we show that this particular representation, [3, −1, 1, 1], can be
interpreted as a vector in the four‐dimensional space defined by C2v
that can be decomposed into the contributions along four one‐
dimensional unit vectors. These unit vectors are referred to as
“irreducible representations” of a group. These irreducible
representations can be considered a way of representing the
transformation properties of the very simplest objects, such as a
general point. Under C2v, a point lying along the z‐axis (the rotation
axes) would transform symmetrically (+1) under all symmetry
operations of the group and thus have a representation of {1, 1, 1, 1}
for the four operations E, C2, σxz, and σyz. A point on the x‐axis,
however, would transform antisymmetrically (−1) under C2 and σyz
and thus would have a representation {1, −1, 1, −1}. Similarly, a point
on the y‐axis will transform antisymmetrically under C2 and σxz and
has a representation {1, −1, −1, 1}. It can be shown (see below) that
these irreducible representations are the equivalents of unit vectors
in a space whose dimension is given by the number of symmetry
elements in a group. Thus, C2ν, which has four symmetry elements,
will have four irreducible representations, three of which have been
visualized so far:
The theoretical derivation of the irreducible representations is quite
complicated and follows directly from the orthogonality theorem of
group theory (see, for example, [1]), but will not be discussed here in
detail. Some of the consequences of this theorem are as follows:

1. The number of irreducible representations (unit vectors) in a


group is equal to the number of symmetry classes. Thus, C2v will
have four irreducible representations, and the C3v point group,
introduced in Section 11.1, has three irreducible representations,
although it has six symmetry operations.
2. If one defines a number li to be the dimensions of the ith
irreducible representations (the largest number X[Γ]), then

(11.11)

In Eq. (11.11), h denotes the order of the group, which was defined in
Section 11.1 to be the total number of symmetry operations in a
group. Thus, for C2v

(11.12)

and

(11.13)

Thus, C2v has four one‐dimensional representations, three of which


are listed above. The fourth irreducible representation can be derived
from the orthogonality theorem, which postulates that the four
representations (being unit vectors) must be orthogonal to each
other. Thus, in addition to the three irreducible representations
listed above, one writes a fourth orthogonal vector as
which results in a (reorganized) table:

(11.14)

Irreducible representations for a few selected other point groups will


be discussed later. Next, the orthogonality of the four
representations shown in Eq. (11.14) will be demonstrated. For this,
an analogy in three‐dimensional space, with unit vectors i, j, and k
will be used. In three‐dimensional space, defined by a Cartesian
coordinate system, these unit vectors can be represented by the
matrix

(11.15)

In the matrix defining the i, j, and k vectors, the orthogonality of the


unit vectors can be demonstrated by summing the products of the
elements corresponding any two unit vectors, for example, i and j,
over the three directions x and y:

(11.16)

In complete analogy, orthogonality in the four‐dimensional space


defined by C2ν is established by summing over all symmetry
operations of a group the products of the X′s for any two unit
vectors, i and j:

(11.17)
For the table of irreducible representations for C2v listed in Eq.
(11.14), the vector dot product, designated by the symbol ⊗, of Γ1 and
Γ2, is

(11.18)
i.e. the two unit vectors Γ1 and Γ2 are orthogonal. It is easy to
demonstrate that any binary combination of the unit vectors
designated Γ1 × Γ4 are, indeed, orthogonal.
Next, it is advantageous to discuss some more complicated symmetry
point groups, such as C3ν, which was introduced earlier:

This dimension (order) of this point group is h = 6, since there are six
symmetry elements in three classes, and some elements occur more
than once in this group. First, the matrix representation of the C3
operation (a clockwise rotation by an angle of θ = 120°) will be
introduced:

(11.19)

Equation (11.19) shows that for a rotation by 120°, the trace of the
transformation matrix is zero.
According to Eq. (11.11), C3v will have three irreducible
representations, but in order to fulfill Eq. (11.12), one of these, Γ3,
must be two‐dimensional:

(11.20)
since Eq. (11.12) requires that

(11.21)

The dimensionality of a representation was alluded to in Eq. (11.11).


If the largest number in an irreducible representation is two, this
representation is referred to as “doubly degenerate,” if it is three, the
representation is “triply degenerate.” The concept of two‐
dimensional representations, for example, can be visualized by a
coordinate system that rotates in space about the z‐axis. In this case,
the x–y plane would be a two‐dimensional representation. Similarly,
one can argue that a representation that has elements larger than
one represents two one‐dimensional directions that cannot be
separated.
In groups where the number of symmetry elements per class is larger
than one, such as the two C3 elements or the three σν planes in C3ν,
the orthogonality condition shown in Eq. (11.17) needs to be modified
to include a factor g that indicates the number of times an operation
occurs in a group. In C3ν, for example, this factor g would be two for
the C3 operations and three for the reflection planes. The
orthogonality condition then reads:

(11.22)

Equation (11.22) is the more general form of the orthogonality


condition that was defined previously (Eq. (11.17)). To demonstrate
the orthogonality of the representation listed in Eq. 11.20, one needs
to determine the products

(11.23)
and finds, indeed, that the three irreducible representations are
orthogonal. In Eq. (11.23), the bold numbers indicate the “g” factors,
and the following digits the values Xα(Γi) and Xα(Γj).
In general, one reserves the designations Γ1 × ΓN for molecular
representations (such as the symmetry properties of molecular
vibrations, or molecular orbitals; see below) and utilizes a different
nomenclature for the irreducible representations of a group. This
nomenclature is as follows: the totally symmetric representation of a
group, i.e., the irreducible representations whose X(Γ) values are all
+1, is generally given the designation Ax, where the subscript x can be
“g,” “1,” or “1g” (depending on whether or not a group includes a
center of inversion operation). In both C2ν and C3ν, the totally
symmetric representation is designated A1. The subscripts g and u
denote symmetric (gerade) and antisymmetric (ungerade) with
respect to a center of inversion.
There is always a totally symmetric representation in a group that is
always written as the first row in the character table. The irreducible
representation that transforms symmetrically (+1) with respect to
the highest axis of symmetry will be given the designation A2, or A2g.
Representations that transform antisymmetrically (−1) with respect
to the major axis of symmetry are given the symbol B. Doubly
degenerate (two‐dimensional) representations are referred to as E
and three‐dimensional representations as T. With these
designations, the irreducible representations (Eq. [11.14]) of the
symmetry group C2ν can be written as

(11.24)

The construct given by Eq. (11.24) is referred to as a “character


table.” Each symmetry group has its own characteristic character
table, some of which are listed in at the end of this chapter. Similarly,
the character table for C3ν is given by

(11.25)

Any representation Γ for a given symmetry group, such as the one


given in Eq. (11.10), can be decomposed into irreducible
representations just as a vector in three‐dimensional space can be
decomposed into its Cartesian components along the unit vectors.
This is accomplished by projecting the vector Γ onto the directions of
the irreducible component unit vectors to determine its component
along this direction. The representation derived in Eq. (11.10) for the
transformation properties of a Cartesian coordinate system under
the operations of the C2v point group will serve as an example of how
this decomposition is accomplished. The representation of the
transformation properties of a Cartesian coordinate system

(11.10)
henceforth will be referred to as ΓCCS, where the subscript CCS
stands for “Cartesian coordinate system.” The decomposition of ΓCCS
into contributions from the four unit vectors A1, A2, B1, and B2 is
carried out via the reduction (projection) formula:

(11.26)

In Eq. (11.26), ni indicates the number of times an irreducible


representation i occurs in the reducible representation, here ΓCCS.
This can be summarized in the following set of equations:
(11.27)

In these equations, the factor “g” is indicated in boldface, the X(Γi) in


gray typeface, and the X(ΓCCS) in regular typeface. The results
presented in Eq. (11.27) indicate that the “reducible representation”
ΓCCS can be decomposed into the following unit vectors:

(11.28)
One can easily show that 1 A1 + 1 B1 + 1 B2 yields the original
reducible representation:

The same procedure will be utilized, in Section 11.3, to derive


symmetry species of the individual vibrational modes of polyatomic
molecules. At this point, a few more comments about the character
tables introduced above are appropriate. In addition to the
information given in Eqs. (11.24) and (11.25), character tables of
molecular point groups contain two more columns, as shown below:

(11.29)

The first of these additional columns contains entries of the form Tα


and Rα, where α stands for x, y, or z, and T and R for translation and
rotation, respectively. Thus, this column indicates how a translation
or rotation of the coordinate would transform under the symmetry
operations of the group. It can be visualized easily that a translation
of a Cartesian coordinate system along the z‐direction transforms
symmetrically under all operations of the group. Often, the
translational components Tx, Ty, and Tz are just written as x, y, and z
in character tables. The reason for this will be appreciated later
(Section 11.4) since the Cartesian components of the dipole operator,
μx, μy, and μz, have the same transformation properties as do the
translations in the x‐, y‐, and z‐directions; thus, this column in the
character table determines whether or not a given representation is
allowed in any spectroscopy that depends on the change in the
molecular dipole moment or one of its components. A rotation of the
coordinate system about the z‐axis transforms as a C2 operation,
which transforms as A2 (see Eq. (11.9)).
The last column in the table denoted as Eq. (11.29) indicates the
transformation properties of binary combinations of Cartesian
coordinates. This column, therefore, determines whether or not a
given symmetry representation is allowed in Raman spectroscopy,
since in Raman spectroscopy, one or more of the polarizability tensor
components must change during a vibration and these tensor
components transform as the binary combinations of the Cartesian
coordinates.
11.3 Symmetry Representations of
Molecular Vibrations
As pointed out in the introduction to this chapter, the main goal of
the discussion of group theory is to determine which transitions in a
molecule are allowed in vibrational or electronic spectroscopy. This
will be demonstrated here for vibrational spectroscopy to determine
whether a normal coordinate's representations transform like x, y, or
z for IR spectroscopy or any binary combinations of x, y, or z for
Raman spectroscopy.
Thus, one needs to determine how a given normal coordinate
transforms under the symmetry operations of a group. In order to
accomplish this, one assumes that the normal modes of vibration can
be written as linear combinations of x, y, or z Cartesian displacement
coordinates. Thus, one attaches a Cartesian displacement
coordinate system to each atom and follows how each of these
coordinate components transforms under the symmetry operations
of the group. This is shown for the normal modes of the water
molecule (see Section 5.2) in Figure 11.6. As before, when deriving
the reducible representation of a coordinate system under the
symmetry operations of a group, one needs to establish the traces of
the transformation matrices for each of the atomic displacement
coordinate systems. Figure 11.6b shows the molecule after applying a
C2 operation. As before, one designates the new coordinates (after
the symmetry operation) as the primed coordinate and the original
coordinates as the unprimed ones, and one obtains the
transformation matrix for the C2 operation as shown in Equation
11.30:
(11.30)

Figure 11.6 (a) Cartesian displacement vectors for the water


molecule. (b) Cartesian displacement vectors after a C2 operation.
The trace of this transformation matrix is −1. Similarly, the trace of
the transformation matrix under the identity operation is 9, since
each Cartesian displacement vector component is unchanged under
the E.
From Eq. (11.30), it becomes obvious that the “molecular”
transformation matrix (i.e., the 9 × 9 element matrix 11.30) for the
water molecule contains three transformation sub‐matrices, one for
each individual Cartesian displacement coordinate system, and that
these transformation sub‐matrices are located at the diagonal of the
overall transformation matrix if the particular atom does not change
its position with another atom during the symmetry operation. If the
atoms change position during a symmetry operation, the sub‐
matrices will occur at off‐diagonal positions.
The σν′ (=σxz) operation exchanges the positions of atom 1 and 3, and
the transformation is

(11.31)

with a trace of 1.
For the σν (=σyz) operation, all atoms remain in place; therefore, all
displacement coordinate sub‐matrices appear on the diagonal, and
the trace of the overall transformation matrix is 3.
Thus, the (reducible) representation of the nine displacement
coordinates, from which the normal modes of vibration can be
constructed, is

(11.32)
Next, the reduction formula discussed (Eq. (11.26)) will be used to
determine the contributions of the four irreducible representations
to the reducible representation of the displacement coordinates for
water:

(11.26)

which reveals
(11.33)

Thus, the reducible representation ΓH2O = [9 −1 3 1] can be


decomposed into or reduced to

(11.34)
At this point, one should check that these contributions, indeed, add
up to nine degrees of freedom (3+1+3+2 = 9), since nine
displacement coordinates were attached to the water molecule.
Furthermore, one may wish to ascertain that the sum of the four
representations, multiplied by their abundance, reproduces the
original reducible representation:

This decomposition presented in Eq. (11.33) revealed nine degrees of


freedom, since three Cartesian degrees of freedom were assigned to
each atom (cf. Figure 11.6). However, since there are only 3N − 6 or
three degrees of vibrational freedom for the water molecule, the
other six degrees are three translational and three rotational degrees
of freedom. These translational and rotational modes originate from
certain combinations of displacement coordinates: for example, if all
atoms are displaced simultaneously in the z‐direction, the combined
motion is a translation of the entire molecule in the z‐direction.
Similarly, if atom 1 moves in the positive x‐direction and atom 3
moves in the negative x‐direction, the entire molecule rotates about
the z‐axis. As pointed out above, these translational and rotational
degrees of freedom are listed in the character table of the symmetry
group C2v (Eq. (11.29)) in the column containing the entries Tx, Ty,
Tz, Rx, Ry, and Rz. This table reveals that the translation of the water
molecule along the positive z‐axis transforms as the A1
representation, while the Tx and Ty translation transform as B1 and
B2, respectively. In order to obtain the symmetry properties of the
three vibrational modes of water, one subtracts the translational and
rotational representations from the total decomposition of the nine
displacement coordinates given by Eq. (11.36) and obtains

(11.35)

Thus, one ends up with two A1 and one B1 irreducible


representations for the symmetry species of the three normal modes
of water, which were depicted in Figure 5.1.
This procedure is generally applicable for any molecule. For
ammonia with C3v symmetry, for example, the reducible
representation of the Cartesian displacement coordinates is

(11.36)
This reducible representation can be derived easily using Eq. (11.19)
for the rotation matrix applied to each of the Cartesian displacement
coordinate sets. The trace of Eq. (11.19), with θ = 120°, is zero, and
the trace for each of the σν operation is +1. This trace enters into the
overall transformation matrix only for the atoms that do not change
position during the reflection operation.
Using the reduction formula Eq. (11.26) discussed above, the
reducible representation shown in Eq. (11.36) can be decomposed
into

(11.37)
Since each of the degenerate E representations accounts for 2°of
freedom, Eq. (11.37), indeed, represents 12° of freedom. Next, the
translational and rotational degrees of freedom will be subtracted.
These are available from the complete character table for C3ν:

(11.25)

Notice that the two‐dimensional representation E accounts for two


degrees of translation (Tx, Ty) and two degrees of rotation (Rx, Ry).
The vibrational degrees of freedom, therefore, are composed of 2 A1
and 2E irreducible representations, accounting for six (3N‐6 =
12−6 = 6) degrees of freedom. Examples for the analysis of more
complicated systems are given in the literature ([2], Section 2.6.1)

11.4 Symmetry‐Based Selection Rules for


Dipole‐Allowed Processes
Next, the symmetry rules for IR absorption based on the symmetry
species for given vibrational transitions will be discussed. As
discussed earlier, the vibrational energy of a molecule is determined
by the (time‐independent) vibrational Schrödinger equation that
yields the vibrational energy eigenvalues, which are independent of
symmetry, that is, a symmetry operation performed on the
vibrational wavefunction will not affect the energy. This last
statement, in mathematical terms, implies that the vibrational
Hamiltonian, , and any arbitrary symmetry operator, ,
commutate:

(11.38)

that is, the same eigenfunctions ψvib are solutions both for the
vibrational energy as well as symmetry operations.
The total vibrational wavefunction for a polyatomic molecule was
introduced earlier:

(5.13)

which may be written in an abbreviated form as

(11.39)

Here, denotes the ground‐state vibrational wavefunction of


normal coordinate Qi. If a transition along coordinate Qj is excited,
the excited‐state total wavefunction can be written as

(11.40)

For a transition described by

(11.41)
only the properties of the jth normal mode need to be considered
since all other modes remain unchanged:

(11.42)

For this coordinate, the transition moment

(11.43)

needs to be evaluated, which can be written in terms of the Cartesian


components of the transition operator μ:

(11.44)

Equation (11.44) implies that for a transition to occur, the integral


over the excited‐ and ground‐state wavefunctions and at least one
component the dipole operator must be nonzero. The ground‐state
wavefunction for any normal coordinate always transforms as the
totally symmetric representation of the symmetry group since the
ground‐state wavefunction of any normal coordinate is given by

(11.45)

which yields

(11.46)

Equation (11.47) describes a Gaussian distribution in the coordinate


Qj that is totally symmetric. The first excited state of the same
coordinate has the symmetry of the normal coordinate Qj
since the Hermite polynomial for the first excited state:
(11.47)

and thus has symmetry of the excited state that is determined by the
irreducible representation of the particular normal mode of
vibration. Integrals of the form ∫ψ1j(Qj) μz ψ0j(Qj) dz will henceforth
be abbreviated as ∫ f1 f2 f3 dτ, where f3 denotes the ground‐state
wavefunction that always transforms as the totally symmetric
representation (see Eq. (11.46)). f1 and f2 represent the excited‐state
wavefunction and transition operator, respectively. Thus, for the
total integral ∫ f1 f2 f3 dτ to be nonzero or even, the product f1 f2 must
be even as well. This is the case if f1, the excited‐state wavefunction,
transforms as one of the dipole moment components of the group.
Then, its product with f2 contains the totally symmetric
representation of the group. The transformation properties of the
components of the dipole operator were discussion before (see Eq.
(11.29)).
To illustrate the points in the last paragraph, the example of the
representations of the vibrations of the water molecule will be
discussed. Equation (11.37) demonstrates that the vibrations of the
water molecule belong to the irreducible representations A1 + B1: the
symmetric stretching mode and the deformation mode transform
(see Figure 5.1) as A1 and the antisymmetric stretching mode as B1.
The question now arises which of these modes is allowed in IR
absorption. According to the character table for C2v, reproduced
below,

(11.29)
the A1 irreducible representation transforms like a translation along
the z‐direction. Thus, the dipole operator component μz also
transforms as A1. For either of the transitions of A1 symmetry (i.e.,
the deformation and the symmetric stretching mode), the ground
vibrational state is totally symmetric and transforms as A1. The
excited vibrational state for either of these modes also transforms as
A1, as pointed out before. Since the transition moment μz also
transforms as A1, the product of the excited‐state and the dipole
operator component (both A1) certainly contains the totally
symmetric representation, and both transitions are allowed in
absorption.
For the 1 ← 0 transition of the antisymmetric stretching mode of B1
symmetry, one proceeds as follows. Again, the ground‐state
vibrational mode transforms as A1 (see Eq. (11.46)). The excited state
transforms as B1, as does one of the components of the dipole
operator, μx. Thus, the product f1 f2 transforms as B1B1, and it is easy
to see that this product contains the totally symmetric representation
(or transforms as the totally symmetric representation) of the group,
and the antisymmetric stretching vibration of B1 symmetry is allowed
as well.
In summary, the discussion in Section 11.4 demonstrated that any
vibrational transition will be allowed in absorption if its irreducible
representation contains a component of the electric dipole moment,
μx, μy, or μz, (or the translational directions Tx, Ty, or Tz).

11.5 Selection Rules for Raman Scattering


The principles discussed in the previous section can be applied to
Raman scattering as well, but the Cartesian components of the dipole
transition moment in Eq. (11.46) need to be replaced by the
polarizability elements, since Raman spectra arise from the changes
in polarizability α during a normal coordinate (see Eq. [5.46]). Thus,
whereas IR transitions require
(5.44)

the corresponding condition for a transition in Raman spectroscopy


is

(5.46)

As elaborated upon in Chapters 4 and 5, a direct transition from a


lower to a higher vibrational state requires absorption of exactly one
photon; this absorption process is mediated by the transition dipole
operator. Raman spectroscopy, on the other hand, is a process that
involves the interaction of two photons. Thus, the transition moment
is determined by expressions of the form

(11.48)

where ψ0 and ψ1 are the vibrational ground and excited state,


respectively, of a given normal coordinate Qk and ψint is a virtual
(vibronic) intermediate state (see Figure 5.7). Since two transition
moments are formally involved in the transition process, different
selection rules apply for Raman spectroscopy. The irreducible
representations that support Raman transitions, therefore, have
binary combinations of the Cartesian coordinates listed, e.g. x2, y2,
z2, xy, xz, yz, or others. The character table for the symmetry group
C2v, shown in Eq. (11.29), has these binary combinations listed in a
column to the right of the column containing the translational and
rotational components. In complete analogy to the selection rules for
absorption, a vibration transforming as a given irreducible
representation will be allowed in Raman scattering if at least one of
the binary combinations of the Cartesian coordinates is listed.
Inspection of the character tables below reveals that symmetry
groups that contain the inversion symmetry element, i, never have
irreducible representations that simultaneously contain elements of
the dipole operator (i.e. x, y, and z) and any binary combinations of
these coordinates. This observation leads to what is known as the
“mutual exclusion rule” that states that in a point group that contains
a center of inversion, i, vibrational transitions cannot be
simultaneously active in Raman scattering and IR absorption. Thus,
in molecules such as CO2 or octahedral SF6, vibrations that are
Raman allowed are IR forbidden, and vice versa. Isomeric molecules
such as cis‐dichloroethene (C2ν) and trans‐dichloroethene (C2h) can
be identified easily by the fact that for the latter molecule, the mutual
exclusion principle holds.

11.6 Character Tables of a Few Common


Point Groups
References
1 Cotton, F.A. (1990). Chemical Applications of Group Theory, 3e.
New York: Wiley.

2 Diem, M. (2015). Modern Vibrational Spectroscopy and Micro‐


Spectroscopy: Theory, Instrumentation and Biomedical
Applications. Chichester, UK: Wiley.

Problems
1. For the C4ν point group for which the character table is given
below, demonstrate the orthogonality of the irreducible
representation, A1, A2, B1, B2, and E:

2. For the C2h point group:


a. Write the 3 × 3 transformation matrices for a Cartesian
coordinate system for each of the four symmetry
operations.
b. Show by matrix multiplication that C2 ⊗ C2 = E.
c. Show by matrix multiplication that C2 ⊗ σh = I.
d. Show that Au and Bu are orthogonal.
3. Generate a 3 × 3 matrix representation for the following
operators: (a) C6, (b) S4, (c) i.
4.
a. For the C3v point group, define the 3 x 3 transformation
matrices that describe the E and any of the three σν
operations for a Cartesian coordinate system.
b. Determine by matrix multiplication the symmetry operation
that results from consecutive application of two σν
operations.
c. Show that the A2 and E irreducible representations are
orthogonal.
5. Consider the planar molecule trans‐diazene, trans‐H–N=N–H.
a. What is the symmetry group to which this molecule
belongs?
b. How many vibrational normal modes do this molecule
exhibit?
c. Between 3000 and 3300 cm−1, this molecule exhibits two
vibrations, one in Raman, the other in IR spectroscopy.
Describe these two vibrations in terms of their atomic
displacements and symmetry species (irreducible
representation).
6. For the Td point group with the following symmetry elements: E
8C3 3C2 6S4 6σd
a. Decompose the reducible representation 15 0 −1 −1 3 into
the contributions of the irreducible representations.
b. Ascertain that these representations account for 15° of
freedom and that the contributions of these irreducible
representations produce the reducible representation.
c. Which are the translational and rotational degrees of
freedom?
Appendix 1
Constants and Conversion Factors
Avogadro's constant NA = 6.022 × 1023 [mol−1]
Boltzmann's constant k = 1.381 × 10−23 [J/K]
Electron charge e = 1.602 × 10−19 [C]
Electron mass me = 9.109 × 10−31 [kg]
Gas constant R = 8.314 [J K−1mol−1]
= 0.698 [cm−1 K−1 molecule−1]

Nuclear magneton

=5.0508 × 10−27 [J/T] = [A m2]


Permittivity ε0 = 8.854 × 10−12 (F/m = C2/[J m])
Planck's constant h = 6.626 × 10−34 [Js]
ħ = 1.054 × 10−34 [Js]
Proton mass mp = 1.673 × 10−27 [kg]
Rydberg constant Ry = 2.179 × 10−18 [J]
Velocity of light c = 2.998 × 108 [m/s]
Energy Conversion

1. [eV] = 8065.5 [cm−1] = 1.602 × 10−19 [J]

Since
Appendix 2
Approximative Methods: Variation and
Perturbation Theory

A2.1 General Remarks


Aside from the simplest quantum mechanical systems – the particle
in a box, the rigid rotor, the harmonic oscillator and simple one‐
electron systems such as the H, He+, Li2+, etc. atoms and ions – most
of the quantum mechanical problems cannot be solved analytically
since the differential equations describing these systems cannot be
solved explicitly. Thus, approximate methods must be used to solve
any problems of higher complexity. At first sight, this mathematical
limitation seems to seriously affect the usefulness of quantum
mechanics. However, there exist methods to solve the differential
equations by approximation methods, which are the subject of this
appendix.
Approximate methods can be applied since one of the postulates
(Postulate 5) of quantum mechanics (see Chapter 2) stated that the
real eigenfunctions φ(x) of an operator form a complete
vector space. Functions that are not eigenfunctions of can be
written in terms of a series expansion of the true eigenfunctions φ(x)
as follows:

(2.10)

where the φ(x)'s are the true eigenfunctions, and the expansion
coefficients an indicate how much each wavefunction looks like the
true eigenfunction of the operator. Typical examples of these
approximate methods were encountered in the discussions of earlier
chapters, for example the orbital approximation described in
Chapter 7: since the multi‐electron problem encountered in the He
atom cannot be solved, due to the electron–electron correlation, we
assumed that the true wavefunctions of the He atom may be
approximated by a linear combination of the individual hydrogen‐
like wavefunctions, approximating the electron–electron correlation
by the effective charge. Subsequently, the expansion coefficients in
Eq. 2.10 above need to be established that produce the best
agreement with experimental results. This can be accomplished in
two ways via variational or perturbational methods.

A2.2 Variation Method


Let us assume we have a system that obeys the Schrödinger equation

(A2.1)

where the Hamiltonian is so complicated that the explicit solution


of Eq. (A2.1) does not exist. However, there exists a somewhat
simpler Hamiltonian that provides eigenfunctions φ that form a
complete vector space. Hence, we write the solutions ψ(x) as a series
expansion in a set of trial functions φ(x) as given by Eq. (2.10). The
variation method establishes that the energy expectation value of the
operator in the space of the trial functions, φ is always larger (or
equal at best) than the true energy. Thus, one needs to show that:

(A2.2)

where

(A2.3)

that is, that the approximate functions have higher energy than the
true lowest energy eigenvalue of the original problem (notice that
this variational formalism works for ground states only).
Thus, we need to establish that
(A2.4)

(A2.5)

Here, ∫ψ*ψ dτ = 1 because the functions ψ are normalized. Next, we


substitute the series expansion given by Eq. 2.10 into Eq. A2.4:

(A2.6)

(A2.7)

(A2.8)

(A2.9)

The last simplification was again due to orthonormality of the


eigenfunctions φk. It follows that Eq. (A2.4),
can be written as

(A2.10)

Since the ak in Eq. (2.10) is positive, and since (Ek − E0) is positive as
well (since E0 is the ground state), we have shown that by
substitution of trial function that obey the same boundary
conditions, we can be assured that the approximate energy for the
ground state is never below the real energy value. Then we may vary
the expansion coefficients in the series expansion

until the observed energy is a minimum closest to the


(experimentally observed) energy. For the application in this book,
the variational method has less importance than, for example, for
molecular orbital calculations since it can only be applied to ground
state energy systems. In spectroscopy, by nature of the transitions
involved, we need to consider at least one excited state. Thus, the
perturbation method has much higher importance here and will be
discussed in more detail.

A2.3 Time‐independent Perturbation Theory


for Nondegenerate Systems
Again, let us assume we have a system that obeys the Schrödinger
equation

(A2.11)

Furthermore, let us assume that the Hamiltonian is so


complicated that explicit solution of Eq. (A2.11) is impossible, but
that there is a similar Hamiltonian for which exist explicit
solutions

(A2.12)

At this point, one defines a “perturbation Hamiltonian” such that

(A2.13)
where λ is a scaling parameter so that one may gradually and
incrementally apply the perturbation. An example would be the (one‐
dimensional) anharmonic oscillator Schrödinger equation:

(A2.14)

where f and g are anharmonicity constants.


It is a reasonable assumption that both the perturbed wavefunctions
ψn and the perturbed eigenvalues En in Eqs. (A2.11) and (A2.22)
depend on how much perturbation is applied:

(A2.15)
Thus, one expands the wavefunctions and eigenvalues in a power
series in λ; the first‐order perturbations to the wavefunctions and
energies are

(A2.16)

and

(A2.17)

where the first‐order perturbed energies are given by

(A2.18)

or
(A2.19)

Equation (A2.19) implies that the first‐order correction to the energy


is found by calculating the expectation value of the perturbation
Hamiltonian using the unperturbed wavefunctions. It can be shown
([1], Section 9.2) that the perturbed wavefunctions are:

(A2.20)

For example, the first‐order perturbed wavefunction ψ1 would be

(A2.21)

Assuming that the perturbed energies get smaller, as compared


to , we may truncate the summation in Eq. (A2.21) after a few
terms, as demonstrated in the next section.

A2.4 Detailed Example of Time‐independent


Perturbation: The Particle in a Box with a
Sloped Potential Function
In the following section, an example of a typical perturbation
problem is presented. This example is somewhat lengthy and best
presented as its own section. The example to be discussed was
mentioned before in Chapter 2, Figure 2.8, and is based in the
quantum cascade laser that may be described as a particle in a box
(PiB) with finite energy barriers and a sloped baseline, as shown in
Figure 2.8. The sloped baseline in a quantum well can be achieved by
suitable doping of the semiconductor materials.
For this example, it is assumed that the potential energy barrier is
still infinitely high, but that inside the energy well of length L = 4, the
potential energy is given by

(A2.22)

that is, a straight line, as shown in the sketch at left.


Thus, the potential energy for this particular problem is

(A2.23)

The first perturbed energy level is given by (see Eq. [A2.19]):


(A2.24)

where the are the unperturbed PiB wavefunctions. For a box


with length 4, these are (cf. Eq. [2.38]):

(A2.25)

Thus, within the box,

(A2.26)

(A2.27)

These integrals can be solved by hand, or in this case here (which


was a take‐home exam question in one of the authors classes) by
using the Mathematica software by typing
Integrate[Sin[Pi x/4]*Sin[Pi x/4], [2]]
and
Integrate[x*Sin[Pi x/4]*Sin[Pi x/4], [2]]
and obtaining the values 2 and 4, respectively. Thus, and
.

Rather than integrating Eq. A2.26 in parts,


Integrate[(2‐(x/2))* Sin[Pi x /4]* Sin[Pi x/4], [2]] will give
the same result. Higher‐order perturbed energies are obtained in an
equivalent manner:
(A2.28)

by evaluating
Integrate[(2‐(x/2))* Sin[2Pi x /4]* Sin[2Pi x/4], [2]] for
and
Integrate[(2‐(x/2))* Sin[3Pi x/4]*Sin[3Pi x/4], [2]] for .

In all cases, the perturbation integral has the value 1; thus, the
perturbed energy levels are just raised by one unit as compared to
the PiB with a flat bottom.
Next, the perturbed wavefunctions will be evaluated according to
Eqs. (A2.20) and (A2.21):

(A2.20)

(A2.21)

To evaluate these, we need to calculate integrals of the form:

(A2.22)

We again use Mathematica to solve, for example, the integral

Integrate[(2‐(x/2))* Sin[Pi x /4]* Sin[2Pi x/4], [2]];

(A2.23)
The following summarizes these results. For the first‐order perturbed
ground state wavefunction ψ1, we find:

(A2.24)

(A2.25)

(A2.26)

(A2.27)
Thus,

(A2.28)

(A2.29)

The first‐order perturbed first excited wavefunction ψ2 is:

(A2.30)

(A2.31)
(A2.32)

(A2.33)

(A2.34)

The first‐order perturbed second excited wavefunction ψ3 is:

(A2.35)

(A2.36)

(A2.37)

(A2.38)

(A2.39)

The gist of this example is that we have expanded the perturbed


wavefunction in a series expansion in the unperturbed
wavefunctions, scaled by the energy differences between the
perturbed eigenvalues. The results of these calculations are
presented in Figure A2.1. The left panel shows the standard PiB
energy eigenvalues and (squared) wavefunctions. Panel (B) shows
the PiB with the sloped potential energy barrier and the resulting
energy eigenvalues. Notice that the ground state E1 is no longer at
energy value 1, as in the unperturbed PiB, but at E1 = 2. (All energies
given in units of [h2/8πmL2], see Eq. [2.33]). That is, the electron in
the ground state avoids the sloped energy barrier by moving up one
notch. This is true for all energy eigenvalues that are increased by
one unit. Notice furthermore that the maxima of the wavefunction is
shifted to the right to avoid the potential energy perturbation. This is
particularly obvious for the n = 2 and n = 3 states where the
probability of finding the electron shifts toward the right.
Figure A2.1 Comparison between the unperturbed (a) and
perturbed (b) particle‐in‐a‐box eigenvalues and wavefunctions. The
squared wavefunctions are shown. See text for details
This example demonstrates the inner workings of perturbation
theory, and the results make intuitive sense since the electron tends
to avoid the perturbed side of the energy well and escapes to the
right. For the quantum‐cascade laser discussed in Chapter 2, this
implies that electron has a higher probability density at the side of
the well with the lower potential energy, and consequently, has a
better chance of tunneling through the (right) barrier.

A2.5 Time‐dependent Perturbation of


Molecular Systems by Electromagnetic
Radiation
Time‐dependent perturbation theory was introduced in Chapter 3 to
define the response of a quantum mechanical system to a
perturbation that acts for a limited time, and in most cases, is due to
electromagnetic radiation impinging on a sample. Here, we assume
that the system is originally in a stationary state, often the ground
state, and responds by making a transition to an excited state if
certain conditions are fulfilled.
Mathematically, one describes the time dependent perturbations as
follows. The quantum mechanical system, in the absence of a
perturbation, is described by the time independent Schrödinger
equation

(A2.40)

where E0 and ψ0 are the stationary state and energies and


wavefunction of the system. When the time‐dependent perturbation
is applied, the system needs to be described by the time dependent
Schrödinger equation
(A2.41)

Here, Ψ(x, t) is a wavefunction that depends both and spatial and


temporal coordinates:

(A2.42)

and again is the perturbation operator. In the absence of the


perturbation, the solutions of the time dependent Schrödinger
equation

(A2.43)

are

(A2.44)

The functions are the eigenfunctions given in Eq. (A2.40). When


the perturbation is present, the functions given in Eq. (A2.44) are no
longer the eigenfunctions of Eq. (A2.41) but need to be replaced by
an expansion

(A2.45)

(A2.46)

where the coefficients ck themselves are time‐dependent to account


for the changes the system will undergo as a response to the
perturbation.
The wavefunctions given in Eq. A2.46 then are substituted into Eq.
A2.41 and the dependence of the coefficients ck on time is obtained
(see [1], p 227). When the perturbation is explicitly written as being
due to electromagnetic radiation (see Chapter 3), the perturbation
operator can be written as

(3.2)

The expansion coefficients are integrated from time t = 0 when the


perturbation starts to a later time t = t'. The time‐dependent
expansion coefficients are obtained (after an additional number of
steps described in the literature [1]) to be

(3.14)

This equation was previously described as one of the fundamental


equations governing the interaction between electromagnetic
radiation and matter. It contains the conditions that a) the electric
field must be present, the transition moment 〈ψn ∣ μ ∣ ψm〉 must be
nonzero, and the energy matching condition between the photon and
the molecular system must be fulfilled.

Reference
1 Levine, I. (1983). Quantum Chemistry. Boston: Allyn & Bacon.
Appendix 3
Nonlinear Spectroscopic Techniques
The advent of tunable, high power, and ultrashort pulse lasers has
spawned the development of new optical techniques that have truly
revolutionized the field of spectroscopy. Perhaps with the exception
of FT–NMR techniques, using various pulse sequences to create
multidimensional NMR processes, no other spectroscopic method
has experienced such an explosive expansion during the past 30
years as has nonlinear spectroscopy.
The term “nonlinear” implies a dependence of the induced effects on
the square or cube of the laser intensity. The first of the nonlinear
Raman effects – the hyper‐Raman effect, a noncoherent three‐
photon effect, and coherent anti‐Stokes Raman scattering – were
carried out at the Ford Motor Company Research laboratory in 1965
using a ruby laser for excitation [1]. This scientific achievement
occurred a mere five years after the first experimental verification of
a visible laser in 1960 (which was, incidentally, also a ruby laser). In
the present times of financial hardship for any scientific endeavor, it
may come as a surprise that this research was performed in the
laboratory of a private company. Subsequently, several other
nonlinear phenomena have been reported, some of which will be
introduced in this chapter.

A3.1 General Formulation of Nonlinear


Effects
With the advent of high‐power pulsed lasers, several novel
spectroscopic effects were discovered, where the size of the effects
depend nonlinearly on the strength of the electric field of the exciting
electromagnetic radiation, and therefore, referred to as “nonlinear
spectroscopies.” The electric field in a pulsed laser can exceed
1010 V/m, which is about 100,000 times stronger than the field
strength of a common continuous wave laser. At these high laser
fields, the induced dipole moment is no longer represented by Eq.
(5.31)

(5.31)

but needs to be rewritten as a series expansion to included higher‐


order contributions to the induced electric dipole:

(3.1)

where is known as the first hyperpolarizability tensor of rank 3 (a


3 × 3 × 3 matrix). Equation (A3.1) is often written in terms of
macroscopic dielectric susceptibilities:

(3.2)

where χ(n) is a tensor of rank n + 1, known as the dielectric


susceptibility, which relates the induced electric macroscopic
polarization P to the electric field E of the exciting radiation. In Eqs.
(A3.1) and (A3.2), the first terms on the right‐hand side, describe the
polarizability and macroscopic polarization, respectively, which are
responsible for the effects discussed previously in Chapter 5. The
second term in Eq. (A3.2) is responsible for hyper Raman scattering,
whereas the second and third terms in Eq. (A5.2) account for
nonlinear effects such as frequency‐doubling and nonlinear Raman
effects such as coherent anti‐Stokes Raman scattering (CARS).

A3.2 Noncoherent Nonlinear Effects:


Hyper‐Raman Spectroscopy
Hyper‐Raman scattering results from the second term in Eq. (A3.2),
and is a nonlinear, noncoherent form of nonlinear Raman
spectroscopy. Hyper‐Raman and hyper‐Rayleigh spectroscopies are
three‐photon processes depicted schematically in Figure A3.1a. Two
photons of frequency ω1 (up arrows) create two virtual states, shown
by the dashed lines. A hyper‐Rayleigh photon at frequency 2ω1, or a
hyper‐Raman photon at frequency 2ω1 – ωm is created from the
upper virtual state, where ωm is the frequency of a molecular
vibrational quantum. This is shown by the downward arrows in
Figure A3.1b. The two laser photons need not necessarily have the
same frequency: the hyperpolarizability can also mix photons of
different frequencies, ω1 and ω2, resulting in what has been referred
to as “nondegenerate” hyper‐Rayleigh and hyper‐Raman scattering
[2] with frequencies ω1 + ω2 and ω1 + ω2 − ωm, respectively. This is
shown in Figure A3.1b. Just as the hyper‐Rayleigh effect can be
viewed as the incoherent form of frequency doubling, the
nondegenerate hyper‐Rayleigh effect is the noncoherent analog of
sum‐frequency generation (see below).
In analogy to the Raman scattering tensor,

(5.48)

the hyperpolarizability for initial and final and intermediate states i,


f, r, and s can be defined in terms of the transition moments Mfr,
Mrs and Msi (in the far‐from resonance approximation) as
(A3.3)
Figure A3.1 Schematic energy level diagram for degenerate (a and
nondegenerate (b) hyper Rayleigh and hyper‐Raman scattering. See
text for details.
In Eq. (A3.3), the subscripts f and i denote the final and initial states,
and s and r the two intermediate virtual states. Comparison of Eqs.
(5.48) and (A3.3) suggest that the selection rules should be different
for Raman and hyper‐Raman spectroscopy since two or three
photons are involved in these processes, respectively. In fact, low
symmetry vibrations often are allowed, and more intense, in hyper‐
Raman spectroscopy, whereas totally symmetric vibrations may be
very weak or entirely forbidden.
Since the selection rules are all different for Raman, hyper‐Raman,
and infrared spectroscopy, the hyper‐Raman effect complements the
two other forms of vibrational techniques in the sense that it allows
the observation of vibrational modes that cannot be observed in
Raman or infrared spectroscopy. The torsional vibration in
tetrachloroethene, C2Cl4, for example, which transforms as the Au
irreducible representation of the D2h point group, is not active in
either infrared absorption or Raman scattering, but is observed in
hyper‐Raman spectroscopy at 110 cm−1.

As pointed out above, the hyper‐Rayleigh effect is the noncoherent


analog of the second harmonic generation, also known as frequency
doubling that will be introduced in the next section.

A3.3 Coherent Nonlinear Effects


There is an important distinction between the hyper‐Raman effect,
the first form of nonlinear Raman spectroscopy discussed here, and
the coherent forms of nonlinear spectroscopy, such as the stimulated
Raman, inverse Raman, Raman gain, and coherent anti‐Stokes‐
Raman effects.
In the former technique, the hyper‐Raman scattered photons are
scattered noncoherently (spontaneously) into all 4π steradians, just
as in the case of spontaneous Raman scattering. Also, in both Raman
and hyper‐Raman effects, all normal modes are excited
simultaneously when a molecule is exposed to the exciting laser
radiation. In the coherent, nonlinear Raman effects, on the other
hand, the scattered light exits the sample as a coherent beam with
the properties of laser light. Since some of the nonlinear Raman
techniques require two input laser beams, only one normal mode,
determined by the frequency difference between the two laser beams,
is excited, leading to a much higher scattering efficiency. Thus, some
of the nonlinear effects are very strong (albeit still not easily
observed), whereas the incoherent hyper‐Raman effect is so weak
that it is still not a particularly practical technique.
In the remainder of this section, some of the theory underlying the
coherent nonlinear techniques will be introduced. The equation most
commonly used as a starting point for the discussion of nonlinear
effects is Eq. (A3.2).

A3.3.1 Second Harmonic Generation


A very common nonlinear optical effect will be discussed first,
namely coherent frequency doubling. This process, which is also
known as second harmonic generation (SHG), is mediated by the
second‐order dielectric susceptibility χ(2). Two photons of angular
frequency ω are combined in a non‐centrosymmetric crystal into one
photon with angular frequency 2ω. Since the molecular system
involved in this frequency‐doubling process is left in the original
state after the new photon is created, the momenta and energies of
the photons involved must be conserved. The momentum
conservation is indicated by writing the electromagnetic fields as

(A3.4)

where k is the wave (or momentum) vector of the photon, defined by

(A3.5)
and

(A3.6)
The conservation of momenta requires that k3, the momentum of the
frequency‐doubled photon, is given by

(A3.7)
where k1 and k1 are the momenta of the original photons. Eq. (A3.7)
is often referred to as the phase‐matching condition in nonlinear
optics. Figure A3.2 depicts the phase‐matching condition for
frequency doubling. Here, Eq. (A3.7) and Figure A3.2a predict that
the frequency‐doubled photon emerges from the (nonlinear) crystal
material collinearly with the incident beam.
Typical nonlinear materials used for second harmonic generation are
LiIO3, KNbO3, LiNbO3, KH2PO4 (KHP), KD2PO4 (KDP),
LiB3O5·(LBO), β‐BaB2O4 (BBO), GaSe, KTiOPO4 (KTP), and
(NH4)H2PO4 (ADP), where the commonly used (engineering)
abbreviations, given in parentheses, strike horror into the hearts of
chemists. Frequency doubling has become such a commonplace
technique that green laser pointers, available for under $40, contain
a frequency‐doubled diode‐pumped solid‐state laser, which gives a
nice green (532 nm) spot on a reflective screen.

Figure A3.2 (a) Phase matching diagram for frequency doubling


(SHG). (b) Phase matching diagram for sum‐frequency generation
(SFG). (c) Directions of the incident beams 1 and 2 to create beam 3
in SFG.
When the photons incident on the nonlinear crystals do not have the
same frequency, they still can interact and combine to form a new
photon according to

(A3.8)
and

(A3.9)
This process is known as sum‐frequency generation (SFG); second
harmonic generation or frequency doubling can be considered a
special case of SFG. Since the vectors k1 and k2 have different
lengths, the phase‐matching condition appears, as shown in
Figure A3.2b, with the direction of the incident and emitted photons
given by the arrows. This is shown in Figure A3.2c that depicts that
the incident beams 1 and 2 have to intersect at the angle derived
from the vector addition shown in Figure A3.2b to fulfill the phase‐
matching criterion.
Since the direction of the incident and emitted photons are different,
and the refractive indices within the nonlinear crystal may differ
along the different directions, Eq. (A3.8) should be written as

(A3.10)
To account for the phase matching conditions of nonlinear processes,
the wave vector notation of the incident and emitted fields is
included in the equations for the induced polarization. For sum‐
frequency generation, for example, the second‐order susceptibility
term is written to include the phase‐matching condition as

(A3.11)

Here, the sign convention in the exponent indicates an emitted


photon at ω3 and incident photons at ω1 and ω2.

A3.3.2 Coherent Anti‐Stokes Raman Scattering (CARS)


For the further discussion of nonlinear Raman effects, the following
conventions will be used. The medium is exposed to various
electromagnetic fields at frequencies ωa, traveling in the z‐direction.
Such a field is represented by

(A3.12)
where the subscript a denotes any of the individual radiation fields.
In CARS, which is probably the most commonly used nonlinear
Raman technique, there are three such radiation fields incident on
the molecule to create a fourth photon, the CARS photon, according
to the energy level diagram shown in Figure A3.3. In the following
paragraphs, the four events necessary for the creation of a coherent
anti‐Stokes photon are described as if they occurred consecutively. In
reality, the four processes do not occur as separate events but are a
four‐wave mixing phenomenon mediated by the third‐order
nonlinear susceptibility χαβγδ (see below).
In CARS, the sample is illuminated by two lasers, one of them with a
fixed wavelength, usually referred to as the pump laser ωP or ω1, and
a second tunable laser referred to as the Stokes frequency ωS or ω2. A
photon ħωP at the pump frequency promotes the system into a
virtual state, shown by the lower dashed line in Figure A3.3a. A
photon ħωS from the laser at the (Stokes) Raman frequency causes
the system to populate the vibrationally excited state, shown by the
upper solid line. The vibrationally excited state in turn interacts with
a second pump photon, ħωP, to populate another virtual state that
undergoes a transition back to the ground state. The energy released
in this last step is carried off by a photon of frequency

(A3.13)
where ħωM is energy of one of the molecule's vibrational modes.
Thus, the wavelength of the emitted photon is that of an anti‐Stokes
Raman process, and the emission of the anti‐Stokes photon occurs
only if the wavelength of the tunable Stokes laser fulfills the
condition

(A3.14)
For this process, in analogy to the discussion of SHG (Eq. [A3.11]),
the term responsible for the CARS process can be written as:
(A3.15)

Figure A3.3 (a) Schematic energy level diagram for the CARS
process. (b) Phase matching condition. θ denotes the angle between
the pump and Stokes beams. θ′ is the angle between the pump and
the CARS beam that is emitted along the dotted line.
where the sign associated with each term in the exponential indicates
whether a photon is annihilated or created. This exponential
expression contains the four wave vectors of the interacting
electromagnetic fields. Since the molecular system is left in the
original state after the creation of the CARS photon, the wave vectors
need to add up to zero. This leads to the phase‐matching condition
for CARS, which can be written as

(A3.16)

and visualized in Figure A3.3b. It implies that the pump and Stokes
beams must intersect at an angle given by the vector addition in
Figure A3.3b for CARS photons to be generated.
The CARS intensity scattered at ωAS is given by

(A3.17)
where all symbols have their usual meaning and z is the distance
over which phase matching is valid.
At this point, it is appropriate to investigate the form of the third‐
order susceptibility tensor of rank 4, used in Eq. (A3.15). This tensor
has 81 (=34) elements, of which only 21 are nonzero in isotropic
media. In fact, only tensor elements for which all four indices are the
same (e.g., χxxxx) and those for which there are two pairs of identical
indices (e.g., χxxyy, χxyyx or χxyxy) are nonzero.
The 21 nonzero elements exhibit only four different numeric values,
commonly referred to as χ1111, χ1122, χ1221, or χ1212. With that, Eq.
(A3.15) can be rewritten as

(A3.18)

where D is an integer factor between 1 and 6 that indicates how often


each susceptibility term must be counted [3]. This equation assumes
that the fields are polarized along the x‐axis.

The anti‐Stokes photons leave the sample as a collimated, coherent


laser beam, for which the spectral resolution is given by the line
width of the exciting lasers. As pointed out before, spectral
information on only one normal mode at a time is obtained if the
Stokes laser is scanned to cover the spectral range.
CARS spectroscopy has taken a huge step forward when it was
demonstrated in microscopic measurements [4]. In the large solid
angle of the light cone in a microscope objective, the phase‐matching
angle is always fulfilled [5]. In an elegant optical arrangement
reported by Kano [6], femtosecond laser pulses from a Ti:Sapphire
laser were split to obtain pulses that were used as CARS pump pulses
after filtering them to ca. 20 cm−1 bandwidth. The other part of the
split beam was directed to a photonic crystal fiber to create a
coherent super‐continuum pulse for Stokes excitation. In this way,
CARS spectra of 2000 cm−1 width could be collected simultaneously
in a microscope set‐up (cf. Figure A3.4).
Figure A3.4 Broadband micro‐CARS spectra of cellular
components: (a) nucleolus, (b) chromosome, (c) cell membrane, and
(d) background (from ref. 6).
A3.3.3 Stimulated Raman Scattering (SRS) and
Femtosecond Stimulated Raman Scattering (FSRS)
Another nonlinear Raman effect due to the third‐order susceptibility
is stimulated Raman spectroscopy. In stimulated Raman scattering,
the sample is illuminated with only one laser at the frequency ωP. If
the intensity of this pump laser increases past a threshold level, the
intensity of the Stokes–Raman scattering becomes sufficiently large
that nonlinear mixing of the Raman radiation field with that of the
pump laser causes coherent laser output at ωS to occur, where ħωS =
ħ(ωP − ωM) (see Eq. [A3.14]). This effect can convert up to 50% of
the incident photons into Raman scattering, compared to an
efficiency of spontaneous Raman spectroscopy on the order of 10−10–
10−12.
SRS also has taken on a new direction with the development of
femtosecond stimulated Raman spectroscopy, referred to as FSRS,
which is a logical extension of SRS. Here, broadband femtosecond
pulses are mixed coherently with a narrow pump laser frequency,
and the same frequency mixing described before for SRS takes place.
The major difference in FSRS is that all frequencies contained in the
broadband pulse simultaneously can mix with the pump pulse; thus,
the entire Raman spectrum can be probed at once. This is shown
schematically in Figure A3.5, taken from [7]. This Figure depicts the
narrow Raman pump pulse and the broadband femtosecond probe
pulse. Typically, this pulse is about 20 fs long and has a natural
linewidth of about 1600 cm−1. In the presence of a sample that
exhibits allowed Raman transitions, some of the pump photons are
transferred into the probe beam at the frequencies of the Raman
modes. Ratioing the probe beam profile collected with and without
the pump pulse yields the desired Raman spectrum, shown as the
trace on top of Figure A3.5. This spectrum corresponds to a single
laser pulse data acquisition (ca. 20 fs).
Figure A3.5 Schematic diagram of FSRS. See text for details
(from ref. 7).
This time resolution is about the same as the low wavenumber
vibrational frequency (a vibration at 500 cm−1 has a vibrational
period of 60 fs); thus, FSRS can probe the time evolution of a
vibrational mode and vibrational dephasing. This time scale is of
significant interest since the corresponding low‐frequency vibrations
sample nuclear motion along the reaction coordinates, which may be
described in terms of the normal modes of the molecular systems
during chemical reactions. Thus, FSRS offers very fast access to
assess molecular reaction dynamics, and an entire branch of Raman
spectroscopy has evolved around these concepts, and interested
readers are referred to review articles by the Mathies group [7]. A
number of items from these reviews are summarized below: first, the
FSRS signal is phase‐matched according to
(A3.19)

which implies that the FSRS photons are emitted collinearly with the
probe pulse. Second, the FSRS appears to violate the Heisenberg
uncertainty principle in that the time‐frequency product Δν Δt is
about an order of magnitude better in FSRS than expected from the
uncertainty principle that predicts [7]

(A3.20)
This may be understood in terms of disentanglement of energy and
time resolution because the broadband femtosecond pulse provides a
molecular polarization with extremely high time resolution, whereas
the not‐time resolved detection of the FSRS photon provides
independent and very high wavelength (frequency) resolution. Third,
FSRS spectra appear very similar to the spontaneous Raman spectra
and are devoid of the line shape distortions observed in CARS.
Furthermore, the spectra are linear in the concentration of the
chemical to be analyzed.
A3.4 Epilogue
This chapter explored some of the techniques that make nonlinear
spectroscopy one of the most versatile optical techniques to study
molecular structure and dynamics. Several of these techniques have
experienced explosive growth during the past decade, mostly due to
the availability of pulsed lasers with extremely short pulses, high
repetition rates, and high power. Results of novel techniques and
improvements appear at every major conference on advanced
spectroscopic methods, and at present, new developments seem
unlimited.

References
1 Maker, P.D. and Terhune, R.W. (1965). Study of optical effects
due to an induced polarization third order in the electric field
strength. Physical Review 137: A801–A818.
2 Ziegler, L.D. (1990). Hyper Raman Spectroscopy. Journal of
Raman Spectroscopy 21: 769–779.

3 Harvey, A.B. (ed.) (1981). Chemical Applications of Non‐linear


Raman Spectroscopy. New York: Academic Press.

4 Cheng, J.‐X. et al. (2001). An epi‐detected anti‐Stokes Raman


scattering (E‐CARS) microscope with high spectral resolution and
high sensitivity. The Journal of Physical Chemistry B 105: 1277–
1291.
5 Evans, C.L. et al. (2005). Chemical imaging of tissue in vivo with
video‐rate coherent anti‐Stokes Raman scattering microscopy.
Proceedings of the National Academy of Sciences of the United
States of America 102 (46): 16807–16812.
6 Kano, H. (2008). Molecular vibrational imaging of a human cell
by multiplex coherent anti‐Stokes Raman scattering
microspectroscopy using a supercontinuum light source. Journal
of Raman Spectroscopy 39: 1649–1652.
7 Lee, S.‐Y. et al. (2004). Theory of femtosecond stimulated Raman
spectroscopy. The Journal of Chemical Physics 121 (8): 3632–
3642.
Appendix 4
Fourier Transform (FT) Methodology

A4.1 Introduction to Fourier Transform


Spectroscopy
Infrared (IR) spectroscopy can be carried out experimentally in two
quite different ways. One uses a color sorting device, such as a
monochromator, to isolate one particular IR “color” band and sends
it into the sample where it may be partially absorbed. A spectrum is
collected by sequentially isolating different colors and plot the
absorbance at each color. This is a time‐consuming process since
about 1800 “color bands” need to be analyzed sequentially for a
standard IR spectrum extending from 400 to 4000 cm−1, assuming a
spectral resolution of 2 cm−1. Furthermore, certain aspects of
detector noise make this approach less favorable than an approach
where all frequency bands are sampled simultaneously via an
“interferogram” from which the spectrum is constructed by what is
known as a Fourier transform (FT). Both data acquisition methods
give the same spectral information, but the FT‐methodology is nearly
1000‐fold faster.
The same is true for nuclear magnetic resonance (NMR)
spectroscopy. The spectral information can be obtained by scanning
the external magnetic field at constant radio frequency (or scanning
the radio frequency at constant external magnetic field) to obtain a
spectrum. This may be viewed as the “classical” NMR experiment
and corresponds to the first data acquisition mode in infrared
spectroscopy, described above, where a monochromator is used. In
NMR spectroscopy, FT methodology is now used nearly exclusively
in which the external magnetic field is kept constant, but instead of
varying the frequency of the radiofrequency excitation, a short pulse
of radio frequencies is applied to the sample, which changes its
overall magnetization. When returning to its original magnetization,
the sample emits the energy difference as a radio frequency signal
containing all frequency components known as the “free induction
decay” (FID). This signal is collected by a radio frequency receiver
and subject to FT to give the same spectral information that would
have been observed in the classical NMR experiment. However, the
“pulse FT” method is hundreds times faster and, furthermore, allows
for different NMR experiments that are impossible or near
impossible to carry out in classical NMR mode.
So, what is this magical “Fourier transform” procedure? Basically, it
is a mathematical method that allows signals to be collected in a
different (and more suitable) “domain” and transform them back
into the domain that spectroscopists are more familiar with. No
information is gained or lost in the process, and it is possible to
analyze the interferogram (in IR spectroscopy) or the free induction
decay (in NMR spectroscopy) directly. However, it will be shown that
the information required in spectroscopy, such as chemical shifts or
group frequencies, are much more easily discerned after FT.
To indicate qualitatively how Fourier Transform methodology works,
we resort to an example from music that demonstrates how data can
be represented in different domains.

A4.2 Data Representation in Different


Domains
Consider, for example, the musical note “A,” the 440 Hz sound
produced by a violin's or guitar's “A” string. A plot of such a sound
wave is shown in Figure A4.1a which shows the amplitude of the
sound wave as a function of time. Since this signal varies periodically
with time, it can be represented as a sine or cosine wave with
frequency ν = 440 Hz and an intensity (amplitude). However, one
could depict the same information by plotting the intensity in a
graph that has frequency on the abscissa. This results in a graph
shown in Figure A4.1b. This panel indicates that all frequencies,
except 440 Hz, have no amplitude or are not represented, while the
frequency at 440 Hz appears with a given amplitude. The “frequency
domain” representations shown in Figure A4.1b has the advantage
that a sound signal that contains more than one frequency can be
interpreted more easily. This is shown in Figure A4.2. Here, a signal
consisting of two sine waves, at 440 and 880 Hz, is shown in (a).
This more complex and harder to interpret amplitude vs. time
representation is much clearer when depicted in the amplitudes vs.
frequency plot shown in (b).
The two representations depicted above are mathematically related
by a Fourier transform. To introduce the concept of Fourier
transforms, it is best to first discuss the concept of Fourier series and
transition to Fourier transforms from there.

A4.3 Fourier Series


Fourier series expansion or harmonic analysis extracts appropriately
weighted harmonic components from a general periodic waveform.
Any function f(x), which is periodic between −π and +π (or L to +L,
or 0 to 2π), can be expanded in this interval by a Fourier series. The
Fourier series expansion of the function f(x) is defined by

(A4.1)
Figure A4.1 Representation of data in different domains. (a)
Graph of the intensity vs. time of a sound wave (the musical note “A,”
top trace) and its first (middle trace) and second harmonic (bottom)
or octaves. (b) Representation of the same information in an
intensity vs. frequency display.
Figure A4.2 (a) Intensity vs. time and (b) intensity vs. frequency
plot of a sound signal consisting of two notes at 440 and 880 Hz.
where the expansion coefficients cn are given by

(A4.2)

In the case of real functions, the expansion given in Eq. (A4.1) takes
the form

(A4.3)
Figure A4.3 Approximation of a square wave function (heavy
black line) by a scaled sum of harmonic frequencies (n = 1 to n = 9).
See text for details.
with the real expansion coefficients given by

and

(A4.4)
An example of Fourier series expansion will be presented next.
Consider a square wave, given by Eq. (A4.5):

(A4.5)

This function is shown as the heavy black trace in Figure A4.3 and is
assumed to repeat periodically. Substituting Eq. (A4.5) into the
equations for the expansion coefficients an and bn (Eq. [A4.4]) and
integrating from 0 to 2π, one finds that all the terms an will be zero:

(A4.6)

(A4.7)

whereas the terms bn assume the values:

(A4.8)

for even values of n, and


(A4.9)

Thus, a square wave can be expanded into an infinite series of all odd
harmonics, scaled by 1/n:

(A4.10)

This is shown in Figure A4.3 for n = 1 to n = 9. It is obvious from this


graph that an increasing number of higher harmonics improves the
fit between the square wave and the sum of all the co‐added
harmonics. Similar expansions can be carried out to approximate
any periodic function (such as a saw tooth function and other, more
complex functions) by a sum of harmonics.

A4.4 Fourier Transform


Next, the standard concepts of the Fourier transform will be
introduced. A logical connection between the principles of Fourier
expansions and Fourier transforms can be made by substituting

(A4.11)
into Eqs. (A4.1) and (A4.2) to obtain

(A4.12)

(A4.13)
and letting the interval, L, over which the function is expanded, go to
infinity. Thus, k gets very small, and one can substitute the sum in
Eq. (4.12) by an integral:

(A4.14)

(A4.15)

where C is a normalization constant.


Thus, one may view the process of taking a Fourier transform as a
harmonic analysis with infinitely small increments in the frequency
intervals. Equations A4.14 and A4.15 are called a “Fourier pair” and
each equation in such a pair completely defines the other. Therefore,
no information is lost when the signal is transformed from one
domain to the other.

Figure A4.4 Examples of Fourier transforms (FTs). (a) The FT of


a delta function is a cosine function (black: real part, gray: imaginary
part). Note that the real and imaginary parts are 90° out of phase, (b)
The FT of a broad spectral distribution is a sharp, symmetric
interferogram. The FT's shown here were calculated via the
Microsoft EXCEL fast FT implementation. The interferogram shown
in (b) was unfolded (see text).
Next, some important properties of Fourier transforms will be
discussed. Equations (A4.14) and (A4.15) show that a Fourier
transform is a complex operation; thus, the output of FT calculation,
in general, have a real and an imaginary part.
This is shown, for example, for a shifted δ‐function, , which
one may think of an infinitely narrow‐band (monochromatic) light
source, in Figure A4.4a. Its FT is a periodic cosine function with
frequency , and the imaginary part is phase‐shifted (or sine)
function shown by the gray trace in Figure A4.4a. The Fourier
transform of a broad spectral distribution gives a very sharp
interferogram, shown in Figure A4.4b.

A4.5 Discrete and Fast Fourier Transform


Algorithms
The equations given so far for Fourier transform pairs describe the
process and concepts but not a practical method to carry out Fourier
transforms of a set of data. This transformation between the two data
domains is performed by an algorithm known as fast Fourier
transform (FFT) algorithm.
FFT is based on the principles of the discrete Fourier transform since
the interferogram is not sampled continuously, but at discrete times,
the interferogram is obtained as a one‐dimensional vector of digital
values. For such a situation, the process of taking the Fourier
transform can be written as

(A4.16)

(A4.17)
In Eqs. (A4.16) and (A4.17), N is the total number of data points, T is
the sampling interval, and n and k are the running indexes in g and
G space, respectively. These equations are the discrete (point‐by‐
point) versions of Eqs. (A4.14) and (A4.15). It is interesting to note
that Eq. (A4.16) resembles very much the equation at the starting
point of Fourier series expansion (Eq. [A4.1]).
Setting n/NT = m and kT = p, Eq. (A4.17) can be written as

(A4.18)

where

(A4.19)
Equation (A4.19) can be cast into matrix notation:

(A4.20)
Thus, the computation of a discrete Fourier transform from g
(interferogram) to G (spectrum) space is reduced to computing a
(complex) transformation matrix Wmp and multiplying the vector of
discrete points with this matrix. For typical spectroscopic
applications, a sample set may consist of 8K data points (i.e. 8192
points in the g[p] vector). The Fourier transform operation,
according to Eq. (A4.20), requires for each of the 8192 points in G
space an 8K × 8K matrix to be multiplied by an 8K vector. Such
matrix manipulations are slow, since for each data point, 8K
multiplications and 8K additions are required.
This problem was alleviated by the FFT algorithm, developed by
Cooley and Tukey [1], which avoids the problem of a large number of
multiplications and additions by factoring the W matrix into sparse
matrices that have many zero elements. It can be shown that such
factoring is always possible, but the factoring will require reordering
the entries in the G and g vectors. Furthermore, the FFT algorithm
only works for data vectors that have integer powers of 2 (256, 512,
1024, etc.) entries. A detailed discussion of the FFT algorithm is
beyond the scope of this chapter, and the reader is referred to the
literature [1, 2]. With the increased computational power of modern
desktop machines and the implementation of FFT routines in the
MATLAB environment, FFT computations can be carried out for a
1024 point data vector in a few milliseconds.

A4.6 FT Implementation in EXCEL or


MATLAB
A final comment is appropriate about the presentation of the FFT
results, for both MATLAB and EXCEL implementations. As pointed
out in the Caption to Figure A4.4, all Fourier transform examples
shown in this Appendix were carried out using Microsoft EXCEL.
In both FT–IR and FT–NMR, data are collected in the time‐domain
(time after the broadband excitation pulse in NMR, or time of the
interferometer mirror travel after passing the “zero‐path difference”
point). These time points will be referred as the “time zero point” in
the time domain. The FT transforms the intensity vs. time
interferogram into intensity vs. frequency space.
Figure A4.5 Panel (a): Real part of a reverse transform of a
spectrum back to interferogram domain. Notice the unfolded
representation of the interferogram. Panel (b): Imaginary part of
interferogram shown in (a). Panel (c): Interferogram shown in (a)
folded to demonstrate the symmetry of the real part of the
interferogram. Panel (d): Interferogram shown in (b) folded to
demonstrate the antisymmetric nature of the imaginary part of the
interferogram.
A spectroscopist, therefore, expects an interferogram similar to
Figure 8.7a, that is, with a starting point at “time zero” point (in FT–
NMR), or at the center burst or “zero‐path difference” (in FT–IR
spectroscopy). However, if a spectrum is reverse transformed into
the time domain using the FFT algorithm as included in MATLAB or
EXCEL, both the real and imaginary parts are presented in a “folded”
form, as shown in Figure A4.5, Panels (A) and (B). Here, the “time
zero” or the “center burst” is shown on the far left of the graph, with
the next data point at the far right of the graph; that is, the reverse
transform is unfolded such that points 1–256 are moved to positions
257–512, and the points 257–512 are moved to positions 1–256 in a
512 point transform. After un‐folding, the interferograms shown in
Panel (C) and (D) are obtained. Notice that the real part of the
interferogram is symmetric about the “time zero” point, whereas the
imaginary part is antisymmetric about this point.

References
1 Cooley, J.W. and Tukey, J.W. (1965). An algorithm for the
machine calculation of complex Fourier series. Mathematics and
Computation 19: 297–310.
2 Brigham, E.O. (1974). The Fast Fourier Transform. Englewood,
NJ: Prentice Hall.
Appendix 5
Description of Spin Wavefunctions by Pauli
Spin Matrices
The introduction of electronic spin in Chapter 7 followed an
argument that compared the eigenvalue of the total angular
momentum operator

(6.43)

and the eigenvalue of a Cartesian component, Lz

(6.44)

to the total spin operator, , and its Cartesian components , and


it was argued that, in analogy, the eigenvalues of the total spin
operator for an electron should be

(7.28)

and the eigenvalues of the operator

Here, the angular momentum quantum number J can have integer


values ≥0, and K from –J to +J, the spin quantum numbers “s” for
an electron are confined to the value ½. Thus, the eigenvalues for
and operators are

(7.29)
(7.30)

In Eqs. (7.28) and (7.29), the spin functions α and β were not further
defined, except that they were postulated to be orthonormal
functions:

and

(A5.1)

and described as “spin‐up” and “spin‐down” or “clockwise” and “anti‐


clockwise” motion. This latter description is somewhat simplistic
since it implies a truly spinning particle which the electron is not.
Rather, it was pointed out that the spin is an inherent quantity that is
much more difficult to visualize.
Historically, it is interesting to note that the wave mechanical
approach pursued by Schrödinger did not include the spin quantum
number s (or ms). However, the alternative formulation (based on
matrix mechanics) of quantum mechanics by Heisenberg, developed
nearly simultaneously to Schrödinger's work, incorporated all four
quantum numbers n, l, ml, and ms for the hydrogen atom. At this
point in time, experimental work had demonstrated that there exist
two spin states and that each spin state has an (experimentally
verifiable) eigenvalue of , leading to an energy difference of the
two states of

(A5.2)
(see Appendix 1)
Since there are two eigenvalues, a matrix‐based theoretical
formalism, based on Heisenberg's matrix mechanics, was devised by
Pauli [1] in which the spin Hamiltonians consist of 2 × 2 matrices,
and the spin functions of 2 × 1 column vectors, as discussed in the
next section. These following equations are presented without proof
and should be construed as a short introduction into matrix
mechanics and spin matrices.

A5.1 The Formulation of Spin


Eigenfunctions α and β as Vectors
In matrix mechanics, the wavefunctions are represented by vectors
and the operators as matrices. We start by discussing the spin
wavefunctions that we know have two eigenvalues (see Eq. 7.29
above). However, the Stern–Gerlach experimental (Section 7.5)
revealed that any linear combination of these eigenfunctions α and β
are also valid eigenfunctions. Thus, any linear combination

(A5.3)

is also a solution of the spin Hamiltonian. This insight is due to the


results of the Stern–Gerlach experiment that revealed that any
number of mutually perpendicular magnetic fields split a beam of
silver atoms (with a single unpaired electron) passing through them
into two beams, since the components of the spin momentum,
cannot be determined simultaneously.
In the matrix formulation of electron spins, the spin wavefunction
ψαβ is represented by a column vector:

(A5.4)

Arguments about the orthonormality of the spin wavefunctions (see


Eq. A5.1 above) leads to the final form of the spin wavefunctions α
and β as

(A5.5)

A5.2 Form of the Pauli Spin Matrices


The spin operators and its Cartesian components
are defined by the Pauli spin matrices as

(A5.6)

The derivation of these spin matrices is beyond the scope of this


discussion. To simplify the calculations below, the normalization
factor often is omitted, and the matrices are written as

(A5.7)

Since the total spin operator, is given by

(A5.8)

it can be evaluated by squaring each of the spin matrices and


summing them up:

(A5.9)

Thus,
(A5.10)

and

(A5.11)

in agreement with Eq. (7.29) above.


Following the argument cited above that the spin operators follow
the same commutation rules as the angular momentum operator, we
evaluate the commutators:

(A5.12)

(A5.13)

Similarly, it can be shown easily that the other two commutators of


the Cartesian components are

(A5.14)
Thus, it was shown that the spin matrices, indeed, reproduce the
commutation properties of the Cartesian components of the angular
momentum operator discussed in Chapter 6. By similar arguments,
it can be shown that
(A5.15)

The other two Cartesian components also commute with the σ2


operator.

A5.3 Eigenvalues of the Spin Matrices


At this point, the eigenvalues of the spin operators can be evaluated
easily. For example, the relationships

(7.29)

can be verified by simple matrix multiplication:

(A5.16)

(A5.17)

and

(A5.18)
(A5.19)

Finally, the orthonormality conditions

and

(A5.1)
can be verified as follows. To form the inner products of two spin
vectors, one has to transpose the first vector (in fact, one has to form
the adjoint vector, where all terms are the complex conjugates.
However, since in this example the vectors are real, the transpose
equals the adjoint vector):

(A5.20)

Thus, the results of the matrix mechanical treatment of the spin


functions provide the same results that were obtained by just
transferring the results from the angular momentum operators
and , but it has been demonstrated here how the matrix
mechanics formalism works.
Reference
1 Pauli Matrices (2020). Wikipedia, the Free Encyclopedia.
https://en.wikipedia.org/wiki/Pauli_matrices.
Index

a
absorption 1, 3, 7–12, 32, 33, 39–45, 51, 56–59, 62, 63, 75–81, 84,
87–89, 96, 104–113, 115, 123, 125, 138, 147, 156, 159–161, 172–174,
179, 180, 182, 183, 187, 189, 190, 192, 193, 214, 216, 217, 235
absorptivity 42, 45
adenine (A) 180
ammonium sodium tartrate 187
androstane 190, 191
angular momentum 19, 93–95, 97–100, 103, 110, 117, 127–129,
132–135, 137, 139, 157–160, 171, 172, 189, 253, 255–257
anharmonic force constant 60
anharmonic oscillator 59–62, 87, 125, 173, 225
anharmonic potential 56, 61, 176, 182
anharmonicity 49, 61–63, 65, 77, 89, 112, 225
anomalous dispersion 12, 77–81, 187
anomalous Zeeman 127
anti‐Stokes Raman process 83, 84, 238, 239
associated legendre polynomials 101, 102, 104, 109
associative law 203
asymmetric molecule 186–191
asymmetric top rotor 96–98, 110
atomic absorption 3, 7–10, 161
atomic number 120, 154, 155
atomic radii 156
atomic spectra 2, 159, 160
atomic spectroscopy 10, 157, 160–162
Aufbau principle 153–155

b
beat frequencies 82, 83
Beer–Lambert law 43, 162
benzene 175, 180, 202
blackbody radiation 3–5
blueshift 183, 185
Bohr radius 117, 122, 123
Boltzmann distribution 4, 44, 84, 90, 105, 106, 221
bond order 170, 172, 173, 176
Born–Oppenheimer approximation 86, 164, 165, 174–176, 199
boundary conditions 23, 24, 28, 51, 53, 99, 225
Brewster angle 47
1,3‐butadiene 179

c
camphor 190, 192
carbon disulfide 76
carbonyl chromophore 178–179, 190
carotene 180
Cartesian component 70, 97–99, 103, 128, 132, 133, 139, 209, 210,
215, 217, 253, 255, 256
Cartesian displacement coordinates 69–72, 85, 211–214
Cauchy integral 81
center of inversion 88, 171, 200, 201, 208, 217
centrifugal distortion 112
centrifugal effects 107
character table 85, 208–210, 213, 214, 216–218
characteristic vibrational frequency 51
chemical bonding 115, 163–167
chemical shift 138, 140–141, 143–146, 244
chirality 185–187, 189, 192, 193
chloroform 88–90, 108, 201
chromophore 178–181, 190–193
circular dichroism (CD) 81, 179, 185–193
circularly polarized light 81, 185–187, 190
Clebsch–Gordan expansion 157, 158
coherent anti‐Stokes Raman scattering (CARS) 184, 233–235, 237–
239
nonlinear Raman effects 233, 234, 236, 237, 240
collisional energy loss 182–183
color band 243
combination band 77, 89
commutation 18, 103, 128, 133, 134, 255, 256
commutators 19, 97, 255, 256
conservation of momenta 236
constructive and destructive interference 2
C3 operation 207
Coulomb's law 116
C2ν point group 203
crystal field theory 181
cytosine (C) 180

d
de Broglie equation 6, 16
degeneracy 120, 121
detector noise 243
2‐deuterobutane 190
diamagnetic response 140
dichloroethene 217
dielectric susceptibility 139, 234, 236
differential operator 16
1,3‐dimethylallene 187
dipolar coupling 191
dipole‐allowed absorption 40–42, 124

dipole‐allowed transitions 76–77, 124, 172, 199


dipole moment 38, 40, 62, 76, 77, 81–83, 87, 88, 96, 105, 112, 113,
124, 189, 191, 193, 199, 210, 216, 233
dipole strength 179, 189, 191
Dirac 1, 128, 129
discrete Fourier transform 248, 249
disentanglement 241
dissymmetric molecules 187, 188, 191–192
DNA bases 180
doubly degenerate 88, 181, 208

e
effective nuclear charge 151–152
Einstein coefficients 42–45
elastic scattering 81
electric and magnetic dipole transition moments 189
electric dipole moment 38, 189, 191, 216
electric dipole operator 188, 189
electric field 2, 3, 38, 40, 82, 83, 90, 109, 138, 188, 189, 231, 233,
234
electric transition moment 3, 190
electromagnetic radiation 2, 3, 10, 11, 18, 25, 37–46, 51, 56, 78, 80,
81, 83, 131, 134, 136–139, 146, 147, 189, 230–231, 233
electron configuration 127, 158
electron correlation 151, 152, 169
electron–electron correlation 223
electron–electron repulsion 151, 164
electronic circular dichroism and optical rotation 185–193
electronic spectra 173
of diatomic molecules 173–177
of polyatomic molecules 177–181
electronic spectroscopy 115, 160, 163, 211
electronic transition moments 84, 85, 176
electron paramagnetic resonance 131, 134
electron spin 11, 126–129, 131, 134, 152, 254
enantiomeric forms 186, 190
enantiomers 185–187, 189, 190
energy eigenvalues 17, 22, 24, 28, 51–56, 60, 74, 95, 96, 99–103,
118, 121, 123, 125, 142, 153, 165, 191, 214, 224, 229, 230
energy expectation value 165, 168, 224
energy states 3, 8–10, 28, 34, 40, 43–45, 55, 63, 90, 115, 127, 128,
131, 134, 136, 139, 140, 147, 155, 163, 176, 195
estrogen 190
exciton couplet 190, 192
exciton formalism 192
excitonic state 191
exciton model 191–192
expectation value 17, 40, 83, 165, 168, 224, 226
external magnetic field 128, 129, 131, 137–140, 144, 145, 147, 243

f
fast Fourier transform (FFT) algorithms 248–249
femtosecond stimulated Raman scattering (FSRS) 240–242
first order non‐linear susceptibility 183
fluorescence microscopic imaging 181
fluorescence resonance energy transfer 181
fluorescence spectroscopy 174, 181–185
fluorochlorobromomethane 186, 187
formaldehyde 178, 179, 202
Fourier series 244–247, 249
Fourier transform 131, 146–147, 243–244, 247–249
Franck–Condon factor 176, 184
Franck–Condon principle 174–177, 181–184
free induction decay (FID) 139, 146–148, 243, 244
frequency doubling 234–237
FT‐IR spectroscopy 147, 249, 250
full width at half maximum (FWHM) 77
g
gain medium 46, 47
gas laser 46
Gaussian distribution 79, 169, 215
Gaussian profile 79
glyceraldehyde 187
green fluorescent protein (GFP) 182
group frequencies 93, 178, 244
group representations 204–211
guanine (G) 180

h
half‐filled shell 155
Hamiltonian 17, 22, 23, 27, 28, 51, 132, 141, 143, 144, 151–152, 157,
158, 164, 165, 168, 175, 214, 224–226, 254
Hamilton operator 17
handedness 185, 186, 189
harmonic analysis 244, 247
harmonic oscillator 38, 49–66, 69, 70, 73, 74, 76, 87, 101, 102, 104,
110, 111, 125, 173, 223, 225
harmonic oscillator Schrödinger equation 51–56
harmonic oscillator wavefunctions 21, 52, 55
harmonic potential energy function 60
Hartree–Fock method 152, 153, 155
He–Ne laser 46, 47, 159
heat capacity 91
Heisenbergs matrix mechanic 254
Heisenberg's uncertainty principle 15, 19, 55, 241
helical molecules 187
α‐helix 192
Hermite's differential equation 51, 118, 125
Hermite polynomials 53, 54, 57, 101, 102, 104, 215
high spin 181
highest occupied molecular orbital (HOMO) 32, 163, 180
hollow cathode lamp (HCL) 3, 161
homonuclear diatomic molecule 62, 105, 168–172, 195
hot bands 62–65, 79, 89
Hund's rule 154, 171, 172
hydrogen atom 2, 7–10, 12, 20, 28, 33, 93, 103, 115–129, 151, 156,
159, 164, 166, 168, 187, 201–203, 254
hydrogen‐like orbitals 169

hyper‐Raman scattering 234

hyper‐Raman effect 139, 233, 235, 236


hyperpolarizability tensor 234

i
identity element 186, 200, 203
improper axes of rotation 200
induced dipole moment 82, 83, 233
induced electric dipole 233
inelastic scattering 81
inertial tensor 95
infrared absorption spectroscopy 62, 63, 76–81, 88, 111, 214, 235
infrared spectroscopy 93, 235, 243
intersystem crossing 182, 183
intrinsic fluorescence 181
inverse operation 203
inversion symmetry element 88, 217
ionization energy 6, 155
ionization potential 155
irreducible representations 181, 189, 204–209, 212–214, 216, 217,
220
IR vibrational circular dichroism (VCD) 192
isotopic species 63–65, 75
isotopic splitting 64

j
Jablonski diagram 182–183
J‐coupling 143, 145

k
Kramers–Kronig transform 12, 80, 81
Kronecker symbol 25, 54

l
Lagrange's equation of motion 71
LaGuerre differential equation 117
LaGuerre polynomials 117, 119
Laplace operator 116
Larmor frequency 139, 147
laser theory 43
Legendre differential equations 118, 125
Legendre polynomials 101, 109
lifetimes 77–79
ligand field theory 181
line shapes 77–79
linearly polarized light 187
linear molecules 75, 76, 94, 96, 103, 105–109
lithium 160
Lorentz force law 136
Lorentzian band 78, 79
lowest unoccupied molecular orbital (LUMO) 32, 180
low spin 181

m
magnetic dipole transition operator 189
magnetic moment 128, 134–136, 139, 147, 189
magnetic transition moment 3, 188–192
magnetic quantum numbers 103, 117, 126, 127
magnetization 11, 129, 131, 137–139, 147, 148, 189, 243
magnetogyric ratio 135
mass‐weighted Cartesian displacement coordinates 70–72
Maxwell's equation 1, 2, 5, 6
3‐methylheptane 190
mirror image 186, 187
molar extinction coefficient 42, 43, 76, 77, 178–180
molecular orbital theory 163, 164, 168–170
moment of inertia 75, 93–95, 98, 103, 107, 108, 112, 132
momentum conservation 236
Morse potential 59
multidimensional NMR processes 233
multi‐electron systems 133, 151–162
multiple spin interaction 144–145
multispin system 141–146
mutual exclusion rule 217

n
natural broadening 79
Nd:YAG laser 90
net magnetization 129, 139, 147–148, 189
Newtonian mechanics 1
Newton's second law of motion 50
noncoherent nonlinear effects 234–235
noncoherent three‐photon effect 233

nondegenerate hyper‐Rayleigh effect 234


nonlinear spectroscopic techniques 233–242
nonlinear spectroscopy 90, 233, 235, 242
normal coordinate analysis 69, 70, 72
normalization 25–27, 41, 53, 100, 123, 166, 247, 255
normal modes of vibration 69–72, 74, 75, 77, 211, 212
nuclear magnetic resonance (NMR) 93, 129, 131–148, 243
nuclear magneton 135, 221
nuclear spin 129, 134–139, 141, 147
nuclear spin energy states 131, 139

o
oblate symmetric top rotor 96
oblate top rotor 108
olefins 179–180
one‐dimensional representation 206
optical activity 185, 187
asymmetric molecules 188–191
dissymmetric molecules 191–192
manifestation of 187–188
vibrational 192–193
optical rotation 185–193
optical rotatory dispersion (ORD) 81, 187
orbital angular momentum quantum number 117
orbital approximation 151, 152, 223
orthogonality condition 208
orthogonality theorem 206
orthonormal functions 253
orthonormality condition 254, 256
orthonormal vector space 25, 54, 120

p
parallel envelope 113
paramagnetic 155, 170
parity 21, 41, 42, 56, 57, 65, 76, 171
particle‐in‐a‐box 17, 21, 23–26, 29–31, 33, 37, 38, 40–42, 49, 58,
65, 66, 125, 223, 226–230
particle‐in‐a‐2D‐box 27–28, 33, 73
particle–wave duality 5, 6
partition function 91
Pauli exclusion principle 18, 152–154
Pauli spin matrices 253–256
P‐branch 112

2,3‐pentadiene 186, 187


periodic chart of elements 115, 155
perpendicular 2, 113, 201, 202, 254
perturbation Hamiltonian 225, 226
perturbation method 34, 37, 60, 144, 225
perturbed wavefunction 225, 226, 228, 229
phase‐matching 236–239
phenylalanine 180
phosphorescence 178, 182, 183
photoelectric effect 1, 2, 5–7, 12, 155
photon mass 7
Planck's constant 4, 5, 15, 16, 32, 95, 221
polarizability 40, 81–86, 96, 185, 193, 199, 217, 234
polarizability tensor 85, 86, 185, 193, 210
polyenes 31–32
poly‐L‐Pro (II) 192
population inversion 45–47
potential energy 16, 20–22, 24, 27, 29, 30, 33, 34, 49, 50, 55, 56,
59–60, 70–73, 89, 95, 97, 116, 125, 176, 182, 226, 227, 230
principal axes of inertia 95, 99
prolate symmetric top rotor 96, 108
prolate top rotor 97, 108, 109
proper axes of rotation 200–201
pseudo‐Voight function 79
pulse Fourier transform (FT) NMR spectroscopy 138, 146
pulse FT method 131, 243

q
Q‐branch 112, 113
quadrant or octant rules 190–191
quantized energy level 24
quantum cascade lasers 31–34
quantum dots 31–34
quantum mechanics 1–12, 15–21, 24, 28, 37, 39, 49, 69, 97, 103,
115, 128, 223, 254
quantum number 8, 18, 24, 26, 28, 53, 56, 57, 61, 62, 77, 100, 101,
103, 107, 109, 112, 116–118, 120, 124, 126–129, 133–135, 145, 153,
154, 157–160, 173, 178, 253, 254

r
racemate 187
racemic mixture 187
radial distribution function 122, 123
Raman optical activity (ROA) 192
Raman scattering 12, 39, 62, 81, 82, 84, 87, 88, 173, 217, 233–236,
240
Raman spectroscopy 81–86, 96, 185, 192, 210, 211, 217, 234, 235,
240, 241
random coil 192, 193
Rayleigh scattering 81, 84
R‐branch 112
recursion formula 52–54, 57, 101, 102, 104, 109, 118
redshift 183, 185
reducible representation 209–214
reflection 42, 47, 201, 204, 208, 214
reflection by mirror plane 186, 200, 201, 204
refraction 42, 79, 80
refractive index 11, 12, 42, 79–81, 84, 187, 188
resonance condition 40, 86, 138, 146
resonance/off‐resonance 11
resonance Raman process 185
resonance Raman spectroscopy 85
resonator structure 46, 47
rigid rotator 107
rotation–reflection axes 200, 201
rotational constants 95, 96, 103, 108, 112, 113, 173
rotational kinetic energy 94, 95, 132
rotational quantum number 100, 112, 158, 173
rotational Raman spectrum 96
rotational Schrödinger equation 97, 99–103, 115
rotational spectroscopy 93–113, 117, 140, 163
rotational strength 189, 192
rotational wavefunction 93, 101
rot–vibrational transitions 110–113
ruby laser 233
Rydberg constant 8, 10, 118, 221

s
scattering tensor 85, 86, 234
Schrödinger equation 16, 18, 20, 23–24, 29, 37, 49, 73, 93, 99, 115,
116, 125, 151, 157, 175, 224, 225
second harmonic generation (SHG) 184, 235–237
second‐order dielectric susceptibility 139, 236

second‐order nonlinear susceptibility 185

second‐row diatomic molecules 169


selection rules 40, 57, 85, 102, 104, 109, 124, 137, 160, 172, 199,
217, 235
for absorption 57
for dipole‐allowed processes 214–216
for electronic transition 178
for harmonic oscillator 56–60, 102
for homonuclear diatomic molecules 171–172
for IR and Raman spectroscopy 87–88
for one‐spin nuclear system 137–138
for particle in a box 40–42
for Raman effect 85
for Raman scattering 217
for rotational transitions 104
for transitions in atomic species 160
series expansion 17, 52, 53, 223–225, 229, 233
SF6 75

β‐sheet 192
shielding effect 138, 140, 141, 156
singlet state 158, 172, 182, 183
singlet to triplet transition 160, 172, 178, 183
Slater determinant 18, 152, 153
sloped baseline 226
sodium D line 127
sodium doublet 127
sodium 589nm line 159, 161
spatial quantization 103, 129, 133, 139
spectral transitions 3, 9, 37
spectroscopic transitions xi
spherical harmonic functions 33, 98, 99, 101, 102, 115, 117, 119, 121
spherical polar coordinates 16, 98, 116, 117
spherical top rotor 96, 109
spin angular momentum 103, 128, 129, 132–134, 139, 156, 157, 159,
160, 171
spin functions 128, 134, 136, 152, 153, 253, 254, 257
spin–lattice relaxation 147
spin multiplicity 154, 158, 159, 171
spin‐orbit coupling 158, 161

spin‐pairing energy 155


spin spectroscopy 93, 129, 131
spin–spin coupling 143–146
spin state population 137–140
spin states 11, 128, 134, 136–140, 144, 145, 158, 254
spin wavefunctions 132, 133, 142, 152, 153, 253–254
spontaneous emission 45–47, 78, 182
spontaneous, nonresonant Raman process 185
spontaneous Raman 184, 236, 240, 242
square wave 246, 247
Stark splitting 109
stationary states 3, 8, 24, 37–40, 43, 51, 78, 85, 97, 230
stationary‐state wavefunctions 25, 37, 40, 83
statistical thermodynamics 91
Stefan–Boltzmann law 5
stereochemistry 188, 190, 191
Stern–Gerlach experiment 127, 254
stimulated emission 34, 40, 43–45, 48
stimulated Raman scattering (SRS) 240–242
Stokes Raman frequencies 83, 84, 238
successive application 203–205
sum‐frequency generation 234, 237
superimposable image 186
symmetric top rotors 96, 97, 108
symmetric tops 96, 108, 109
symmetry group 186, 200–204, 209, 213, 215, 217
symmetry operation 186, 200–207, 210–212, 214
symmetry representation 199, 210–214

t
Taylor series 82
term symbols 127, 156–161, 171–172, 195
testosterone 190
tetrachloroethene 235
tetramethylsilane 141
thermal ellipsoids 75
thermal energy 90, 105
3rd law of thermodynamics 49, 55
third‐order nonlinear susceptibility 238

third‐order susceptibility tensor 239


thymine (T) 180
time‐dependent perturbation 37–40, 230–231

time‐dependent Schrödinger equation 18, 37–39, 230

time‐dependent wavefunction 18, 38, 39, 83

time‐independent perturbation theory 225–226


toluene 180
total orbital angular momentum 157–159, 172
total spin angular momentum 139, 159, 160, 171
trace 25, 41, 42, 187, 205, 207, 211, 212, 214, 241, 245, 246, 248
transformation matrix 72, 204, 207, 211, 212, 214, 249
transition coupling 191–192
transition metals 154, 155, 181
transition moment 3, 39–45, 47, 56–59, 84–87, 104, 105, 107, 124,
175, 176, 188–192, 215–217, 231, 234
translational and rotational degrees of freedom 213, 214
trial function 23, 118, 224
triplet state 158, 171, 172, 183
triply degenerate 113, 181, 208
tunnelling 29–31, 34, 56, 230
two‐photon fluorescence 183–184

two‐photon process 39, 85, 87

u
unbound particle 27–31
uncertainty principle 15, 19, 55, 78, 241
v
valance‐shell electron‐pair repulsion (VSEPR) model 93
variation method 169, 224–225
vector space 17, 19, 22, 25, 54, 97, 101, 120, 200, 223, 224
vibrational dephasing 241
vibrational infrared and Raman spectroscopy 69–91
vibrational optical activity 192–193
vibrational Schrödinger equation 51, 52, 54, 60, 73, 214
vibrational spectroscopy 52, 62–65, 69, 70, 78, 79, 90, 93, 97, 178,
193, 199, 211
vibronic absorption spectrum 173–174
Voigt function 79

w
wave vector 2, 29, 237, 239
Wien's law 5
work function 6

z
Zeeman effect 126, 127
zero‐point vibrational energy 74
WILEY END USER LICENSE AGREEMENT
Go to www.wiley.com/go/eula to access Wiley’s ebook EULA.

You might also like