Robert H Lipson - David L. Andrews - Molecular Photophysics and Spectroscopy (2021)
Robert H Lipson - David L. Andrews - Molecular Photophysics and Spectroscopy (2021)
Robert H Lipson
Department of Chemistry, University of Victoria, Victoria, BC, Canada
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system
or transmitted in any form or by any means, electronic, mechanical, photocopying, recording
or otherwise, without the prior permission of the publisher, or as expressly permitted by law or
under terms agreed with the appropriate rights organization. Multiple copying is permitted in
accordance with the terms of licences issued by the Copyright Licensing Agency, the Copyright
Clearance Centre and other reproduction rights organizations.
Certain images in this publication have been obtained by the author(s) from the Wikipedia/
Wikimedia website, where they were made available under a Creative Commons licence or stated
to be in the public domain. Please see individual figure captions in this publication for details. To
the extent that the law allows, IOP Publishing [and full partner name (if applicable)] disclaim any
liability that any person may suffer as a result of accessing, using or forwarding the image(s). Any
reuse rights should be checked and permission should be sought if necessary from Wikipedia/
Wikimedia and/or the copyright owner (as appropriate) before using or forwarding the image(s).
Permission to make use of IOP Publishing content other than as set out above may be sought
at [email protected].
David L Andrews and Robert H Lipson have asserted their right to be identified as the authors of
this work in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.
DOI 10.1088/978-0-7503-3683-3
Version: 20210701
IOP ebooks
British Library Cataloguing-in-Publication Data: A catalogue record for this book is available
from the British Library.
US Office: IOP Publishing, Inc., 190 North Independence Mall West, Suite 601, Philadelphia,
PA 19106, USA
Dedicated to Karen Ann
Ever held in loving memory
Contents
vii
Molecular Photophysics and Spectroscopy (Second Edition)
viii
Molecular Photophysics and Spectroscopy (Second Edition)
Bibliography 16-1
Index I-1
ix
Preface to the second edition
x
Preface to the first edition
xi
Author biography
David L Andrews
David Andrews is known for advances in the theory of quantum and nonlinear optics,
fundamental photonics, energy transport, and optical vortices. He has published 400
papers and over twenty books, including as author or co-author Lasers in Chemistry,
Optical Harmonics in Molecular Systems, Optical Nanomanipulation, and an
Introduction to Photon Science and Technology. David has taught extensively at the
University of East Anglia and run numerous spectroscopy courses for industry: he has
twice been Chair of the Royal Society of Chemistry Molecular Spectroscopy Group.
Andrews is a Chartered Chemist and Chartered Physicist, and he is the 2021 President
of SPIE, the international society for optics and photonics.
Robert H Lipson
Rob Lipson has published extensively in the areas of laser spectroscopy, nano-
science, and novel material for photonics applications. He has taught a wide variety
of undergraduate and graduate level physical chemistry courses at Western
University and the University of Victoria (UVic). Administratively, he has served
as Chair of the Department of Chemistry at Western and two terms as Dean of
Science at UVic. In addition to his academic work, he was also a long-serving Senior
Editor of the Canadian Journal of Chemistry, and a panel member on numerous
funding and research integrity committees for the federal Government of Canada.
Lipson is a Fellow of the Chemical Institute of Canada.
xii
List of symbols
Mostly of the symbols used in the following text are standard, including several
unfortunately overworked symbols that commonly have more than one meaning.
Usually, the correct meaning in any appearance is clearly given by the context, and
definitions are given in each of the relevant sections.
xiii
Molecular Photophysics and Spectroscopy (Second Edition)
P degree of polarisation
q electron coordinate; also charge
Q generalised nuclear coordinate
Qe nuclear coordinate in equilibrium position
r bond length; also florescence anisotropy
r radial displacement vector
re equilibrium bond length
R distance
RH Rydberg constant
Ri rotation about the axis denoted by the subscript
R0 Förster distance
s spin
Sn improper rotation about an axis of 2π/n radians; also nth singlet state of a molecule
S total spin; also label for a singlet electronic state
t time
T absolute temperature; also kinetic energy; also transmittance
T tetrahedral; also 3D irreducible representation in group theory
Ti spatial translation along the Cartesian axis denoted by the subscript
u unified atomic mass unit
u ungerade (odd parity)
v velocity; also vibrational quantum number
V potential energy; also volume
VM molar volume
w beam radius
xe anharmonicity constant
z propagation distance
Z nuclear charge
Zeff effective nuclear charge
α polarizability; also absorption coefficient; also fine structure constant
α′ mean polarizability derivative
Δν full-width at half-maximum linewidth
ε extinction coefficient (molar absorption coefficient)
ε0 vacuum permittivity
γ′ polarizability anisotropy derivative
θ angle
Γ transition rate
κ orientation factor
λ wavelength
Λ axial component of angular momentum
μ electric dipole
ν frequency
ν´ scattered frequency
ν0 classical frequency of vibration; also frequency of peak resonance
ν wave-number
Π electronic state of a linear molecule with Λ = 1
π off-bond axis molecular orbital
ψ general wavefunction; also electronic wavefunction
Ψ vibrational wavefunction
Φ molecular wavefunction
ρ density; also density of states; also depolarization ratio
xiv
Molecular Photophysics and Spectroscopy (Second Edition)
σ absorption cross-section
σ on-axis molecular orbital; also reflection operation in group theory
Σ axial projection of atomic spin; also diatomic or linear molecule state label when Λ = 0
τ decay lifetime
ω circular frequency
Ω axial component of total angular momentum
xv
IOP Publishing
Chapter 1
Introduction to molecular spectroscopy and
photophysics
observed dependence on optical wavelength provides the crucial basis for spectro-
scopic interpretation. The origin of this dependence is of course energy conservation:
the requirement that the balance of photons, delivered to and emerging from each
particle in the course of a single process, must match the transition energy lost or
acquired by that particle—and the latter energy change is itself constrained by the
restriction of moving between quantum energy levels.
The capacity to infer structural information from optical interactions thus owes
its origin to the fact that atoms and molecules have discrete levels, determined by
quantum mechanical principles, and given by the solutions of the Schrödinger
equation. Accordingly, since photon energies are related to frequency by
E = hv = hc/λ, (where h is Planck’s constant, c is the speed of light, ν is the optical
frequency and λ the wavelength), only certain frequencies of light can induce
changes from one level to another. Always, some quantum level other than the
stable (lowest energy) ground state is involved. The reason we can deduce so much
from molecular spectra is that the molecular transitions responsible are very strongly
linked with other characteristic properties of the matter—primarily their atomic
framework. This link between spectroscopic features and molecular constitution
provides the basis for a whole host of applications, familiar in our everyday world.
The field of photophysics is primarily concerned with the properties and
behaviour of molecular excited states. These are usually very short-lived—they
are almost always less than a millisecond in lifetime, and frequently present for times
on the scale of nanoseconds (10−9 s). What happens to the energy these states
contain, as they ‘relax’ to states of lower energy, is itself a rich field of study. By a
variety of mechanisms, energy that most often has been captured from light itself is
thereby available not only for subsequent re-emission as light; it can be transported
away over microscopic scales to initiate other processes. Most familiar and
important to us are the primary processes of vision, and of photosynthesis, each
of which entails a rich interplay of molecular photophysical mechanisms. We could
not find our way around the world, nor have any food to sustain our lives even if we
did, without photophysics.
In successive chapters, this book aims to build an increasingly rich and detailed
picture of how light is absorbed, what then happens to the energy acquired, and
what information the study of such processes can reveal. As indicated, we shall focus
on molecular matter which, with the exception of metals and minerals, accounts for
most of the matter we see in the everyday world. We shall start with the simplest
examples, including those few molecules we do not directly see—the molecular
gases. Once the general principles are established, we shall cover the field from the
lowest to the highest energy molecular transitions, from free rotations in a gas to
multiphoton absorption in complex biological media.
1-2
Molecular Photophysics and Spectroscopy (Second Edition)
Under ambient conditions most gas-phase samples are comprised of only atoms or
small molecules. Larger molecules, whose standard state phase is seldom gaseous,
can usually be present only in low concentrations—which is why trace detection
techniques are required for their observation. At any instant, the atoms or molecules
in a gas are well separated in space, and their random motions produce a temper-
ature-dependent Doppler broadening (due to motion-dependent frequency shifts) in
their absorption and emission spectra.
Whatever the nature of the sample, transitions that occur at visible and ultraviolet
wavelengths generally involve transitions between electronic states—for most gases,
photons of sufficient energy in fact have wavelengths just outside the visible
spectrum, in the near-ultraviolet region. Atomic spectra may exhibit a multitude
of lines, but those lines appear clearly separated and, compared to molecular
spectra, much narrower. In contrast, under normal ambient conditions, the visible/
ultraviolet spectra of condensed phase materials will most often exhibit broad,
overlapping features with no simply describable shape. And of course, in any
molecule, the electron sharing involved in chemical bonding between the atoms
means that the electronic properties are entirely different from the constituent
atoms; absorbing light will not excite individual atoms.
In regular atomic or simple ionic crystals (metals, semiconductors and salts, for
example) there is a uniformity of multiply-repeated unit cell structures that allow
bulk properties to emerge. But our visible world is mostly made up of condensed
phase materials: amorphous solids, liquids, solutions etc. Such substances most often
comprise relatively large molecules, each of which is electronically distinct; it is only
the non-bonding forces of attraction between them that prevents their complete
spontaneous evaporation. The interactions experienced by individual molecules can
vary much more widely from site to site, however, according to their local electronic
environment. Moreover, whereas every atom comprises a single nucleus, polyatomic
molecules can rotate and have internal vibrations whose number grows linearly with
the number of atoms per molecule—and these can also be involved in electronic
transitions, as we shall see in a later chapter.
There is one other, fundamental difference that plays into the linewidth of
electronic transitions in the condensed phase, and that is the uncertainty effect of
excited state lifetimes that are often in the sub-nanosecond regime. In gases, in which
the atoms or molecules are essentially isolated and infrequently collide, radiative
emission is a primary means of excited state decay. However, in the condensed
phase, the availability of a host of other internal decay and energy transfer processes
can dramatically shorten excited state lifetimes. Heisenberg’s time-energy
Uncertainty Principle then decrees that the associated transition frequencies become
less well-defined: they acquire greater breadth.
For all these and other reasons that we shall look at in due course, the electronic
spectra of most molecular materials display very broad bands with only their peak
values precisely quantifiable. But there is a flip side to this seemingly problematic
complexity; unique possibilities then arise to derive additional information from
other measurements, such as optical polarisation effects and excited state lifetimes.
And even more importantly, spectroscopy in other regions of the electromagnetic
1-3
Molecular Photophysics and Spectroscopy (Second Edition)
spectrum can directly identify individual molecules, and indeed reveal their atomic
composition and structure.
1-4
Molecular Photophysics and Spectroscopy (Second Edition)
Table 1.1. Wavelength λ, frequency ν, wavenumber ν, and photon energy E, for selected spectral regions, and
the types of atomic and molecular motions with which they are associated.
1-5
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 1.1. Variation of electric and magnetic fields in propagating light, the oscillations having a repeat
length that defines the wavelength. In free space the two fields are always mutually orthogonal: (a) plane
(linearly) polarised light, where the two fields are exactly in phase; (b) circularly polarised light (right-handed
case) where the fields have a π/2 radians phase difference.
frequency of the electric field. The form of the field oscillations is most often
represented as a plane wave, as shown in figure 1.1(a), but it may for example take
the form of a helix (of either left- or right-handedness) as in figure 1.1(b). These two
cases are known as plane polarisation and circular polarisation, respectively.
Circular polarisations convey angular momentum, ±ℏ per photon according to
handedness (the positive sign for left when looking down the propagation axis of the
incoming wave), thereby revealing another key property of photons; the quantisation
of spin. Many of the most important properties of light, especially in the realm of
spectroscopy, owe their origin to the fact that photons have an intrinsic unit spin,
S = 1. This categorises photons in the class of elementary particles known as bosons,
i.e. particles of integer spin whose collective behaviour is subject to Bose–Einstein
statistics. This contrasts strongly with the electrons with which they interact, whose
own half-integer spin renders them fermions (being subject to Fermi–Dirac statistics).
One immediately important aspect of spin is that a collection of bosons can exist
in the same quantum state—and photons can therefore produce highly coherent,
monochromatic and directional beams, which is of course a major feature of lasers.
(Conversely, not even two electrons can exist in the same quantum state, as the
periodic table makes very apparent.) We shall encounter numerous other evidences
of photon spin in the course of the following chapters, because it is a feature that
underpins almost every one of the selection rules in spectroscopy—rules that
determine whether or not a quantum transition can take place, even if the photon
has exactly the right energy to match a gap in molecular energy levels. These rules
still operate in the more common case where light is plane-, rather than circularly
polarised, since plane polarisation acts as a superposition of opposite circular
polarisations.
To round off the most significant properties of photons it is worth finally observing
that despite having zero mass, they do possess linear momentum, p = h/λ per photon.
This neatly exhibits both the quantum and relativistic nature of the particle, since
energy, momentum and ‘rest mass’ m0 are linked by the relation E 2 = p2 c 2 + m 02c 4 .
With m0 = 0 and relation (1.1), the result p = h/λ immediately follows. However, linear
momentum has very little direct influence in most forms of spectroscopy; it mostly
comes into its own in the sphere of optical manipulation of atoms and molecules; for
example, trapping samples in space and reducing their kinetic energies to near zero.
1-6
IOP Publishing
Chapter 2
Atoms and molecules: developing principles of
electronic structure
state: that is the energy needed for ionization; removal of the electron leaving behind
a proton H+. This energy is often expressed in electron-volt units, ~13.6 eV, where
1 eV is around 8066 cm−1. It should be noted that this lower limit for ionization is
defined as where the kinetic energy of the freed electron is zero. The energies of all
H atom states are referenced to this zero, and therefore have negative (binding)
energies. In this regard the ground state is the most bound state of the system.
Spectroscopy does not involve measuring the absolute energies of states, rather
the energy differences for transitions between them. Hence, it makes more sense
to pick the ground state as a ‘zero’ of energy and measure excited state energies
relative to that level. Thus, when an atom (or molecule) absorbs light it goes from
the lower state (usually the ground state) to an excited state that is higher in
energy. Conversely, when an atom (or molecule) transitions from an excited state
to a lower state by emission or some other deactivation mechanism, it loses
energy. Regardless, the transition energies are reported as positive quantity,
ΔE = Eexcited state − Elower state. In spectroscopic studies of the light emitted in a
discharge lamp containing elemental hydrogen (H2) one observes a series of
atomic H-atom lines (called transitions) originating from high n-levels and
terminating on specific lower n-states. Here, the energy level differences between
H atom levels with quantum numbers labelled n1 and n2 can be written as:
⎛1 1 ⎞
ΔE = E n2 − E n1 = RH⎜ 2 − 2 ⎟ . (2.2)
⎝ n1 n2 ⎠
Transitions terminating on n1 = 1 form the Lyman series, which lies mainly in the
extreme ultraviolet (λ < 100 nm), while the group of transitions terminating on
n1 = 2 and n1 = 3 are called the Balmer and Paschen series, respectively. The former
lies in the visible range while the latter is observed in the infrared. There are further
series at even longer wavelengths. One important point here is that the energy levels
in hydrogen (and all other atoms as well) are not equally spaced. Almost half the
energy required to ionize H—to remove its electron completely—is required to excite
the atom from its ground state just to the first excited state, n = 1 to n = 2. This is
comparable to the energy to excite transitions from n = 2 up to the ionization limit,
even though there are an infinite number of n-states in this region.
Before delving deeper into the quantum features of the H atom, some comments
regarding spectral lines is warranted. In describing the precise electronic structure of
atoms and simple molecules, more intricate quantum calculations will deliver
extremely accurate results for the energy levels of the ground and most excited
states. Indeed, advanced computational techniques based on quantum mechanical
models can give very precise figures for the excited state energies of even quite large
molecules. However, it is usually only gas phase samples that exhibit sharp lines in
their absorption or emission spectra. Hydrogen atom spectra are indeed composed
of sharp (narrow linewidth) features, corresponding to transitions between different
electronic states—but even here, or for other atoms, how ‘sharp’ a line is will depend
on the sample conditions, the presence or absence of electric and magnetic fields, the
light source, and the detection system being used. Many details will depend on the
2-2
Molecular Photophysics and Spectroscopy (Second Edition)
experimental setup, and of course this is one of the key reasons for the wide
application of laser instrumentation: to achieve the narrowest possible linewidth for
the light source.
In addition to the generic lineshape issues relating to the nature of the sample,
discussed in chapter 1, a host of other line-broadening mechanisms generally contribute
to an effective linewidth for any spectral feature. As we observed, the effects of
uncertainty line-broadening due to finite excited state lifetimes are significant; so too are
the broadening due to the Doppler effect of random particle motions, and through
stochastic perturbations produced by inter-particle collisions. These fall into two
categories of line-broadening relating to intrinsic and motion-dependent origins, known
as homogeneous and inhomogeneous respectively: each is generally associated with a
specific form of lineshape function. Homogeneous lifetime broadening, whose magni-
tude depends for each transition on the specific excited state, produces a frequency
response with a Lorentzian lineshape, f (ν ) ∼ 1/[(ν − ν0 )2 + ( 12 Δν )2], where v0 is the
resonance frequency and Δv is the FWHM (full-width at half-maximum) linewidth. On
the other hand, inhomogeneous collisional and/or Doppler broadening give rise to a
Gaussian shape, f (ν ) ∼ exp[−4 ln 2((ν − ν0 )/Δν )2]. When both phenomena are
present, the two lineshapes can be convoluted to yield a Voigt profile.
Returning now to the H atom, theory reveals that there is another quantum
number, symbol ℓ, known as the azimuthal quantum number, that signifies the
orbital angular momentum of the electron as it travels around the nucleus. Although
it is not determinable from any of the prominent frequencies in the spectrum—note
that it does not feature at all in equation (2.2)—it is spectroscopically very
important, as we shall see. Specifically, it is related to the square of quantised
orbital angular momentum of the electron in each nℓ orbital in units of ħ2, and to the
shape of the orbital itself. However, that shape, whose boundaries define a volume of
space around the nucleus where one is most likely to find the electron if measured,
does not tell us about the actual trajectory of the electron within that space. Indeed,
quantum mechanically the idea of a trajectory is a meaningless concept. It can be
shown that the value of the square of the angular momentum is = ℓ(ℓ + 1) ħ2.
The quantum number ℓ is also quantised and its possible values are related to n.
For a given n value, ℓ can take on the values 0, 1, …, n − 1, such that for n = 1, ℓ = 0
alone. Orbitals with ℓ = 0, 1, 2 and 3 are labelled s, p, d and f, respectively: the
electronic ground state is therefore labelled 1 s. Similarly, for n = 2, 3 and 4, there are
2, 3 and 4 possible states, as listed in table 2.1. In principle there are orbitals with
ℓ > 4 (g-orbitals, h-orbitals, etc.) but they are usually not considered because the
highest ground state orbitals in the periodic table are the 5f states of the actinides.
Since the hydrogen atom energy levels given by equation (2.1) are independent of
ℓ, i.e., different states for a given n have the same energy, they are called degenerate.
The basic shapes of the ℓ-orbitals, determined as the angular part of the wave-
functions in the solution of the Schrödinger equation, are similar for every n, except
that as n increases, they grow in volume. In contrast, as ℓ increases they exhibit an
increasing number of radial and angular nodes (spatial positions where one will
never measure the electron).
2-3
Molecular Photophysics and Spectroscopy (Second Edition)
Table 2.1. Observed orbitals and n, ℓ quantum numbers for the first four energy
states of H.
N ℓ State labels nℓ
1 0 1s
2 0, 1 2s, 2p
3 0, 1, 2 3s, 3p, 3d
4 0, 1, 2, 3 4s, 4p, 4d, 4f
Figure 2.1. Shapes of the 1 s, 2p and 3d orbitals of the hydrogen atom. Shaded lobes indicate an opposite sign for
the angular wavefunction, compared to unshaded lobes. The subscripts dependent on (x, y, z) indicate the
directions in which the orbitals point. There is no directional preference for s-orbitals, which have a characteristic
round shape. The 4f-orbitals have more complex structures and are not presented. Note that s- and d-orbitals do
not change sign by inversion through the nucleus; that is going from (x, y, z) to (−x, −y, −z)—they are of even
spatial parity: however, the p-orbitals do—they have odd parity.
Several of the most important orbital shapes are shown in figure 2.1, in which
shading indicates the sign of the wavefunction—an aspect that proves to be
important in the rules governing the types of optical transition. Clearly, some are
symmetric about the nucleus, while others are not; this behaviour is known as spatial
parity. It can be shown that the parity eigenvalue for one-electron systems is given
by (−1)ℓ. It follows that s and d orbitals are even with respect to inversion (parity +1),
2-4
Molecular Photophysics and Spectroscopy (Second Edition)
and are therefore labelled as g-states, where g stands for gerade, the German word
for even. Similarly, p and f states are odd with respect to inversion (parity −1) and
are labelled as u-states, deriving from ungerade, the German for odd. It is worth
recalling, however, that the modulus square of the wavefunction is proportional
to the charge density distribution in space—and this of course has a constant
negative sign.
The standard designation of Cartesian axes, used in describing the shape and
orientations of atomic orbitals in figure 2.1, invites more comment. There are no
unique directions in empty three-dimensional space, and atoms themselves have
spherical symmetry—so any such labelling has to be arbitrary. At least some of the
degeneracies therefore become obvious. However, physically imposing a space axis
by placing the atom within a constant magnetic field breaks those degeneracies and
leads to line splittings in the spectrum. We can then deduce that there is a third
quantum number mℓ, accordingly termed the magnetic quantum number, which tells
us for a given n the number of orbitals of a given ℓ. Formally, mℓ ranges from −ℓ to +ℓ
in integer steps, leading to a total of 2ℓ + 1 degenerate states. Thus, each n-level, for
n ⩾ 1 has one s-orbital, while for every n ⩾ 2 there are also three p-orbitals, and
so on.
There is one further, additional quantum number for the electron associated with its
spin s. It is related to the square of an intrinsic spin angular momentum of the electron.
Although the quantum behaviour of the electron spin resembles that of its orbital
motion, it does not in fact depend on spatial coordinates. Spin is something therefore
that is usually added on in an ad hoc way to the solutions described above (although it
does naturally emerge using a relativistic formulation of quantum mechanics). Similar
to ℓ, the square of the spin angular momentum is given by s(s + 1)ħ2 and every electron
has s = ½. The number of spin states for s = ½ is labelled by the spin quantum number
ms; every electron has only two allowed spin states, ms = + ½ (spin up) and ms = −½
(spin down). In many chemistry books the spin up function is denoted by α, and the spin
down function likewise β. While the use of the spin terms ‘up’ and ‘down’ is pervasive,
it is a little unfortunate because it implies some specific spatial orientation, which is not
the case.
There is one further spectroscopic feature to introduce before we move on from
hydrogen to more complex systems. In any light absorption or emission process it
does not, in general, simply suffice for the energy gap between the initial and final
states of the atom to be matched by the energy of a photon. In fact, additional
criteria have to be satisfied; it is necessary to conform to selection rules. This is an
aspect of special significance for not only for optical transitions in atoms but also in
molecules, where considerations of parity, the forms of charge distribution, and
angular momentum all come into play.
In both absorption and emission, the strongest transitions correspond to a dipolar
form of coupling with the electric field of the radiation. More specifically, the
property that determines the strength of such interactions transition electric dipole,
through which an atom or molecule interacts with the electric field of optical
radiation. It is important to emphasise the quantum nature of this parameter, a
2-5
Molecular Photophysics and Spectroscopy (Second Edition)
where the prime indicates an upper state. If the wavefunction contains the complex
number i = −1 then one must also use the complex conjugate of upper state
wavefunction (where i → −i ), as indicated by the asterisk. It can be shown that there
are no restrictions on transitions in H between different n-levels, but the change in
the azimuthal quantum number must satisfy Δℓ = ±1. This is sometimes referred to
as the Laporte selection rule. Ultimately, this is related to the angular momentum
properties of light. As we have seen, fundamental states of a photon are either left-
hand circularly polarised with angular momentum +ħ, or right-hand circularly
polarised with angular momentum −ħ. It follows that when an H atom absorbs one
photon the change in the angular momentum of the atom must either be Δℓ = +1 or −1
due to conservation of angular momentum. Hence, the first allowed transition in H is
from the 1 s → 2p. However, the transition 1 s →2 s is said to be forbidden. This
principle also satisfies parity, as the electric dipole operator has odd parity; the
integrand in equation (2.3) is therefore an even quantity when transitions are
allowed.
While light can interact with other moments such as the magnetic dipole and/or
the electric quadrupole, the resultant transitions are usually weaker than the electric
dipole transitions by several orders of magnitude and only become dominant in
spectral regions where electric dipole transitions are not observed.
We can now wrap up our consideration of hydrogen by noting that equation (2.3)
can be adapted to include electron spin, as;
2-6
Molecular Photophysics and Spectroscopy (Second Edition)
2-7
Molecular Photophysics and Spectroscopy (Second Edition)
indicate how many electrons are in the labelled nℓ shell. Essentially, as we move from
one atom to another, through the periodic table, electrons move into shells of
progressively lower binding energy.
Two other important principles arise, however, which the electron configurations
alone do not identify. The first is the Pauli Exclusion Principle, which specifies that
no two electrons in any orbital can have the same set of electronic quantum
numbers. Put another way, two electrons at most can occupy any specific (n, ℓ, mℓ)
orbital, and they must have opposite spin quantum numbers, ms; that is, one will be
spin up and the other, spin down. The second principle is Hund’s Rule, which states
that it is more energetically favourable to first fill a set of (n, ℓ, mℓ) orbitals singly
with electrons having the same spin orientation, before the accommodation of
further electrons results in any of those orbitals becoming doubly occupied and
spin-paired.
For a given ground or excited state electronic configuration the total orbital
angular momentum L of a many-electron system can be calculated from the
quantised vector sum L = Σ⃗i ℓi . As an example, consider two orbital angular
momenta ℓ1 and ℓ2. The permitted vector sums give rise to a resultant whose
magnitude lies in the range L = ℓ1 + ℓ2 → ℓ1 − ℓ2 in integer steps. Repeating the
procedure for the next electron, and so on, it is easy to see that a single multielectron
configuration can lead to many possible values of L. As each ℓ is an integer, the
resultant L values are also integers. States of different L are called terms, and they
are represented in analogy to H atom states with capitalised labels S, P, D, F, …, for
L = 0, 1, 2, 3, …, respectively. The ML components which range from –L to +L in
integers can be resolved in a magnetic field, which lifts their degeneracy.
Since each electron has a spin s = ½, one can also derive the total spin of the
system S using the vector sum S = Σ⃗i si whose permitted quantum magnitude is
similarly limited to the range s1 + s2 → s1 − s2 , again in integer steps. If the
number of electrons in the atom is even then the total S will be an integer; an odd
number of electrons leads to multiples of ½. The resultant MS components, ranging
from –S to +S in integer steps are resolvable in an applied magnetic field.
It is useful to define the spin multiplicity as the value of (2 s + 1). Two of the most
common cases of spin multiplicity also commonly arise for organic molecules. When
S = 0 (all electrons are spin-paired) the spin multiplicity = 1 and the term is referred
to as a singlet state. Similarly, if S =1 (two of the electrons in the same set of n, ℓ
orbitals have the same spin orientation), then (2 s + 1) = 3. The resultant term is
called a triplet state.
The orbital motion of the electrons in an atom (for L ≠ 0) creates a magnetic field
which can itself interact with magnetic moment of the total spin of the atom (again
provided S ≠ 0). This so-called spin–orbit interaction can couple the total L and S
angular momenta together to create different total angular momenta which are
labelled J. For light atoms with a relatively small number of electrons (typically < 40)
the L and S values considered above are summed vectorially to yield a total angular
momentum of magnitudes J = (L + S) → (L−S); a term with specific L and S will
often give rise to several values of J. In this approach to understanding the interaction
2-8
Molecular Photophysics and Spectroscopy (Second Edition)
of spin and orbital angular momenta, called the Russell–Saunders (or L−S) coupling
scheme, the resultant terms are labelled by 2s+1LJ, and their individual energies differ.
Once again, the energies of individual MJ components, ranging from −J to +J in
integer steps, can be split by a magnetic field.
There is one further subtlety that should be noted: there may be restrictions on S
(and therefore J) due to the Pauli Exclusion Principle for terms derived from
electrons within an nℓ configuration versus terms derived from electrons in different
nℓ states. For example, a 1s2 configuration yields only a 1S0 term (the two electrons
must be spin-paired) while a 1s12s1 configuration allows both 1S0 and 3S1 terms (the
two electrons may have either the same or opposite spin orientations).
As we move on through the periodic table and examine atoms with increasingly
high atomic numbers, and large numbers of electrons, a different kind of behaviour
becomes apparent. One factor that becomes more and more important is a direct
coupling of spin and orbital angular momenta—a relativistic effect whose influence
grows with the fourth power of Z. As the atomic number increases, the resultant
spatial confinement of a large number of circulating electrons about the nucleus will
lead (by the Heisenberg Uncertainty Principle) to them having increasingly high
momenta—so it makes sense that relativistic effects become more significant.
When the number of electrons in an atom is large, the relatively larger spin–orbit
interaction will first couple the individual ℓi and si angular momenta to create
individual total angular momenta ji. Again, the ji values can be derived by adding
each ℓi and si vectorially. The total J angular momenta are found by first vectorially
adding individual values, say j1 and j2, to obtain intermediate J values. Those values
are then added vectorially to another, say j3, and so on successively until all the
electrons have been accounted for. Here, L and S are not well defined, and what are
now described as j–j coupling terms are labelled simply by J values alone—which can
have either integer or half-integer values, depending on whether there is an even or
odd total number of electrons. The number of J-states and their numerical labels
found this way will be the same as those derived using Russell–Saunders coupling,
but the actual energy ordering of the terms may be significantly different.
The parity of a multielectron atom is determined by the sum of individual electron
orbital angular momenta in the electronic configuration for a particular state; that is
by( −1)∑i ℓi . Accordingly, a term may be even (g) or odd (u) regardless of L. The
selection rules for transitions between Russell–Saunders terms, and between j–j
coupling terms are provided in table 2.2.
Again, the forbidden transitions reflect the conservation of angular momentum
requirement for the absorption (or emission) of a photon by the atom.
2-9
Molecular Photophysics and Spectroscopy (Second Edition)
Table 2.2. Selection rules for electric dipole transitions between Russell–Saunders terms, and between j–j
coupling terms.
g↔u g↔u
ΔS = 0
ΔL = 0, ±1 but L = 0 → L = 0 forbidden
ΔJ = 0, ±1 but J = 0 → J = 0 forbidden ΔJ = 0, ±1 but J = 0 → J = 0 forbidden
ΔMJ = 0, ±1, but ΔMJ = 0 forbidden if ΔJ = 0
whole molecule can take place. However, the fundamental quantum transitions
associated with these very important forms of motion take place at much longer
wavelengths, and they have a very different character; we shall return to consider
them in later chapters.
An obvious difference between atoms and diatomic molecules is that the latter
have cylindrical symmetry, the internuclear bond introducing a unique, physically
distinguishable axis. Accordingly, the electronic states of diatomics are usually
labelled by the absolute magnitude of the component of orbital angular momentum
along their bond axis, Λ. By analogy with atoms, such states are labelled by the
corresponding capital Greek letters: Σ, Π, Δ, Φ, … when Λ = 0, 1, 2, 3…,
respectively. Here the orbital motion can be thought of as clockwise or counter-
clockwise rotation of the electron about the bond axis. Therefore, states with Λ ⩾ 1
are doubly degenerate since to a first approximation the two motions are indis-
tinguishable in energy. The case of Σ-orbitals (Λ = 0) requires special attention with
respect to their symmetry on reflection in any plane containing the molecular axis.
This reflection plane (and the operation of reflection) is denoted by σv—the first of
several molecular symmetry properties to be discussed in more detail in the next
chapter. Based on this mirror symmetry, Σ states that are symmetric on reflection are
labelled Σ+while those that change sign are denoted as Σ−. Such superscripts are not
required for states with Λ > 0 because of their double degeneracy.
Molecular terms are typically denoted by 2s+1ΛΩ, where Ω = (Λ + Σ) is the
component of total angular momentum along the molecular axis. Note the italics: Σ
does not mean Λ = 0: it is instead the standard symbol for the component of the
electron spin S along the bond axis, which ranges from +S to −S in integer steps.
The selection rules here are ΔΛ = 0, ±1, ΔS = 0, ± ↔ ±, g ↔ u ; the latter of
these applies only when both atoms comprising the molecule are the same—i.e. for
homonuclear diatomics. When the two atoms are different—heteronuclear diatomics—
the g and u labels are no longer meaningful.
2-10
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 2.2. (a) Molecular orbitals and energies (ascending on the vertical scale) for a homonuclear diatomic
A—A, formed from the constructive and destructive addition of the electron wavefunctions for two ns orbitals,
one on each atomic centre. (b) Molecular orbitals for a hetermuclear diatomic A—B from the constructive and
destructive addition of two ns orbitals, one on each atomic centre. Inversion symmetry denoted by g/u is only
meaningful for the homonuclear case.
almost all molecules are held together by chemical bonds between neighbouring
atoms. This does not mean that the features in a molecular spectrum can in general
be attributed to individual bonds, for both electronic and nuclear motions
commonly extend beyond individual bonds, sometimes to a very substantial effect.
But certain bond structures do exhibit characteristic features—the carbonyl unit
C=O where the C and O atoms are held together by a double bond (see below) is a
prime example—and there the principles that apply to simple diatomics do lend
useful insights. Later, we shall encounter molecules whose electronic orbitals are
more substantially delocalised.
A chemical bond is formed when the atomic orbitals (AOs) on the two atom
centres overlap to form molecular orbitals (MOs). A simplistic expectation that
sharing electrons between two nuclei in a bond will lower the energy, relative to that
for the separated atoms, which explains why most molecules are stable entities.
Forming linear combinations of separate atomic orbitals can be done in two ways,
with different energies. Consider figure 2.2(a), where ns orbitals on each atomic
centre are combined to form a homonuclear diatomic molecule A—A. Similarly, in
figure 2.2(b), the ns orbitals associated with the atoms making up a heteronuclear
diatomic A—B are combined. For convenience we take the horizontal axis between
the atoms to be the bond axis. In both cases, adding the orbitals constructively (that
is, combining them with the same +/− phase) produces significant electron density
between the atoms. This is a bonding MO and is labelled nsσg for A—A and nsσ for
A—B. Accordingly, these are termed sigma bonds. Recall that the g/u labels apply if
the two atoms making up the diatomic are the same. Otherwise, those labels are
omitted. The shared electron density between the atoms lowers the energy, as
expected.
When the orbitals are added destructively however, (i.e. with opposite phase) the
resultant electron density between the atoms necessarily exhibits a nodal plane that
2-11
Molecular Photophysics and Spectroscopy (Second Edition)
bisects the bond axis. This is not conducive for bonding; it signifies that electrons
cannot move freely between the atoms. The resultant MO is in fact higher in energy
than that for the separated atoms, and is therefore termed anti-bonding. These states
are labelled nsσu* and nsσ* for A—A and A—B, respectively where the asterisk
superscript * denotes an anti-bonding configuration. The main difference for a
heteronuclear diatomic, compared to the homonuclear case, is that the relative
energies of the ns orbitals are different and therefore the shapes of the resultant MOs
are not symmetrical. As drawn in figure 2.2(b) the bonding MO has more B-atom
character, while the anti-bonding MO has more A-atom character—reflecting the
higher AO energy of the free atom A. An example of a molecule which only has
σ bonds is tetrahedrally-shaped methane, CH4 (figure 2.3). The sigma bond electron
density is localised between the C and H atoms in each C–H bond.
These principles can be extended to MOs formed from combinations of other
atomic orbitals having different n and ℓ. For example, in figure 2.4(a) a bonding
orbital is formed by constructively adding the 2pz orbitals on, say, two atoms to
form a 2pσg molecular orbital (it is conventional to assign the z-axis to the
internuclear bond). The other combination forms a 2pσu* anti-bonding orbital
with a nodal plane between the two atoms, as in figure 2.4(b).
Since there are three p-orbitals for every n ⩾ 2, MOs can also be formed using the
npx and/or npy orbitals on each atom. In figure 2.4(c) shows constructive formation
of electron density from either the 2px and or 2py atomic orbitals, on each side of the
bond axis (above and below, as shown) but not directly between the atoms. This MO
is represented as 2pπu; it has a bonding nature, termed a pi bond, though it is less
strong than 2pσg. The destructive combination in figure 2.4(d) with a nodal plane
between the atoms is anti-bonding in nature: the resultant MO is designated as
2pπg*.
A double bond is formed when the p-orbitals form one σ-bond and one π-bond.
The remaining atomic p-orbital not involved in molecular bonding is called a non-
bonding orbital, denoted by n. (This is the standard notation for a type of MO,
contrasting with σ and π; beware confusion with the principal quantum number n.)
Figure 2.3. Representations of methane. (a) Bonding structure in a conventional organic chemistry 3D
representation (dashed line indicating a bond heading ‘backwards’ relative to the plane of the page, the
thickened line a ‘forwards’ bond relative to that same plane, the straight line a bond in that plane); (b) atom
positioning, carbon at the centre of a regular tetrahedron of hydrogen atoms (bonds not shown).
2-12
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 2.4. (a) 2pσg bonding molecular orbital formed from two 2pz atomic orbitals; (b) 2pσu* anti-bonding
molecular orbital formed from 2pz orbitals; (c) 2pπu bonding molecular orbital from either two 2px or two 2py
atomic orbitals; (d) 2pπg* anti-bonding orbital formed from either two 2px or two 2py atomic orbitals. The red
dashed line indicates the anti-bonding nodal plane between the atoms.
Figure 2.5. Conventional depiction of methanal, the pair of parallel lines denoting a double bond.
Similarly, p-orbitals forming one σ-bond and two π-bonds yield a triple bond.
Double and triple bonds are commonly formed in many organic molecules.
An example of a planar molecule with a double bond, shown in figure 2.5, is
methanal (CH2O; informally known as formaldehyde). This molecule has a double
bond composed of one σ-bond and one π-bond located between the C and O atoms.
The π-system does not extend to the C–H bonds. While sigma and pi bonds can be
formed from combinations of s, p and d atomic orbitals, the combinations described
above are the most common in organic molecules—which represent by far the
largest number of molecules we encounter in the everyday world.
We can now begin to appreciate some of the key elements of electronic
spectroscopy, involving the absorption of photons in the UV/visible region.
Assuming selection rules allow (a matter we pursue in the next chapter), it is usually
the transition which bridges the smallest energy gap that are the most probable. As
shown in figure 2.6, the strongest electronic transitions in organic molecules will take
place from the highest occupied molecular orbital (HOMO; usually the π-system if
there are multiple bonds) to the lowest unoccupied molecular orbital (LUMO) which
2-13
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 2.6. Energy level diagram (energy ascending on the vertical scale) for a generic organic molecule,
indicating the possible types of electronic transitions, and their relative energy ordering.
are typically anti-bonding in nature (σ* or π*). However, the type of transitions
observed will depend on the exciting light source. In general the energies E of the most
common transitions are ordered: E(π* ← n) < E(π* ← π) < E(σ* ← n) < E(σ* ← σ).
In terms of the absorption wavelength, σ* ← σ and σ* ← n transitions are typically
strongest for λ < 200 nm, while π* ← π and π* ← n dominate in the UV/visible range.
2-14
IOP Publishing
Chapter 3
Polyatomic molecules: orbitals, symmetry and
group theory
3-2
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 3.2. The left-hand depiction of benzene shows six p-orbitals, one per carbon atom, which combine to
form the molecular orbital shown on the right, with electron density above and below the ring plane.
3-3
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 3.3. Symmetry elements of the benzene molecule. The C6- and S6-axes coincide and are coming out of
the horizontal plane of the figure indicated by the green box (σh). The red lines which correspond to the three
C2’ axes perpendicular to C6 also indicate the edges of three σν planes coming out of the page, containing C6
and one of the C2’ axes. The blue lines correspond to three C2’ axes perpendicular to C6, and the edges of the
three σd planes coming out of the page, containing the C6 and one of the C2’ axes.
abelian. As an example, table 3.1 shows products of the symmetry operations for an
equilateral triangle (depicted in figure 3.4), i.e. {E, C31, C3−1, σA, σB, σC}.
It is important to note here that this group exemplifies four key properties that
apply to all proper groups, namely that they must satisfy the following:
1. Closure. No new symmetry elements are generated through multiplication.
2. The group contains E.
3. The elements are associative; that is (OiOj)Ok = Oi(OjOk).
4. Each element has an inverse such that the product of Oi and its inverse
yields E.
Each row and column of a group multiplication table contains every operation once,
similar to a Sudoku puzzle. Given this information, there are several algorithms one
can use to establish which unique combination of symmetry elements, i.e. which
point group a molecule belongs to. These are named using Schoenflies notation, and
are presented in table 3.2. The term ‘point group’ indicates that only the relative
point positions of atoms within the molecule are significant. For crystalline solids,
whose symmetry additionally involves the relative spatial translations of unit cells,
space groups are employed. For regular crystals (a great many of which are of
atomic or ionic rather than molecular constitution) there are 230 distinct types of
space group; their study is beyond out present remit, where we are primarily
concerned with the optical properties of individual molecules.
3-4
Table 3.1. Multiplication table for the group symmetry elements of an equilateral triangle.
3-5
C31E = C31 C31C31 = C3−1 C31C3−1 = E C31σA = σC C31σB = σA C31σC = σB
C3–1 C3−1E = C3−1 C3−1C31 = E C3−1C3−1 = C31 C3−1σA = σB C3−1σB = σC C3−1σC = σA
σA σA E = σA σAC31 = σB σAC3−1 = σC σA σA = E σA σB = C31 σA σC = C3−1
σB σ BE = σ B σBC31 = σC σBC3−1 = σA σB σA = C3−1 σB σB = E σB σC = C31
σC σC E = σC σCC31 = σA σCC3−1 = σB σC σA = C31 σC σB = C3−1 σC σC = E
Molecular Photophysics and Spectroscopy (Second Edition)
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 3.4. Symmetry elements of an equilateral triangle with its C3-axis in the z-direction out of the plane of
the page. The red lines denote the vertical planes containing C3.
of the field of linear algebra, but it is sufficient here simply to note that matrices of
the correct dimensions can be added, subtracted and multiplied. Many square
matrices, but not all, have an inverse O−1 such that the product inverse
O−1O = OO−1 = I, where I is the Identity matrix whose diagonal elements equal
1 and all others are 0. The only definition we need to introduce is the character of
matrix, χ, which is the sum of diagonal elements: χ = ∑i oii . Another common name
for this character is the trace; in group theory the former designation is more
prevalent.
Matrices are introduced because symmetry operations can be written in matrix
form. A logical basis set to construct a matrix representation of a symmetry
operation Γ(O) comprises the unit vectors pointing in the x-, y-, and z-directions:
( )
iˆ, jˆ, kˆ , for example:
⎛1 0 0 ⎞ ⎛− 1 0 0 ⎞
Γ(E) = ⎜⎜ 0 1 0 ⎟⎟ ; Γ(i) = ⎜⎜ 0 − 1 0 ⎟⎟ .
⎝0 0 1 ⎠ ⎝ 0 0 − 1⎠
On the left here is the 3 × 3 identity matrix, and on the right the matrix for spatial
inversion; their characters are 3 and −3, respectively. Matrices for all other
operations can be similarly derived.
One of the powerful aspects of group theory is that unit vectors are not the only
basis functions that can be used to derive matrix representations of symmetry
operations. For example, one might use as bases the directions of the bonds in a
molecule, the inter-bond angles, the orbitals on each atom, or so on, depending on
the type of problem being considered. Some of these choices could in principle lead
to very large matrix representations indeed. Consider again, for example, placing
three Cartesian unit vectors at each vertex of the equilateral triangle shown in
figure 2.4. The dimension of each resultant matrix representation, describing how all
9 of those vectors ‘move’ at the same time under a symmetry operation, would
already be 9 × 9.
3-6
Molecular Photophysics and Spectroscopy (Second Edition)
Table 3.2. Major point groups collected into categories, and the symmetry operations needful for each. Point
groups labelled ‘uncommon’ are not representative of any common molecule.
Rotational C groups
Cn E, Cn Abelian
Cnh E, Cn, σh Even n: operations include E, Cn and
its repetition, σh, i, various Sn.
Odd n: operations include E, Cn and
its repetition, σh, various Sn.
Cnν E, Cn, nσν
Dihedral D groups
Dn E, Cn, nC2’
Dnh E, Cn, nC2’, σh, nσν Even n: the nσ planes divide into one
set of n/2 σν planes and one set of
n/2 σd planes.
Dnd E, Cn, nC2’, nσd Odd n: an inversion operation i will
exist.
For all n, several S2n symmetry
operations are possible.
Linear point groups
C∞v E, 2C∞, ∞σν
D∞h E, 2C∞, ∞σν, σh, ∞C2’, i
S groups
Sn E, Sn Uncommon
Cubic groups
Td E, 8C3, 3C2, 6S4, 6σd Tetrahedral point group
Oh E, 8C3, 6C2, 6C4, 3C2, i, 6S4, 8S6, Octahedral point group
3σh, 6σd
Ih E, 12C5, 12C52, 20C3 15C2, i, 12S101, Icosahedral point group
12S103, 20S6, 15σd
Additional cubic groups
T E, 4C3, 4C32, 3 C2 Uncommon
Th E, 4C3, 4C32, 3 C2, i, 4S6, 4S65, 3σh Uncommon
O E, 8C3, 6C2, 6C4, 3C2 = C42 Uncommon
I E, 12C5, 12C52, 20C3 15C2 Uncommon
3-7
Molecular Photophysics and Spectroscopy (Second Edition)
The set of matrices for all of the symmetry operations applicable to a specific
molecule, regardless of their dimensions, also form a mathematical group. It can
hugely simplify applications if the matrix representation is expressed in terms of the
smallest dimensional matrix representations possible. These are known as irreducible
representations (IRs), because they can be reduced no further; any other representa-
tion is said to be reducible. The irreducible representations are 1 × 1 (which means its
character is the matrix itself!), 2 × 2 or 3 × 3, but no larger—except in the very rare
case of icosahedral symmetry. We can simplify further by asking how the characters
χ of these matrix IRs behave under symmetry operations. The advantage of
considering characters over the any explicit matrix representation is they are matrix
invariants; the individual matrix elements may vary according to the choice of
coordinate system, but the characters do not. The labels of the IRs are called
Mulliken symbols:
• A one-dimensional IR is labelled A if it is symmetric under a Cn rotation; that
is, χ(Cn) = +1. Otherwise, if χ(Cn) = −1 the IR is labelled B.
• Subscripts 1 and 2 are used depending on whether the IR is symmetric or anti-
symmetric, χ = +1 or −1, respectively, under any C2’ rotation or σν reflection.
• Prime or double prime superscripts indicate whether the IR is symmetric or
anti-symmetric under any σh reflection.
• g and u subscripts, (introduced in chapter 2), indicate whether the IR is
symmetric or anti-symmetric with respect to inversion, i, if an inversion centre
exists.
• Two-dimensional and three-dimensional representations are labelled E and T,
respectively.
In the last of these rules, the labels E and T should not be confused with the
‘do-nothing’ symmetry operation E, or with T groups. In each of these cases,
ancillary labels (primes, subscripts etc) may also be applied, just as with the A and B
representations.
Linear systems are often described with an alternative set of primary symbols.
One-dimensional A representations are in this case labelled Σ, and there is an
unlimited range of two-dimensional (i.e. doubly degenerate) E representations, now
labelled Π, Δ, Φ etc. The virtue of this system is that when applied to linear
molecules such as diatomics, these designations correspond with the magnitude of
the z-component of the angular momentum along the symmetry axis, Λ. As we
observed in section 2.3, terms labelled Σ, Π, Δ, and Φ signify Λ = 0, 1, 2, and 3.
However, the same Mulliken subscripts and superscripts as those described above
still apply, if appropriate.
Let us look in detail at a more typical example, the character table for the
C3v point group, shown in table 3.3. The ammonia molecule, NH3, is a common
example of this kind. The left-hand column of the table lists all the IRs associated
with the point group specified in the upper left-hand corner. The transformation of
the characters for each class of operations is listed in the right adjacent columns.
Commonly such tables include additional information on the right-hand side. Here,
for example, we are told for example how components of x, y and z vectors transform.
3-8
Molecular Photophysics and Spectroscopy (Second Edition)
Table 3.3. Character table for the Schoenflies point group C3ν.
Linear functions,
C3v E 2C3(z) 3σν rotations Quadratic functions Cubic functions
A1 +1 +1 +1 z x2 + y2 , z 2 z 3, x(x 2 − 3y 2 ), z (x 2 + y 2 )
A2 +1 +1 −1 Rz − y(3x 2 − y 2 )
E +2 −1 0 (x , y ) (Rx , Ry) (x 2 − y 2 , xy ) (xz, yz ) (xz2, yz2 )[xyz , z (x 2 − y 2 )]
[x(x 2 + y 2 ), y(x 2 + y 2 )]
Figure 3.5. Rotations about the three Cartesian axes. In the point group C3v, Rz transforms as A2 while
(Rx, RY) are doubly degenerate and transform as E.
In C3v, z transforms as A1, while (x, y) form a doubly degenerate pair and transform
as E. This identification is useful for understanding the behaviour of the components
of the dipole moment, or p-orbitals, in this symmetry environment. The character
table shown also tells us how quadratic functions and cubic functions transform. The
former information is useful for d-orbitals while the latter describes the symmetry
properties of f-orbitals. It should be noted that s-orbitals have no directional
dependence and therefore they always transform under the totally symmetric IR;
for example, A1 for a C3v molecule.As will be discussed later, the quadratic functions
also describe the symmetry properties of the polarizability of a molecular system,
which is important for understanding Raman spectroscopy. Furthermore, the
character table always lists the symmetry behaviour of rotations about the x-, y-,
and z-axes (Rx, Ry, and Rz respectively), as shown in figure 3.5. A rotation vector will
be symmetric if, under a symmetry operation, its sense of rotation remains the same,
but anti-symmetric if it changes between left and right-handed.
To conclude our excursion into symmetry principles, we can now identify their
principal application in the realm of molecular spectroscopy, where they afford the
basis of selection rules for polyatomic molecules. As we established earlier, the most
common form of electromagnetic interaction between light and matter takes place
through electric dipole coupling. In chapter 2, equation (2.4), we saw that the key
3-9
Molecular Photophysics and Spectroscopy (Second Edition)
This represents just one specific instance of a more general rule, expressible as;
∫ ψ ′* μψdτ ≠ 0, (3.1)
where the dipole operator μ sits between an initial wavefunction and the complex
conjugate of the final state wavefunction, in an integral over a generalised space τ —the
latter including every degree of freedom that undergoes change in the transition. For the
electronic transitions of hydrogen this was a simple integral over the three spatial
coordinates of the electron; for electronic and other transitions in polyatomic molecules
it is less straightforward. But here is where symmetry comes powerfully into play.
To ascertain whether the transition integral shown above may be non-zero, one
constructs the direct product Γ(ψ’)⊗Γ(μ)⊗ Γ(ψ) by multiplying the characters in each
column of the character table for the IRs involved. The resultant character set may
be irreducible (if it tallies with one of the representations listed in the table) or
reducible—in which case it can be reduced to a linear combination of IRs. In either
case the transition is only allowed if the direct product equals or contains the totally
symmetric representation. More simply, and equivalently, a transition will be
allowed if the direct product Γ(ψ’)⊗ Γ(ψ) ⊆ Γ(μ) = Γ(x, y, z). It should be noted
that since the dipole moment may have components represented by more than one
IR it follows that a transition between two states in a polyatomic molecule may be
allowed by some components and forbidden for others. We shall see the imple-
mentation of these principles, and many important conclusions that can be drawn
from it, in many of the chapters that follow.
3-10
IOP Publishing
Chapter 4
Electronic and nuclear energy levels in molecules
where q and Q are schematic labels indicating all the electron coordinates, and all
the nuclear coordinates, respectively. The square brackets around Q shown in the
parameters of the electronic wavefunction indicate an indirect, parametric depend-
ence on Q; physically this reflects the fact that, to a very good approximation, the
electrons in a molecule respond instantaneously to any nuclear motion. Thus, in
each electronic state the electrons move on the potential energy surface, V(Q),
formed by the nuclear motion in that electronic state. The nuclear wavefunction
itself factorises into two parts: vibrational and rotational. As we shall see, such a
factorisation of the full molecular wavefunction also enables the corresponding
identification of separate energy terms.
The wavefunctions in quantum terminology are all eigenstates of the electronic,
vibrational, and rotational parts of the molecular Hamiltonian, derived using the
Born–Oppenheimer approximation. This in turn means they are stationary states,
i.e. stable states whose measurable properties do not change with time. All this
changes of course when light (indeed, any electromagnetic field) is present, so
allowing transitions to take place—which is the fundamental basis of spectroscopy.
In principle there are other energies relating to whole-molecule linear motion
(translation) through space, and nuclear spin. The translational levels usually merge
The three contributions on the right of this equation relate to electronic, vibrational,
and rotational energies, respectively, of successively diminishing magnitude.
In terms of energy, the lowest electronic levels are usually very far apart, and
typically the frequencies for transitions between them lie in the visible or the
ultraviolet part of the spectrum. We can readily verify that at room temperature
almost every molecule is its electronic ground state; that is, the Boltzmann population
factors for electronically excited states are close to zero. For example, we can do the
calculation for a typical carbonyl group based on its π–π* transition. The correspond-
ing absorption typically occurs at an ultraviolet wavelength λ of around 200 nm, so we
can calculate exp(−ΔE/kBT) = exp(−hc/kBTλ), where kB is Boltzmann’s constant and
T is the absolute temperature. With kB = 1.381 × 10−23 J K−1, c = 2.998 × 108 m s−1
and h = 6.626 × 10−34 J s, then at T = 298 K we find the result of around 10−105, in
other words no molecules will be in the electronic excited state at room temperature—
in any volume of the world where the concept of ‘room’ could have meaning.
Numerous vibrational levels, with much closer spacing, are associated with each
electronic level. Since the corresponding energy gaps ΔE between adjacent levels are
that much smaller, so too are the frequencies at which vibrational transitions are seen.
The correspondingly longer wavelengths involved in these transitions generally lie in
the infrared part of the electromagnetic spectrum. We can calculate for example that
for the wavenumber ν = 1700 cm−1 we have exp(−ΔE/kBT) = exp(−hc ν /kBT), giving
a result of 2.73 × 10−4. In other words, about one molecule in three or four thousand
will be in its first vibrationally excited state at room temperature. This is still usually a
small enough number to regard as insignificant. So basically, it is safe to assume that,
at room temperature, a molecule undergoing an absorption transition starts from the
ground vibrational level in the ground electronic state. Equally, emission spectra
(associated with excited state decay) will not be observable without prior excitation.
Finally, we have rotational levels, and these are still more tightly packed together:
rotational transitions are therefore observed at yet lower frequencies and longer
4-2
Molecular Photophysics and Spectroscopy (Second Edition)
wavelengths, in the far infrared or microwave region. For a diatomic molecule, for
example, figure 4.1 illustrates the varying spacings of the different types of
energy level.
The fact that the rotational levels are separated by the smallest amounts of energy
has two immediate consequences. First it is clear that at normal room temperatures,
thermal energies (typically corresponding to wavenumbers of magnitude kBT/hc ~
200 cm−1) will greatly exceed the rotational level spacing, which are typically
~1 cm−1 or less, unless these small molecules contain the exceptionally light atom
hydrogen. Consequently, there will usually be a substantial population of numerous
rotational states, not just the lowest one. But secondly, unlike internal electronic and
vibrational motions, rotations can be frustrated in the condensed phase by the steric
effect of neighbouring molecules, so giving rise to a heterogeneous broadening of the
energy levels. Hence, complete energy level quantisation based on free rotation can
only occur in a rarefied medium, and the observation and investigation of rotational
transitions is essentially a gas-phase technique.
For any given molecule, each electronic state of the molecule is usually identified
by some label, often based on the molecular point group and state symmetries, as
described in chapters 2 and 3. For every such state there is an equilibrium nuclear
geometry about which vibrations occur, signifying oscillatory motions about a
configuration of minimum potential energy. The form of each vibration is controlled
by restoring forces associated with the shape of each potential energy surface—and
the detailed form of the latter depends on the electronic state. Thus, each electronic
state has its own set of vibrational states. So too, each vibrational level has its own
set of rotational states.
Figure 4.1. Typical plots of potential energy V against the interatomic distance (bond length) r for a diatomic
molecule, schematically showing (not to scale) the positioning of the quantised electronic, vibrational and
rotational energy levels.
4-3
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 4.2. Independent translational and rotational motions of the linear carbon dioxide molecule.
4-4
Molecular Photophysics and Spectroscopy (Second Edition)
Before considering the form of those vibrations, let us consider another triatomic
molecule, H2O, that is not linear. As shown in figure 4.3, here there are three
translational and also three rotational degrees of freedom for this (and any) non-
linear molecule.
Thus, for any non-linear molecule—we have six of our 3N coordinates accounted
for, leaving (3N − 6) degrees of freedom which correspond to the vibrational modes
of the molecule. In summary, we have:
• (3N − 5) vibrational degrees of freedom for a linear molecule:
• (3N − 6) vibrational degrees of freedom for a non-linear molecule.
Now it turns out that, just like the translations and rotations, there are certain
specific vibrations in terms of which all others can be expressed. These are known as
the normal modes of vibration. Just as a molecule moving in the x-direction cannot
begin to move in the y- or z-direction without some external force being applied, so
too a molecule oscillating in one of its normal modes cannot begin oscillating in a
different mode without being excited in some way. Moreover, each of these modes of
vibration is close to being simple harmonic in form—at least for the lowest energy
quantum levels (see chapter 5).
So, to summarise: the nuclear motion part of the molecular wavefunction
factorises into a product of rotational and wavefunctions for each of the 3N − 6
normal modes of vibration (3N − 5 if the molecule is linear), so that we effectively
now have:
3N − 6(5)
Φ(q , Q ) ≈ ψmelec(q , [Q ])ψ rot(Q ) ∏ ψvvib(
i
m)
(Qi ), (4.3)
i
where the capital pi, Π, signifies a product of the nuclear vibrational wavefunctions
for each normal mode, indicated by the index i.
Figure 4.3. Independent translational and rotational motions of the non-linear water molecule.
4-5
Molecular Photophysics and Spectroscopy (Second Edition)
where x signifies the extent of displacement from the equilibrium configuration (at
the bottom of the quadratic potential well). Also, ω = 2πcν , where ν is the
wavenumber of the vibration, and ℏ = h/2π : hence, we find;
∴ x 2 = ℏ/4πmcν = h /8π 2mcν .
4-6
Molecular Photophysics and Spectroscopy (Second Edition)
Exercise
Using the values for h and c and the data m(O) = 16.00 amu (1 amu = 1.6606 × 10−27 kg)
and ν = 1700 cm−1, evaluate <x2> for the carbonyl bond. (This establishes the
variance, since <x> = 0.)
Answer
The result, 6.197 × 10−24 m2, has a square root which signifies a mean (rms)
variation in bond length of about 2.49 pm. Compare this with the reported
equilibrium bond length which is around 123 pm. The fractional change, almost
exactly a fiftieth, is a fairly typical result.
So it is that in molecular vibrations, the atoms don’t generally move very much,
and therefore the Born–Oppenheimer approximation, separating electronic from
nuclear motions, works well.
We still have to remember that in most regular samples, we are not seeing the
behaviour of any individual molecule in isolation. Whether the molecules are held
together immobile in the static structure of a solid, or constantly buffeted by other
molecules in the mobile surroundings of a liquid, the proximity of those other
surrounding molecules slightly modifies the electronic environment of each mole-
cule, generally contributing to a level broadening. Taking account of these factors,
we can understand why it is that for most molecules the absorption of light, although
a quantum phenomenon, gives broad lines rather than the sharp, discrete lines found
in atomic spectra. As we shall see, this is a very important distinction; as a corollary,
there is a very high density of states (ρ) associated with any energy level within the
vibrational continuum of a given electronic state. (Note, the term density of states
signifies the number of quantum states to be found within a small energy interval—
this terminology has nothing to do with volume.) If we compare the successive
vibrational levels in even a simple series of increasingly large molecules, we can see—
even without taking the corresponding rotational levels into account—how the density
of states rapidly increases with the molecular size.
Exercise
Consider a molecule of formic acid HCOOH (methanoic acid), for which the
fundamental vibration wavenumbers are as given below (note the 3N − 6 rule). To
gauge the density of states, estimate how many vibrational states will generally be
found in a typical 1000 cm−1 energy interval. Note: the first 1000 cm−1 above the
ground vibrational state will not be typical, because only the single-quantum
vibrational levels for ν6 and ν 9 can feature below that threshold (table 4.1).
Answer
For formic acid, the average (1/3.570) + (1/2.943) + … = 7.76; say ~8, the nearest
integer (since one cannot have a fraction of an energy level!). However, any such
estimation ignores the huge number of higher energy levels, rapidly generated by the
so-called overtones such as 2ν6 = 2210 cm−1 or combination bands and overtones
such as ν4 + ν7 = 2012 cm−1; ν5 + 2ν7 + ν9 ≡ ν4 + ν6 + ν7 = 3117 cm−1. Even this is
4-7
Molecular Photophysics and Spectroscopy (Second Edition)
Table 4.1. Vibrations of HCOOH. From: Shimanouchi T 1972 Tables of Molecular Vibrational
Frequencies Consolidated, Nat. Stand. Ref. Data Ser., N.B.S. (U.S.) 1, 39.
ν1 OH stretch 3570
ν2 CH 2943
ν3 CO stretch 1770
ν4 CH bend 1387
ν5 OH bend 1229
ν6 CO stretch 1105
ν7 OCO deform 625
ν8 CH bend 1033
ν9 Torsion 638
without regard to Fermi resonance—an energy level splitting that occurs when two
vibration modes of the same symmetry are close in energy. Overtones and
combination bands will be discussed further in chapters 6 and 7.
Computer calculations show that the number of nuclear vibration levels rises
exponentially with N, the number of atoms in the molecule. It is clear that many
thousands of vibrational lines per 1000 cm−1 will emerge from a similar calculation
even for a relatively small molecule. For example, benzophenone, shown in
figure 4.4, has 24 atoms, properly counting all the carbon and hydrogen atoms as
well as the oxygen, and hence 66 separate individual vibrational modes—all of
which can exhibit combination bands and overtones.
4-8
IOP Publishing
Chapter 5
Small molecule rotational energy levels
and spectra
where the sum is taken over all component particles (in the case of molecules, the
individual atoms) of masses mi at equilibrium distances ri from the centre of mass—
the latter being the point about which rotation occurs. In the case of the diatomic
molecules, the centre of mass has a location defined such that:
m1r1 = m2r2 . (5.2)
Since r1 + r2 = r, we find:
m2 m1
r1 = r; r2 = r. (5.3)
m1 + m2 m1 + m2
Figure 5.1. Key parameters for describing the rotations of a diatomic molecule.
I = m1r12 + m2r22
⎛ m2 ⎞2 ⎛ m1 ⎞2
= m1⎜ ⎟ r 2 + m2⎜ ⎟ r2
⎝ m1 + m2 ⎠ ⎝ m1 + m2 ⎠ (5.4)
m m (m + m1)r 2 ⎛ m1m2 ⎞ 2
= 1 2 2 =⎜ ⎟r = mr r 2 .
(m1 + m2 )2 ⎝ m1 + m2 ⎠
Thus, the moment of inertia in this simple case reduces to the square of the bond
length multiplied by what is known as the reduced mass, mr, (note: some books use
the symbol μ), the latter defined as;
m1m2
mr = . (5.5)
m1 + m2
Now let us think about the rotation itself. The two atoms must have to move at
different speeds, ν1 and ν2, obviously proportional to their distances from the centre
of mass—or else one of them would catch up with the other one (figure 5.2). The
constant of proportionality, ω, is called the angular velocity. So, we can write
ν1/r1 = ν2/r2 = ω. It therefore follows that the kinetic energy of rotation is given by:
1 1 1 1 1 1 2
E= m1ν12 + m2ν 22 = m1r12ω 2 + m2r22ω 2 = (m1r12 + m2r22 )ω 2 = Iω . (5.6)
2 2 2 2 2 2
Note the similarity to the usual kinetic energy formula. Now the angular
momentum of the molecule is given by:
Again, the result is like the linear motion momentum, p = mv. So classically, we can
write the rotational energy in terms of the angular momentum, using the earlier
equations (5.4) and (5.5), as follows:
L2
E= . (5.8)
2I
5-2
Molecular Photophysics and Spectroscopy (Second Edition)
L = ℏ J (J + 1) , J = 0,1,2,... (5.9)
where ℏ = h/2π and J is the rotational quantum number which as noted, can only
take on discrete integer values 0, 1, 2, …. The angular momentum is a signified by a
vector, directed along the axis perpendicular to the plane of rotation with the sense
of a right-handed screw, and with a magnitude L, as indicated in figure 5.3.
It follows that the energies are therefore expressible as:
E = BhJ (J + 1), (5.10)
h
B= , (5.11)
8π 2I
where B is the rotational constant of the molecule, with units of Hz. (Note that
several authors define B differently, introducing a dividing factor of c, the speed of
light, such that its units are cm−1. Others introduce another power of h such that the
units are Joules.) The first few rotational energy levels are shown in figure 5.4.
Because the spacing between rotational levels depends on B, which is inversely
proportional to the moment of inertia, the levels became more and more tightly
packed as the moment of inertia increases—this is true for non-linear molecules, too.
As derived below, the spacings between rotational levels (indicated by the arrows in
figure 5.4) are not constant but increase linearly with increasing J.
One other property of the rotational levels should be mentioned before we go any
further, namely the degeneracy. It is another feature of the quantum mechanical
5-3
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 5.3. Angular momentum vector for the rotating diatomic molecule.
Figure 5.4. First few energy levels of a perfect rotor in units of Bh (vertical arrows also showing allowed
transitions), and degeneracy factors gJ associated with each of the rotational quantum numbers J.
5-4
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 5.5. Permissible orientations of the angular momentum vector for the case J = 2.
If we apply an electric field, the angle that each angular momentum vector makes
with the field vector produces an offset in the energy—for example with J =2 there
are five states, all with a slightly different energy. But with no field, there is no
preferred direction and for a given J, all values of M give the same energy. In
general, there are (2J + 1) values of M, so we have (2J + 1) energetically equivalent
states: we therefore say that the rotational state with quantum number J has
(2J + 1)-fold degeneracy, or
gJ = (2J + 1). (5.13)
The degeneracy factor is important because it means there are more and more states
with the same energy as J increases—and this affects the thermal distribution of
molecules between the various rotational states.
Having found a formula for the rotational energy levels, we now have to
investigate what transitions can be observed in a spectroscopy experiment.
We have already seen in a general way that when a molecule is irradiated with
light of the right frequency, light absorption occurs if and only if there is a non-zero
transition dipole moment. Applying the general formula of equation (7), it follows
that for a transition from a state with rotational quantum number J to a state with
quantum number J´, this transition dipole is given by
5-5
Molecular Photophysics and Spectroscopy (Second Edition)
because molecular rotation doesn’t influence or induce any dipole in the molecule.
The physical significance of the apparently innocuous relation (5.15) is that we can
only have a transition dipole moment if there is a permanent dipole in the molecule—
otherwise no rotational transitions can occur. We can summarise:
So, whereas we can certainly observe the rotational spectra of heteronuclear dia-
tomics CO, HCl etc, we can evidently see nothing from homonuclear molecules such
as O2, N2 which, being centrosymmetric, have no permanent electric dipole. In fact,
this is a general rule, true also for non-linear molecules. For example, benzene does
not give a pure rotation spectrum because it too has no permanent dipole moment.
Still, pure rotational spectra of centrosymmetric molecules can be obtained using a
light scattering phenomenon called the Raman effect, a subject that is covered in
chapter 8.
The other important thing that we find out when evaluating the transition dipole
MJ ′ J is that:
MJ ′J ≠ 0 if J ′ = J ± 1: MJ ′J = 0 if J ′ ≠ J ± 1.
Therefore, transitions do not occur (they are forbidden) unless the rotational
quantum number changes by one unit only. So, we have the selection rule for
rotational spectroscopy:
ΔJ = ±1. (5.16)
ΔJ = + 1 corresponds to an increase in energy by absorption; ΔJ = −1 corresponds
to emission of light. This is consistent with the conservation of angular momentum
where the total angular momentum of the system (light + molecule) is conserved.
So, if we irradiate molecules and observe absorption frequencies for transitions
J → J + 1, these frequencies are given by:
hν = ΔE = EJ +1 − EJ
= Bh(J + 1)(J + 2) − BhJ (J + 1)
(5.17)
= 2BhJ (J + 1)
∴ ν = 2B(J + 1)
and the spectrum accordingly shows absorption lines at 2B, 4B etc (see figure 5.4).
As noted in chapter 4, in each case the absorption intensity will depend significantly on
the population of the state from which the transition occurs. Since rotational levels are
closely spaced, it emerges that many of the states with low J are appreciably populated
at room temperature—the population is governed by the Boltzmann factor (accom-
modating the degeneracy factor pre-multiplier) NJ ~ (2J + 1) e−BhJ(J + 1)/kBT.
It is relatively straight forward to numerically estimate the most populated J-level,
5-6
Molecular Photophysics and Spectroscopy (Second Edition)
kBT
Jmax = − ½. (5.18)
2hB
Equation (5.14) shows the expected behaviour that as the moment of inertia
increases, B decreases, and more J levels become populated for a given T.
An idealised spectrum is shaped as shown in figure 5.6, with a frequency interval
νJ+1 − νJ = 2B between successive absorption lines. As T decreases the rotational
line intensity distribution favours lower J levels while at higher T, the distribution
broadens and shifts its maximum to states with higher J values.
Exercise
Using equation (5.14), find the most populated rotational levels of nitrogen gas at
room temperature, 298 K and at 1200 K. For the nitrogen molecule, the quantity
hB/kB (deceptively known as the characteristic temperature for rotation) is 2.87 K.
Answer
The answer has to be an integer, and the nearest values for Jmax are 7, at room
temperature; 14 at 1200 K. The answer reminds us that the air around us is
rotationally quite excited; whereas for nitrogen the most populated level is Jmax =7,
for oxygen at room temperature (characteristic temperature 1.45 K) Jmax = 10. The
big difference is due to the oxygen molecule having slightly more massive atoms, and
a slightly greater bond length, than diatomic nitrogen; hence its moment of inertia is
larger, and the rotational levels are closer together.
Figure 5.6. Idealised absorption spectrum of a rigid rotor heteronuclear diatomic molecule.
5-7
Molecular Photophysics and Spectroscopy (Second Edition)
5-8
Molecular Photophysics and Spectroscopy (Second Edition)
This is, however, on the assumption that the bond lengths are unchanged on isotopic
substitution—in fact, this is not quite true, and consequently the results are not quite
as accurate as those for diatomic molecules.
Exercise
To determine both the H–C and C≡N bond lengths in hydrogen cyanide, rotation
spectra are collected from two isotopomers, H12C14N and D12C14N. Here deuterium
D is an isotope of H (2H) found for example in heavy water, having mass ~2 u. The
corresponding rotational spectra show adjacent line spacings of 88.632 and 72.416
GHz, respectively. Calculate each of the moments of inertia, IHCN and IDCN; then
using the following formula, work out the C≡N bond length, rCN;
⎛ 9 ⎞ ⎛1 ⎞
rCN = ⎜ ⎟IHCN − ⎜ ⎟IDCN .
⎝ 28u ⎠ ⎝ 6u ⎠
5-9
Molecular Photophysics and Spectroscopy (Second Edition)
Answer
Since the separation between adjacent lines equals 2B, the B values are 44.316 × 109
Hz for HCN and 36.208 × 109 Hz for DCN. Also,
B = h 8π 2I ∴ I =h 8π 2B .
Hence, with h = 6.626 × 10−34 J s; u = 1.661 × 10−27 kg, we find IHCN = 1.8937 ×
10−46 kg m2 and IDCN = 2.3178 × 10−46 kg m2. Using the equation leads rCN = 1.157
× 10−10 m, or 0.1157 nm.
Table 5.1. Frequencies of selected rotational transitions in 16O=12C=32S; observed and calculated results
derived on the basis of fixed, or variable bond length (third and fourth columns, respectively).
5-10
Molecular Photophysics and Spectroscopy (Second Edition)
One then empirically adds additional centrifugal terms of order J n(J+1)n, n ⩾3, as
needed.
Figure 5.7. Examples of the four major classes of rotor in three dimensions.
5-11
IOP Publishing
Chapter 6
Diatomics and triatomics: vibrational energy
levels and spectra
We now turn to the subject of vibrational energy levels and spectra, again starting
with the simplest case, the diatomic molecule. For this analysis, the key parameters
are defined as shown in figure 6.1.
Figure 6.1. Key parameters for a diatomic molecule. The left-hand atom is arbitrarily considered the
coordinate origin for displacement of the right-hand atom, so that its appearance as stationary is for
illustrative purposes only. The molecule experiences a restoring force F on extension (as shown) or
compression of the bond, where r < re.
Table 6.1. Molecular parameters deduced for the hydrogen halides from
rotational and vibrational spectra.
6-2
Molecular Photophysics and Spectroscopy (Second Edition)
Whereas classically there is no restriction upon the value of the total energy
(potential plus kinetic energy) of the system, the quantum mechanical treatment of
the simple harmonic oscillator limits the energy to discrete amounts:
⎛ 1⎞
E = ⎜v + ⎟hν0 (6.3)
⎝ 2⎠
where v is the vibrational quantum number (often given the alternative symbol n.
Equally, the energy expression is commonly cast as E = n + 12 ℏω0 , where ( )
ω0 = 2πν0). The possible values of v are integers 0, 1, 2, …. So, we have equally
spaced levels with ΔE(v) ≡ E(v + 1) − E(v) = hνo. This system exhibits a zero-point
energy of E = ½ hν0 when v = 0. The origin of the latter can be considered a
quantum uncertainty effect, since a zero for both kinetic and potential energy would
contravene the position-momentum uncertainty principle—see the calculation in
chapter 4. The zero-point energy is not dependent on temperature, which means that
hypothetically even at a temperature of absolute zero, the oscillator would still be in
motion—a result that has no classical analogue. The first few levels and wave-
functions for the harmonic oscillator are shown in figure 6.2.
An examination of the quantum mechanical wavefunctions exhibited in figure 6.2
also shows that there is a finite probability that one can measure the bond length of
the quantum oscillator to be either larger or smaller than the values expected at the
classical extremes. This is strictly another quantum mechanical effect called
tunnelling, where one can find a system in regions of energy E < V which are
forbidden classically. One other especially striking quantum feature, exhibited by the
Figure 6.2. Potential energy curve for the simple harmonic oscillator model of a diatomic molecule, with the
potential energy V plotted against bond length r and centred on an equilibrium value re, also showing the
positioning of the first few quantum energy levels and their normalised wavefunctions.
6-3
Molecular Photophysics and Spectroscopy (Second Edition)
ground state alone, is that the probability of finding a given separation between the
two atoms in the molecule is greatest near to the equilibrium bond length. (Recall
that large amplitude features in the normalised wavefunction ψ, of either positive or
negative sign, are associated with the largest values of probability per unit length,
∣ψ∣2.) This is in complete contrast to the classical behaviour.
Although it only becomes fully apparent for higher energy levels, it emerges that
as the vibrational quantum number v continues to increase, the probability of
finding the molecule does more closely approach classical limits. Modelling the
oscillator as a single particle with a reduced mass mr moving back and forth within
the potential well, the probability of finding the particle within a particular range of
positions would classically be equivalent to the fraction of time, within one cycle of
oscillation, spent passing through that region. The particle would accordingly be
most likely found closest to a turning point, where the total energy is entirely
potential energy—as compared to its equilibrium position, where the total energy is
all kinetic energy. Classically, these points correspond to the bond lengths of
maximum extension and compression—the two points where the parabolic potential
curve crosses the horizontal line for a particular energy. This is a manifestation of
the correspondence principle for large quantum numbers. (Note that the notion of
‘turning point’ in this classical sense, which we shall see in later chapters proves to be
a very useful concept in advanced spectroscopy, is not the same as the mathematical
sense of that term as a position where a derivative changes sign.)
For basic spectroscopy, our first question must once again be, what transitions are
allowed between vibrational levels? For the transition v → v′, our transition dipole
moment is given by
Evaluation of the integral reveals that, for the transition moment to be non-zero, the
transition needs to satisfy the selection rule v′ = v ± 1, i.e. any change in vibrational
quantum number is governed by:
Δv = ±1. (6.5)
At this stage we therefore conclude that for each energy level, absorption transitions
(upwards in energy, Δv = +1) can only occur to the level immediately above;
emission transitions (downwards, Δv = –1) go to the level one below. From the
detailed evaluation of equation (6.4) we can also discover the following
proportionality:
⎛ dμ ⎞
Mv ′ v ∝ ⎜ ⎟ . (6.6)
⎝ dr ⎠ r
e
So, the dipole moment of the molecule must change during the vibration for the
transition to be allowed:
6-4
Molecular Photophysics and Spectroscopy (Second Edition)
In the case of diatomic molecules, it is clear that vibrational spectra can therefore
only be recorded for heteronuclear compounds such as CO, HCl, but not for
homonuclear species such as H2, N2. When we come to polyatomic molecules we
shall discover cases where a molecule lacks a permanent dipole moment, but it
acquires dipolar character in the course of a particular vibration—with the result
that the vibration is active in the absorption spectrum. Unlike purely rotational
spectra, it is not enough just to have a dipole moment: hence the necessarily lengthier
statement of the selection rule.
Now from equations (5.18) and (6.1) it is clear that the absorption frequency for
the v = 0 → 1 transition is given by hν = ΔE = 32 hν0 − 12 hν0 (i.e. photon
energy = difference between vibrational level energies). Hence,
ν = ν0. (6.7)
This is not as trivial as it appears: it means that the absorbed frequency of light
equals the classical frequency of molecular vibration, (1/2π ) k /mr from equation
(6.2). Vibrational frequencies are typically in the range 1012–1014 Hz, or hundreds to
thousands of wavenumbers (cm−1); hence absorption bands are measured in the
infrared. When we interpret infrared spectra, we often take it for granted that
absorption frequencies directly signify vibrational frequencies—but even this is true
only for simple harmonic vibrations. By way of contrast, as we have seen, no such
principle applies to rotations, for example; the absorption frequencies measured in
rotational spectroscopy cannot be directly interpreted as frequencies of rotation.
6-5
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 6.3. Potential energy curves for an oscillator: simple harmonic (…..) and anharmonic (___).
6-6
Molecular Photophysics and Spectroscopy (Second Edition)
which has the property that as r approaches ∞, V(r) approaches the value De. The
constant β is specific for the diatomic being considered. If we solve the Schrödinger
equation using a Morse potential we find that the energy levels are given by the
following equation:
⎡⎛ 1⎞ ⎛ 1 ⎞2 ⎤
E = ⎢⎜v + ⎟ − xe⎜v + ⎟ ⎥hν0 (6.9)
⎣⎝ 2⎠ ⎝ 2⎠ ⎦
Figure 6.4. Energy levels shown on the potential curve for an anharmonic oscillator, showing the dissociation
energy D0 as it is experimentally measured, i.e. relative to the ground state, and also as defined with respect to
the bottom of the classical potential energy curve at the equilibrium position (De). Note how the average bond
length, signified by a position midway across the horizontal lines for successive energy levels, increases with
increasing v .
6-7
Molecular Photophysics and Spectroscopy (Second Edition)
Table 6.2. Experimental results for the positioning of the first few vibrational levels of H35Cl, compared with
the results of calculations based on the harmonic oscillator and anharmonic oscillator models. Here the
superscript on the Cl atom symbol indicates which isotope of Cl is being considered.
Figure 6.5. Schematic representation of the differences between the vibrational infrared spectrum of a
supposedly simple harmonic diatomic molecule, and one with anharmonic vibrations.
6-8
Molecular Photophysics and Spectroscopy (Second Edition)
smaller again by another factor of ten. The other new features are transitions that
originate from vibrationally excited states, for which the absorption intensities are
governed by the Boltzmann factor exp(−Einitial/kBT). Since at room temperature
kBT ∼ 200 cm−1, the excited vibrational levels are generally poorly populated, so
that transitions from the vibrational ground rate are by far the strongest. Here we
see a marked contrast to the population of rotational levels, seen in the previous
chapter. However, since the excited state vibrational populations depend on
temperature, the intensity of such bands grows as temperature is increased—hence
the term hot bands is commonly applied. In the harmonic oscillator model, hot
bands would be obscured because their transition frequency is the same as the
fundamental.
From equation (6.9) we find the following expressions for the key transitions:
⎧⎡ 3 ⎛ 3 ⎞2 ⎤ ⎡ 1 ⎛ 1 ⎞2 ⎤⎫
⎪ ⎪
⎧⎡ 5 ⎛ 5 ⎞2 ⎤ ⎡ 3 ⎛ 3 ⎞2 ⎤⎫
⎪ ⎪
⎧⎡ 5 ⎛ 5 ⎞2 ⎤ ⎡ 1 ⎛ 1 ⎞2 ⎤⎫
⎪ ⎪
⎧⎡ 7 ⎛ 7 ⎞2 ⎤ ⎡ 1 ⎛ 1 ⎞2 ⎤⎫
⎪ ⎪
Note that the fundamental frequency is not precisely ν0 , but slightly less; however if
we find the overtones or a hot band we have enough information to deduce xe, and
hence find the true value of ν0 .
Exercise
The ratio of the principal overtone and the fundamental absorption frequencies
satisfies the relation (as follows from the above results):
Using this result, calculate a value for the anharmonicity constant of HCl, given that
the absorption spectrum shows a fundamental absorption at 2885.9 cm−1 with
another weak line, ascribed to the principal overtone, occurring at 5668.0 cm−1.
Answer
0.017 36.
At dissociation, the spacing between adjacent vibrational levels will close in to
zero. The highest v-level supported by the potential, nD, can then be determined by
6-9
Molecular Photophysics and Spectroscopy (Second Edition)
assuming the relationship for ΔE(v) given in equation (6.5) holds. A graph of ΔE(v)
versus v to determine nD is called a Birge–Sponer plot. It emerges that De can be
calculated using:
hνo2
De = . (6.10)
4νoxe
(Note that for both the harmonic and anharmonic oscillators, the depth of the
potential well De is not the same as the measured dissociation energy D0, because of
1
the zero-point vibrational energy; to a good approximation De = D0 + 2 hν0.)
Exercise
For the molecular vibration of H35Cl the anharmonicity constant, determined by
comparing the absorption frequencies for the first few transitions, is evaluated at
0.018. Estimate the number of vibrational levels before a dissociation limit is
reached.
Answer
The dissociation limit should be the point at which the diminishing gap between successive
levels becomes zero. Thus, for this level nD we should have E(nD) = E(nD + 1).
From equation (6.4) we therefore have
⎛ 1⎞ ⎛ 1 ⎞2 ⎛ 3⎞ ⎛ 3 ⎞2
⎜n D + ⎟ − xe⎜n D + ⎟ = ⎜n D + ⎟ − xe⎜n D + ⎟
⎝ 2⎠ ⎝ 2⎠ ⎝ 2⎠ ⎝ 2⎠
1 1 3 9
∴ n D + − xen D2 − xen D − xe = n D + − xen D2 − 3xen D − xe
2 4 2 4
∴ 2xen D = 1 − 2xe
∴ n D = (1 − 2xe )/2xe = 0.964/0.036
= 27 (result for the next largest integer).
6-10
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 6.6. Energy level representations of the rotation–vibration transitions in a heteronuclear diatomic
molecule, shown in order of increasing optical frequency and mapped to the corresponding lines in the
absorption spectrum. The rotational quantum numbers in the ground and first excited vibrational levels are
here designated J and J′ respectively. Note the distribution of intensities amongst the lines in the spectrum,
primarily determined by the thermal Boltzmann population of the initial rotational levels.
6-11
Molecular Photophysics and Spectroscopy (Second Edition)
Here, we use symbols J and J′ respectively for the rotational quantum numbers in
the ground and first excited vibrational levels. (It may be noted many traditional
texts use J″ for the former.) For the P- and R-branch transitions specifically, we
therefore have:
P−branch; J′ = J = 1 ∴ v = v0(1 − 2xe ) + B′(J − 1)J − BJ (J + 1)
∴ v = v0(1 − 2xe ) + (B ′ − B )J 2 − (B ′ + B )J
R−branch; J′ = J = 1 ∴ v = v0(1 − 2xe ) + B′(J + 1)(J + 2) − BJ (J + 1)
∴ v = v0(1 − 2xe ) + (B′ − B )J 2 + (3B′ − B )J + 2B′ .
By analysing the vibration–rotation spectrum we can evidently determine the
rotational constants (and hence moments of inertia and bond lengths) for both
the ground and first excited vibrational states, without resorting to microwave
spectroscopy, albeit usually with less precision.
It is worth remarking that such spectra can appear more complicated if the
elements involved have more than one significant isotope in their natural form of
abundance. The vibration–rotation spectrum of ordinary HCl, for example, in
contrast to the pure isotope spectrum shown on the previous page, reveals each line
to be split into a doublet, because there are two isotopomers, H35Cl and H37Cl; since
these have slightly different reduced masses, their vibrational and rotational energy
levels differ slightly. The intensities of the two resultant lines of each doublet reflect
the ∼3:1 natural abundance ratio of the chlorine isotopes.
6-12
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 6.7. Normal modes of vibration of the carbon dioxide molecule, and their wavenumbers.
Figure 6.8. Normal modes of vibration of the water molecule, and their wavenumbers.
Consequently, the selection rules are exactly as for the diatomic case:
Δvi = ±1 (6.14)
and
⎛ dμ ⎞
Mv ′ v ∝ ⎜ ⎟ (6.15)
⎝ dQi ⎠ Q
e
6-13
Molecular Photophysics and Spectroscopy (Second Edition)
Water has a permanent dipole moment, and it changes in each of its three normal
vibrations—so all of them are active, giving three main bands in the infrared
spectrum. On the other hand, CO2 has no permanent moment—the dipole remains
zero throughout the symmetric stretch vibration ν1—yet the ν2 and ν3 vibrations
remove the centre of symmetry and allow a dipole moment to be formed. Using a
bond dipole model, the variation of molecular dipole moment during the course of
each of the vibrations of CO2 is shown in figure 6.9.
Consider ν2 for example. The dipole oscillates about zero, and the derivative of
the equilibrium position is non-zero—see figure 6.10. Hence, this mode is active in
the infrared spectrum. Similar arguments apply to ν3. The only mode that is not
active is ν1, the symmetric stretch, because for symmetry reasons the dipole remains
zero throughout the course of the vibration. So we notice that a difference in shape,
Figure 6.9. Normal modes of vibration of the carbon dioxide molecule, showing the variation in geometry in
the course of each vibration, and partial charge separations resulting from the greater electronegativity of
oxygen compared to carbon. At each step the resultant dipole μ from addition of the bond moments are also
shown. Note the modern physical chemistry convention that dipoles point in the direction that a free electron
would travel, i.e. pointing from more negative to more positive regions.
6-14
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 6.10. Variation in the electric dipole moment of the carbon dioxide with the coordinate Q2 signifying
the extent of its antisymmetric stretch vibration.
between these two triatomics, water and carbon dioxide, itself produces a clear
difference in the look of their infrared spectra, producing one less band for CO2.
A pictorial interpretation, while immensely insightful, becomes more challenging
although not impossible, as the number of atoms in the molecule increases. The
alternate, direct and reliable approach is to use character tables. The irreducible
representation of every ground vibrational wavefunction (v = 0) Γ(ψv ) is always
totally symmetric. Thus, the problem reduces to determining whether the irreducible
representation a vibrational mode, v′ = 1, Γ(ψv′) transforms as one of the irreducible
representations of the dipole moment operator Γ(μ). In the point group D∞h, the
representation Γ(μ) transforms as Σ u+ (z) ⊕ Πu(x,y). Thus, as shown before, ν2 and ν3
are infrared active but ν1 is not. Similarly, in C2v , Γ(μ ) transforms as A1 (z) ⊕ B1(x)
⊕ B2(z). Thus, all three vibrational modes of H2O have the correct symmetry to be
infrared active.
6-15
IOP Publishing
Chapter 7
Large molecule infrared absorption spectroscopy
When we look at larger molecules, the vibrational spectrum quickly becomes much
more complex than the cases we have looked at so far. Consider benzene, for
example, which is still a relatively small molecule compared to most of the materials
in the living world. It has twelve atoms (recall the common schematic representation
in figure 3.1, which omits the positions of one carbon and one hydrogen atom at
each vertex), and hence (3 × 12) − 6 = 30 normal modes of vibration. In fact, ten of
these are non-degenerate and ten are doubly degenerate, so there are 20 distinct
vibrational frequencies. Although not all of these modes are active in the infrared
spectra, complications arise because of overtones and combination bands.
Fortunately, the appearance of group frequencies in the vibrations of complex
molecules is a redeeming feature which does allow some sense to be made of
the spectra.
Figure 7.1. Infrared vibrational spectra of: (a) cyclohexanone, highlighting the C=O stretch at ∼1715 cm−1;
(b) ethanoic acid, with C=O stretch at ∼1739 cm−1. The two molecules are depicted in the usual schematic
chemical form.
Figure 7.2. Typical infrared absorption regions for chemical groups common in organic molecules.
7-2
Molecular Photophysics and Spectroscopy (Second Edition)
7-3
Molecular Photophysics and Spectroscopy (Second Edition)
Water has unique properties, worth noting at this juncture. Most of the physical
properties of H2O, in the condensed phase (liquid water, solutions, water-containing
substances and ice) are all strongly affected by hydrogen-bonding. It is the strength
of this interaction that keeps water liquid at room temperature. Spectroscopically,
this effect very substantially broadens the peaks in the infrared spectrum of any
substance with water content. So the ν1 and ν3 vibrations shown in figure 6.8 are
together smeared over a broad range of wavenumbers, and their absorption bands
are so strong that they would usually obliterate the signals of other substances
absorbing in the same region. For this reason, infrared spectroscopy is not a very
helpful technique for intrinsically water-containing materials—which includes
almost all biological samples; this is indeed where Raman spectroscopy comes to
the fore, as we shall see in the next chapter.
Infrared spectroscopy is one of the most familiar workhorses in industrial
laboratories; it is relatively fast, highly amenable to computer control, and the
instrumentation is relatively cheap. Pure substances whose infrared molecular
spectra are required may be measured in films, oil-based suspensions, or vaporised
for the purpose. In the analysis of more complex, real-world samples, there may
often be a need for first separating components, by chromatographic techniques for
example. The exact methodology depends very much on the nature of the sample
and the analytical precision required—a topic which goes beyond the remit of this
book. One of the more recent developments, which has come to fruition as a result of
breakthroughs in source and detector technology, lies in security and defence
applications of terahertz imaging. This deploys sensors operating in the range
0.1–20 THz (3–600 cm−1) spanning the long-wavelength infrared and microwave
regions. Here, there is seldom a need to identify specific molecules, but more often
the contrast in spectral response from different materials.
7-4
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 7.3. Absorption (logarithmic scale) spectra of haemoglobin (Hb) and oxyhaemoglobin (HbO2) across a
window of wavelengths in the near-infrared.
7-5
IOP Publishing
Chapter 8
Raman scattering and spectral interpretation
Figure 8.1. Light scattering through a non-absorbing medium. Monochromatic input radiation with optical
frequency ν, encountering atoms or molecules, produces scattered light with frequencies ν′.
The quantum mechanical expression for the polarizability tensor is beyond our
present scope, but it will be useful to observe that it has the properties of a product of
transition electric dipoles (in fact, a weighted sum of such products). On this issue it
is worth noting that a superficially appealing, traditional ‘polarizability ellipsoid’
description proves rather less than helpful; indeed it can be highly misleading if used
to describe inelastic scattering. Such a representation does not allow for more than
three independent polarizability components.
To continue: from equation (8.2) it follows that if the electric field is oscillating at
frequency ν, so too does the dipole oscillate at this frequency, giving rise to its own
electric field at frequency ν. In the language of classical electrodynamics, the
molecule acts like a secondary source; electromagnetic waves are emitted in all
directions with this same frequency, and Rayleigh scattering is therefore observed.
Rayleigh scattering using a monochromatic laser has a strong angular dependence.
The strongest scattering is in the forward direction, coincident with the direction of
propagation of the incident beam, while the weakest signals are in the opposite
direction. This angular effect however is not prominent when using an incoherent
lamp or in sunlight.
8-2
Molecular Photophysics and Spectroscopy (Second Edition)
shifted from the frequency of irradiation. These lines appeared shifted to both the high
and low frequency sides of ν by equal amounts. However, the intensity of lines on the
high-frequency side was observed to fall off rather sharply.
The frequency shifts measured in the scattered light were shown to be character-
istic of the sample, but not due to fluorescence—spontaneous light emission from an
excited state. (This will be covered in chapter 13.) For example, adding a substance
known to quench fluorescence would not prevent the unusual frequencies appearing
in the light emerging from the sample. Also, the shifts were found to have the same
values if a different input frequency was used—the entire spectrum shifts along with
the Rayleigh line. A model spectrum exhibiting the key features is shown in
figure 8.2, with features either side of a Rayleigh scattering line subsequently
designated as Stokes and anti-Stokes. It was Raman who first realised that such
observations could be accounted for by the scheme illustrated in figure 8.3.
In the mechanism responsible for the Stokes frequencies, the molecule scatters
light of a lower frequency, and hence lower energy than the absorbed photon. The
energy difference must promote the molecule to an excited (usually vibrational or
rovibrational) level in the ground electronic state: the difference in the initial and
final state energies is therefore small, compared to the photon energies. For energy
conservation we must have:
Stokes: hν′ = hν − ΔE , ∴ ν′ = ν − Δν. (8.3)
Figure 8.2. Schematic Raman spectrum, here showing the intensity of scattered light as a function of input
optical frequency. Lines typically appear in pairs of dissimilar intensity, with frequency shifts of equal
magnitude to the either side of a much more intense Rayleigh line (ν′ = ν). The high-frequency (anti-Stokes)
lines are less intense than their Stokes counterparts, and progressively diminish in intensity. In practice the
Rayleigh signal may be removed by a cut-off optical filter to save overwhelming the detector.
8-3
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 8.3. Illustration of Raman scattering processes connecting the ground state with just one selected upper
level, compared to Rayleigh scattering, showing the photon energies involved in each case.
In the anti-Stokes case, however, the molecule starts off in the excited vibrational
state and Raman scattering leaves the molecule in its ground state; we therefore have
a shift in the other direction, to a higher frequency:
anti−Stokes: hν′ = hν + ΔE , ∴ ν′ = ν + Δν . (8.4)
Hence the Stokes and anti-Stokes lines are shifted by the same amount Δν from the
Rayleigh line at frequency ν. Also, the intensity of anti-Stokes lines must fall off
rapidly with ΔE because the population of molecules initially in the excited level is
governed by the Boltzmann distribution e−ΔE /kT . As we saw in the previous chapters,
it is such vibrationally excited levels that are indeed responsible for the weak hot
bands in infrared absorption spectra.
It follows, from equations (8.3) and (8.4) that the various shifts in either a Stokes
or an anti-Stokes Raman spectrum can usually be attributed to transitions between
vibrational states, and directly identified with frequencies of molecular vibration.
However, we shall see that Raman spectroscopy is often capable of delivering
information that other spectroscopies such as infrared cannot.
Raman scattering is a very weak process—much weaker than Rayleigh scattering.
It should be noted that usually the input light is not resonant with an intermediate
excited state. Instead the scattering is said to proceed via a virtual state, a quantum
mechanical entity which is a superposition of all real eigenstates of the system. The
intensities of the Raman lines can be increased dramatically however, if the exciting
light frequency approaches an allowed transition in the sample; a process called
resonance Raman (section 8.4).
8-4
Molecular Photophysics and Spectroscopy (Second Edition)
What revolutionised the Raman effect as a spectroscopic method was the advent
of the laser, the obvious choice for a source of intense, monochromatic radiation.
While in principle any input light frequency ν can be used, it can be shown that the
Raman scattering intensity, like Rayleigh scattering, scales as ν4. For this reason, the
blue-green 488 nm or green 514 nm lines of a powerful Ar+ ion laser, were and still
are often used. However, these sources can also generate a lot of background
fluorescence noise. Popular alternatives today are to use a near-infrared 785 nm
diode laser, or the 1064 nm output of a solid-state Nd (neodymium):YAG (yttrium
aluminium garnet) laser which have high powers but tend not to excite sample
fluorescence. Unlike most Ar+ ion lasers, these light sources can be made relatively
compact, leading to the development of hand-held Raman spectrometers for a
variety of applications outside the lab environment. These include the identification
of narcotics, unknown powders, liquids gels and mixtures, pharmaceutical quality
control, and medical anti-counterfeiting.
In principle, Raman signals can also be produced by transitions between
electronic levels—but in practice it is almost always used in connection with
molecular vibrations, so this is what we shall focus on. Evaluation of the transition
dipole moment connecting states with quantum numbers v and v′ reveals that:
⎛μ ⎞
Mv ′ v ∝ ⎜ induced ⎟ (8.5)
⎝ dQ ⎠ Q
e
⎛ dα ⎞
Mv ′ v ∝ ⎜ ⎟ . (8.6)
⎝ dQ ⎠ Q
e
The derivative on the right-hand side of equation (8.6) is often written for shorthand
as α′. The key symmetry criterion for any specific vibration to generate a Raman
signal is thus as follows:
We also find that Mv ′ v is zero, and hence the transition is forbidden, unless
Δv = ±1 (Stokes, anti−Stokes) (8.7)
which, in this respect, is just as for infrared absorption. However, the dependence on
a change in polarizability, rather than dipole moment, is crucial, and it means that
8-5
Molecular Photophysics and Spectroscopy (Second Edition)
quite different vibrations appear in the Raman spectrum. (For small polyatomic
molecules, rotational structure is also exhibited in the vibrational Raman spectrum,
where the selection rules are Δν = ±1, ΔJ = 0, ±2.) Notice especially the requirement
for vibration-induced variation in polarizability. In contrast to a dipole moment,
every molecule has a non-zero equilibrium value polarizability.
Consider first diatomic molecules, with only one vibration. This motion must
always change the polarizability. In visualising polarizability it is often most helpful
and realistic to consider it as something akin to the shape of and a magnitude
proportional to the volume of the electron cloud around the molecule itself.
Consequently, it is immediately evident that even non-polar, symmetric molecules
(e.g. homonuclear diatomics) generate Raman signals. As we saw earlier, the anti-
Stokes line should be much weaker because it requires the molecule to start in an
excited state, and as we have seen in earlier chapters the room-temperature
population of vibrational excited states is generally small.
With polyatomic molecules we have to decide whether or not each vibration is
Raman-active. Figures 8.4–8.6 illustrate how the polarizability varies in each of the
fundamental vibrations of CO2. Clearly the ν1 symmetric stretch is Raman-allowed,
whereas in the ν2 and ν3 vibrations (antisymmetric stretch and bending modes), since
the change in polarizability is symmetric about the equilibrium position, the
derivatives vanish at that point. Hence the latter modes (as well as the counterpart
to the degenerate bending mode) are forbidden in the Raman spectrum.
As with infrared spectroscopy, it is often much easier to use character tables for
more complicated molecules, to establish which vibrational modes are Raman-
active or forbidden. Specifically, a Raman transition will be allowed if the
irreducible representation of the normal mode involved transforms in the same
way as one of the components of the binary product of vectors: x2, y2, z2, xy, xz, yz.
This can be read directly off the right-hand side of the character table.
Figure 8.4. Variation of molecular polarizability in the course of the symmetric stretch vibration of CO2.. The
polarizability changes uniformly during the symmetric stretch and the slope is non-zero at the equilibrium
configuration.
8-6
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 8.6. Polarizability of CO2 during one of its bending modes of vibration.
It was shown in section 6.4 that the ν2 (Πu) and ν3 (Σ u−) vibrations of CO2 are
infrared active but the ν1 (Σ g+ ) mode is not. Examination of the D∞h character
table shows, as did figures 8.4–8.6, that the opposite result holds; namely, ν1 is
Raman-active but ν2 and ν3 are not. With CO2 we therefore have a good illustration
of the mutual exclusion rule for centrosymmetric molecules, which states that
molecular vibrations that are allowed in the infrared spectrum are Raman-
forbidden, and vice versa. A word of caution, however: some vibrations may be
forbidden in both.
The mutual exclusion rule can be generalised as follows:
• Raman-active modes are gerade (g) and infrared inactive
• IR-active modes are of ungerade (u) and Raman inactive
In the present connection gerade means that the vibration looks exactly the same if
we perform a spatial inversion through the centre of symmetry. Applying this to the
vibrations of CO2 we can observe the symmetry principles illustrated in figure 8.7.
8-7
Molecular Photophysics and Spectroscopy (Second Edition)
Exercise
Three unknown substances, A, B and C, are available as samples: two are gases, one
a solid. It is known that each is a simple tetrahalide of the type MF4, where M
represents either carbon, sulfur or xenon. The Raman, infrared and microwave
absorption spectra of each are secured (in the case of the solid, by gently heating to
produce a vapour). Comparing the infrared and Raman vibrational spectra of A, it
is apparent that they contain no features in common. It also transpires that it is only
sample C that will produce a rotational spectrum.
8-8
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 8.8. Vibrational spectra of paracetamol (acetaminophen): (a) Stokes Raman scattering; (b) infrared
absorption. Both spectra courtesy of Nicolet Instruments Inc.
The origins of these differences in behaviour is that all three substances have
differently shaped molecules. CF4 is tetrahedral; SF4 is of trigonal bipyramidal
shape with one position (actually an equatorial one) occupied by a lone electron
pair, and XeF4 is square planar (octahedral with two opposing lone pairs, above and
below). Which of these are compounds A, B and C?
8-9
Molecular Photophysics and Spectroscopy (Second Edition)
Answer
The mutual exclusion of vibrational modes between the infrared and Raman spectra
of A indicates that this is a centrosymmetric molecule, and hence it has to be XeF4
(in fact, this is the solid substance). This principle does not apply to the
non-centrosymmetric CF4 or SF4. Since only C gives a rotational spectrum (and
the other two compounds are therefore presumably non-polar) it has to be SF4.
Hence B is CF4.
3
γ′ =
2 i, j (
∑ αij′ − α¯′δij . ) (8.11)
Figure 8.9. Scattering geometry for the usual measurement of depolarisation ratios. Against the lab-based
Cartesian frame the incident radiation is z-polarised and the light scattered at right-angles is resolved for
components with: (a) z- and (b) y- polarisation.
8-10
Molecular Photophysics and Spectroscopy (Second Edition)
In the summation in the latter equation, the indices i and j are each taken over x, y
and z; δij is the Kronecker delta whose value is 1 if i = j, 0 if i ≠ j. Physically, ᾱ′ is
essentially a measure of the distortion in mean polarizability, produced in the course
of a particular molecular vibration, whilst γ′ measures the corresponding vibration-
induced distortion of the polarizability shape.
For gases and liquids, it emerges that the value of ρ usually lies between 0 and ¾,
for vibrations that are totally symmetric (vibrations during which the symmetry of
the molecule is unchanged—technically transforming under the totally symmetric
representation of the point group: see chapter 3), but equal to ¾ for other vibrations
that lower the molecular symmetry (since ᾱ′ is then zero). Using this principle,
extensive information on the symmetry properties each Raman-active molecular
vibration can be obtained by measurement of the depolarisation ratios for each line
in the Raman spectrum.
Figure 8.10. Ultraviolet/visible absorption spectrum of myoglobin. Raman spectra, measured with input laser
radiation corresponding to three different wavelengths from the absorbing regions, display selective resonance
enhancement of the signals from specific protein components, in each case those closest to the corresponding
absorption centres. Trp and Tyr refer to the amino acids tryptophan and tyrosine, respectively, and the inset
depiction uses the simplified motif representation of protein structures that is common in biochemistry. Data
courtesy of Sanford Asher (University of Pittsburgh).
8-11
Molecular Photophysics and Spectroscopy (Second Edition)
8-12
IOP Publishing
Chapter 9
Electronic and vibrational states in
large molecules
9-2
Molecular Photophysics and Spectroscopy (Second Edition)
vibration integral that signifies the overlap between the vibrational wavefunctions
for the ground and excited state. The square of the latter overlap is called the
Franck–Condon (FC) factor. The electronic integral establishes the selection rules for
the electronic transition, and for this section we will assume that such transitions are
allowed. However, unlike infrared spectroscopy, where there are restrictions on Δv
within the ground state potential (chapter 6), in principle all vibronic transitions are
possible between the two electronic potential energy curves. The intensity of any (vf, vi)
vibronic band however is determined by the magnitude of its FC factor.
Consider figure 9.1(a) where, for generic diatomic potential energy curves the
absorption of light produces a transition from a ground electronic state vi = 0 to an
excited electronic state having a very similar equilibrium geometry. In this case the
largest FC factor will correspond to vf = 0 since both vi and vf have their maximum
vibrational wavefunction amplitude close to the equilibrium bond length, re. The
strongest transition will correspond to (vf,vi) = (0,0). All other transitions where
vf ≠0 will be weaker. On the other hand, if re of the excited state is less than that of
the ground state, or greater—as in figure 9.1(b) and (c) respectively, the strongest
vibronic absorption transitions will be to vf—levels whose outer turning points or
inner turning points, respectively, is closest to the ground state re. In these cases the
strongest transitions will be to (vf, 0) with vf > 0. The intensity of hot bands; that is,
absorptions from vi ⩾ 1, are still expected to be weak since the population of the
ground state vibrational levels is governed by Boltzmann statistics, as discussed
previously for infrared spectroscopy.
9-3
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 9.1. Electronic transitions (denoted by vertical arrows) occurring without change in instantaneous
nuclear configuration, for three cases: (a) an electronic excited state with similar equilibrium geometry to the
electronic ground state. The strongest transition (red arrow) connects the ground vibrational levels of the two
states (0,0). In principle a higher vibrational level of the upper state is accessible as a hot band such as (3,3)
(blue dotted arrow) from an initial level in the electronic ground state, itself vibrationally excited—though their
intensities will be weak due to the Boltzmann statistics. (b) For an electronic excited state with markedly
different geometry—a shorter equilibrium bond length in the upper state—upper vibrational levels are directly
accessible in transitions from the ground state, primarily landing at the far extreme of the vibration. (c) With
an upper electronic level having a longer equilibrium bond length, similar transitions occur, landing at the
most contracted point in the vibration.
9-4
Molecular Photophysics and Spectroscopy (Second Edition)
can expect longer progressions. An example might be (vupper, vlower) = (n,0), (n,1),
(n,2), … where n could be the vibrational quantum number of several excited state
levels. When vupper > 0, there is also the possibility of observing progressions in
different wavelength regions: those originating from the inner turning points of the
excited levels differing from those originating from the outer turning points. This is
due to three factors. The first is that typically emission takes place on the nanosecond
(10−9 s) timescale, while vibrational motions are faster (~10−12 s). Thus, even if an
upper level vupper > 0 is excited specifically at one turning point (say by using
monochromatic laser excitation), the strongest emission bands will still involve both
turning points. The second factor is that the potential energy curves involved are in
general asymmetric; transitions from a specific excited vibrational level will thus probe
different lower state geometries, and hence potentially different lower vibrational
levels with substantial FC factors. Lastly, as will be described later in detail in chapter
12, excited state relaxation from high vibrational levels to those closer to the bottom of
the potential is extremely rapid for larger molecules. Thus, the emission usually comes
at significantly longer wavelengths than those required to excite an absorption.
A sequence corresponds to a group of transitions with a common value of
Δv = vupper − vlower. An example of a sequence might be the grouping corresponding
to Δv = 0; that is, (vupper,vlower) = (0,0), (1,1), (2,2), (3,3), …. The frequencies will be
similar, but usually not the same if the vibrational frequency and anharmonicity in
the upper state differs from those in the lower state. It follows that the frequency
separations between the members of a sequence could come to either higher or lower
relative values, depending on those vibrational constant differences. For larger
molecules—especially in the condensed phase—the vibronic structure, while present,
will be usually extremely difficult if not impossible to resolve due to the high density
of iso-energetic excited vibrational states inhomogeneously broadened by solvent
collisions, or other stochastic variations in the local electronic environment.
For small molecules—diatomics in particular—the paucity of vibrational modes
is such that measuring the decreasing frequency gap between successive absorption
lines in a progression is relatively easy to resolve. Again, in principle, the gap will
approach zero (due to anharmonicity), close to the excited state dissociation limit.
Graphical extrapolation based on this principle thus enables an approximate value
for the dissociation energy to be determined. This is the Birge–Sponer method, noted
in chapter 6.
9-5
Molecular Photophysics and Spectroscopy (Second Edition)
complex, and so allowing electronic transitions to take place between the split levels.
In Ligand Field Theory one uses group theory to construct symmetry-adapted
molecular orbitals to explain the nature of the resultant molecular orbitals involving
the metal and ligands. Here, for simplicity, we will consider only the qualitative
Crystal-Field Theory, to obtain insight into the nature of the visible absorptions of
metal–ligand complexes.
Consider an ML6 octahedral complex (Oh point group) where M and L denote
the metal and ligand, respectively. In crystal field theory the ligands are approxi-
mated as point negative charges. When the metal ion is placed at the origin of a
coordinate system, its d z 2 and d x 2−y2 orbitals form a two-fold degenerate set (with eg
symmetry) that have lobes pointing along the axes directly at the ligands. This raises
the energy of these two orbitals, due to Coulombic repulsion. The remaining three
orbitals (dxy, dyz, dzx) form a degenerate set, with t2g symmetry, that lies lower in
energy because they are orientated between the axes and not directly at the ligands.
The d-orbital splitting patterns for complexes of lower symmetries can be deduced
by notionally removing specific ligands from the octahedral configuration, and
deciding if this yields a net stabilisation of some of the d-orbitals relative to others.
As shown in figure 9.2, for example, forming a square planar ML4 complex with D4h
symmetry, by removing the two ligands orientated along the z-axis, lowers the
energy of the d z 2 orbital in the eg set relative to d x 2−y2 , as well as the energy of the dyz
and dxz orbitals relative to dxy in the t2g set. Thus, the energy ordering of the
d-orbitals in such a case might be E(dxy, dyz) < E(d z 2 ) < E(dxy) < E(d x 2−y2 ). The exact
ordering, however, will depend on the metal ion and ligands involved.
The splitting between the eg and t2g set in an octahedral complex is denoted by Δo
and its magnitude is a strong function of the ligands making up the complex. This
splitting can be substantial. For example, Δo for the Ti(H2O)63+ ion has a magnitude
of 20 000 cm−1. Ligands that yield a large Δo value are called strong-field ligands
Figure 9.2. Crystal field splitting of the d-orbitals in an octahedral (Oh) and square planar (D4h) geometry. The
d-orbitals are labelled by their irreducible representations in the appropriate point group.
9-6
Molecular Photophysics and Spectroscopy (Second Edition)
while those that result in smaller splittings are called weak-field ligands. The ligands
can be empirically ordered in this regard for a given metal and geometry in an
empirical spectrochemical series. A partial list from strong-field to weak-field is as
follows:
CO ≈ CN− > NO2− >NCS− >H2O > OH− >F− >SCN− >Cl− >Br − >I−.
It should be noted that such complexes can involve more than one type of ligand,
and that the splittings also depend on the metal cation and its charge. The resultant
colours of these species either in solution or in the solid phase are often striking. For
example, aqueous solutions of common cobalt, nickel and copper compounds,
which contain the solvated species [Co(H2O)6]2+, [Ni(H2O)6]2+, and [Cu(H2O)6]2+
respectively, are red, green and blue, respectively, indicative of a decreasing Δo.
Figure 9.3 shows these and some of the other colours that such octahedral complexes
can exhibit in aqueous solutions. The d-orbitals fill according to the Hund’s rule,
such that the terms of highest spin multiplicity lie lowest in energy. Electrons will
spin-pair in the t2g set before filling the eg orbitals if the magnitude of Δo is large
(strong-field ligands). Complexes of this sort are termed low-spin systems.
Conversely, electrons will fill both the t2g and eg orbitals before spin-pairing if the
ligands are weak-field: these complexes are designated as high-spin systems.
UV–visible transitions in a ML6 complex involve promoting a d-electron from t2g
to eg. Formally however, d–d transitions are electric-dipole Laporte forbidden.
However, this rule can be relaxed through vibronic coupling, as the transition
connects states whose total wavefunctions include not only an electronic component,
but also a vibrational normal mode. If that normal mode has the correct symmetry
the transition can become allowed—a phenomenon known as the Herzberg–Teller
effect. Physically, the vibrational motion breaks the inversion symmetry of the
molecule. In the case of the transition metal octahedral complexes, the normal mode
must have u-symmetry. These absorptions tend to be weak, even if the metal
complexes have no inversion centre such as those with tetrahedral or trigonal
bipyramidal geometries.
Figure 9.3. Octahedral ML6 complexes formed by transition metal ions in aqueous solution, where the ligand
L is H2O. The d-shell configurations are: Ti3+ (d1); V2+ (d3); V3+ (d2); Cr2+ (d4); Cr3+ (d3); Mn2+ (d5); Fe2+ (d6);
Co2+ (d7); Ni2+ (d8); Cu2+ (d9).
9-7
Molecular Photophysics and Spectroscopy (Second Edition)
The strongest transition metal complex absorptions usually lie at higher energies,
in the UV, and they can be assigned to metal–ligand charge transfer (g ↔ u)
transitions. If an electron is transferred from an orbital lying principally on the
ligand to an orbital lying principally on the metal, or vice versa, the transitions are
termed a ligand-to-metal or metal-to-ligand charge transfer, respectively. These
transitions however can occur in the visible if the metal is easily oxidised (in the
redox sense of losing an electron) and the ligand easily reduced (gaining an electron);
if they do, they are often sufficiently intense enough to mask the crystal field
transitions. Notable in this regard are metal iodide complexes, because I− is readily
oxidised. Charge transfer transitions readily facilitate the analysis of ions such as Fe
and Mg in biologically important systems such as hemoglobin and chlorophyll,
respectively. The molecular structure and visible range absorption spectrum of
chlorophyll a—a key light-absorbing pigment in the process of photosynthesis—are
shown in figure 9.4.
The concept of vibronic coupling can be generalised. Let the irreducible
representation of the components of the electric-dipole moment operator be Γ(x,y,z).
Depending on the point group this could be a direct sum of two or more irreducible
representations. An electric-dipole forbidden transition between states ψi and ψf with
irreducible representations Γ(ψi) and Γ(ψf), respectively, will become allowed if
Γ(ψf ) ⊗ Γ(ψi ) ⊗ Γ(x , y, z ) ⊂ Γ(Qk ), where Γ(Qk) is the irreducible representation of
one of the normal modes of vibration of the molecule.
Vibronic coupling is not just important for transition metal complexes. It can also
be used to explain the absorption spectra of other types of molecules such as
benzene. Based solely on the excited state molecular orbitals of benzene, one might
expect only one absorption band near 184 nm in the vacuum ultraviolet region,
due to a 1E1u ← 1A1g transition. Experimentally, however, symmetry-forbidden
1
B2u ← 1Ag and 1B1u ← 1Ag transitions are observed near 200 nm and 260 nm in the
Figure 9.4. Chlorophyll a. (a) Molecular structure showing the complex coordination of the magnesium atom,
whose doubly charged state is indicated by the oxidation state II. Adapted from original by Yikrazuul <https://
commons.wikimedia.org/w/index.php?curid=7974375>. (b) Absorption spectrum in the visible range; absorp-
tion is principally in the blue and red regions, giving the compound its green colour (see chapter 10).
9-8
Molecular Photophysics and Spectroscopy (Second Edition)
UV, due to vibronic coupling with B2g and E2g normal modes, and as a result they
exhibit resolved vibrational fine structure. To complete the picture for benzene it
should be noted that there is also a very weak 3E1u ← 1A1g transition near 350 nm,
partially allowed through spin–orbit coupling.
9-9
IOP Publishing
Chapter 10
Electronic transitions, colours, and detection
10-2
Molecular Photophysics and Spectroscopy (Second Edition)
is the concentration of the absorbing species usually expressed in moles per litre, mol
L−1 or M. (The over-use of symbols such as α is ultimately unavoidable: in standard
usage the same symbol is used here as for polarizability in earlier chapters. However,
the context should always make it clear which meaning is intended.)
The key to the application of equation (10.1) is that the rate of intensity
diminution must equate to the rate at which absorption promotes molecules to
excited states. The simple measurement of intensity reduction as a quantitative
register of concentration for a specific analyte is known as photometry—the same
term may also be used for emission measurements. When the wavelength is varied so
that a spectral response can be recorded, the method is sometimes called spectro-
photometry. Here, we focus on absorption spectroscopy in the UV–visible range,
associated with electronic transitions.
Before proceeding further, it is worth noting that electronic absorption transitions
generally populate excited states that are, in general, no longer able to undergo
further resonant absorption of the incident radiation. Although this effectively leads
to a small reduction in the concentration of ground state molecules [C ], most of the
excited molecules rapidly return to the ground state by fluorescence (or less direct
methods, some of which are described in the next chapter): such emission occurs in
all directions. For each molecule, the mean interval between the arrival of successive
throughput photons is generally far longer than the excitation decay time, so the
ground state population is very little changed. Under high-intensity conditions, it is
possible that a sufficiently significant population difference could be established
between the ground and excited states—a phenomenon known as saturation, where
the effect of absorption no longer increases with the incident intensity. However,
while possible using high power lasers, this condition is almost never achieved using
UV–visible lamps, the usual light source in most common UV–visible spectrometers.
To continue: the solution to the differential equation (10.1) is a simple exponen-
tial decay function:
I = Ioe−αℓ[C ] (10.2)
where ℓ is the path length through the sample along the z-direction, usually
expressed in centimetres, cm. This result, generally known as Beer’s Law (sometimes
also the Beer–Lambert Law), expresses the intensity of light transmitted through the
sample in terms of the intensity I0 incident upon it. To facilitate expressions based on
logarithms to base 10, it is alternatively written as
I = Io10−εℓ[C ] (10.3)
where ε = α/2.303 is generally termed the extinction coefficient, or molar absorption
coefficient which has units of L M−1. Since the absorption usually covers a range of
wavelengths or frequencies, its value when integrated over the full extent of each
band is directly related to the square of the transition electric dipole for that band. If
we label the peak absorption frequency νf0 and the transition dipole μf0, we can
discover the following relation:
2π 2L(μf0)2 νf0 Lhνf0
∫ εdν = 3 ln 10hcε0
=
c ln 10
B. (10.4)
10-3
Molecular Photophysics and Spectroscopy (Second Edition)
Here, L is Avogadro number, and εo is the vacuum permittivity = 8.8542×10−12 J−1 C2 m−1;
the second expression on the right introduces and serves to define the ‘Einstein
B-coefficient’, whose significance we shall pursue in the next chapter.
Returning to equation (10.3), the product εℓ[C] is often given the symbol A, and
referred to as the absorbance or optical density (OD) of the sample; it is evidently
related to the transmittance T = I/I0 by
A = −log10T . (10.5)
Exercise
Show that the fractional loss of intensity of the light passing through the sample,
ΔI /I0 = 1 − 10−A where ΔI = I0 − I.
Answer
ΔI /I0 = (I0 − I )/I0 = 1 − (I /I0 ) = 1 − T = 1 − 10−A , taking 10 to the power of
each side of (10.4) in the last step. The result is an important basis for quantitative
spectrophotometry.
For example, suppose that on passage through a certain sample, a beam of light
suffers a 14.2% diminution of intensity. The absorbance or optical density A is
defined by I = 10−AI0 , so when I/I0 = 100% − 14.2% = 85.8% ≡ 0.858, then –A = log
(0.858) = –0.0665. Hence the optical density = 6.65×10−2. If the sample cell is 0.1 cm
long and the sample is a solution of 10−3 M concentration, then since A = εℓ[C ], we
have an extinction coefficient for the solute given by ε = Α/ℓ[C] = 665 M−1 cm−1.
Exercise
Oxyhaemoglobin, HbO2, has a molar extinction coefficient in aqueous solution with
very small values for wavelengths at the long-wavelength, red end of the visible
spectrum, but a huge maximum value of 52 430 M−1 cm−1 at around 414 nm.
Calculate the transmittance of light with this wavelength through a 1.5×10−3 M
solution of HbO2, in a 0.3 mm cell.
Answer
Using ℓ = 0.03 cm and the other data as given, T = 10−A = 10−εℓ[C ] = 0.0044, i.e.
about 0.44% transmission. It is an optically very dense sample.
While direct measurement of absorption is common, there can be issues with
sensitivity because one is measuring small reductions in light intensity against a high-
intensity light background. This can be particularly problematic if the intensity of
the light source fluctuates. Although this can be ameliorated by a double beam
configuration where one records the ratio of the transmitted beam to the intensity of
the beam before entering the sample, there are more sensitive techniques to infer
absorptions.
Many molecules fluoresce after absorbing light (see chapters 11 and 12), the
emission often covering a range of wavelengths. Hence, one could measure the total
fluorescence yield as a function of incident wavelength, without needing to resolve
10-4
Molecular Photophysics and Spectroscopy (Second Edition)
Table 10.1. Absorption wavelength maxima, λmax, and molar absorption coefficients, εmax, for various
polyenes; for m > 1 the successive (CH=CH) units are linked by C–C single bonds.
1 162 10 000
2 217 21 000
3 258 35 000
4 296 52 000
5 335 118 000
8 415 210 000
11 470 185 000
15 547 150 000
10-5
Molecular Photophysics and Spectroscopy (Second Edition)
electron confined inside a box of length d with zero potential inside but having
infinitely high potential walls at its boundaries. The energy of the electron in the box
is purely kinetic energy, but because the particle is confined, it is quantised with
energy eigenvalues En that depend on a quantum number, n:
n 2h 2
En = n = 1, 2, 3, … (10.6)
8me d 2
where me is the electron mass, and h is Planck’s constant. Such a system exhibits a
zero-point energy because n ≠ 0; that is, the particle never stops moving—a non-
classical result. The energy level separations ΔEn = En+1 − En scale as 2n + 1; that is,
they increase linearly with n.
Using these results, we can begin to understand the spectra of the polyenes by
modelling these compounds as a one-dimensional box whose length increases with
the number of double bonds, m. The box dimension, d, can be estimated by summing
the appropriate number of known average C–C and C=C bond lengths. Each of the
m double bonds in H(CH=CH)mH contributes two electrons to the energy levels of
the box up to n = m, taking the Pauli exclusion principle into account. The energy
level with n = m corresponds to the Highest Occupied Molecular Orbital (HOMO),
and the lowest energy transition will be from the HOMO to the Lowest Unoccupied
Molecular Orbital (LUMO) which has n = m + 1. As the lengths of the polyene
increases the peak absorption wavelength λmax shifts to the red, as expected for a
particle-in-a-box type system. The initially increasing values of the peak absorption
coefficients εmax, reflect the trend in a transition dipole that lies approximately along
the direction of the π-system. The simple model is remarkably successful in
identifying the main trends; the deviations evident from the values given in table 10.1
arise because the π-electrons of the polyene backbones are not completely delocal-
ised but still exhibit some influence of the alternating single and double bonds.
Furthermore, the box-length is determined more by the extent of the electron cloud
delocalisation than strictly by the dimension of nuclear framework of the molecule.
One naturally occurring conjugated molecule is β-carotene shown in figure 10.1
which is responsible for the distinctive orange colour of carrots. It consists of a long
chain of alternating single and double carbon–carbon bonds (9 double and 10 single
bonds), with a bulky cyclohexene ring on either end.
Substituents that increase the intensity and often the λmax of absorption peaks
when attached to a chromophore are called auxochromes. A set of empirical rules
exists for dienes that allow the absorption maximum of a compound to be
10-6
Molecular Photophysics and Spectroscopy (Second Edition)
determined for different additional chromophores present and are useful for
structure determination. Similar rules have also been deduced for conjugated
carbonyl compounds, and mono- and di-substituted benzenes.
Solvent polarity can also play a role in determining the position of peak maxima
in an absorption spectrum. This effect, called solvatochromism, arises because the
HOMO and LUMO involved in a UV–visible absorption have different polarities,
and therefore interact differently with a polar solvent. Bathochromic (red) shifts
result when the LUMO is stabilised in the solvent relative to the ground state.; this
effect is commonly observed for π* ← π transitions. Conversely, hypsochromic (blue)
shifts, which arise when the HOMO is preferentially stabilised by the solvent, are
common in π* ← n transitions.
10-7
IOP Publishing
Chapter 11
After light is absorbed: photophysics in an
excited electronic state
11-2
Molecular Photophysics and Spectroscopy (Second Edition)
In explaining the photophysics of excitation for any but the simplest molecules,
the widespread adaptation of potential energy diagrams designed for diatomic
species, with their single chemical bond, can lead to some misapprehensions—not
least being a failure to appreciate the huge number of bound electronic excited
states. Even Jablonski diagrams such as the example shown in figure 11.1 can be
misleading in this respect. Quantum mechanical calculations for most molecules
typically reveal tens or hundreds of electronic excited states, each with their own set
of vibrational level continua, all tightly packed together and overlapping on the
energy scale. (Moreover, if molecules are free to rotate then each vibrational level
will have its own set of much more closely spaced rotational levels—though in the
liquid phase, the proximity of other molecules means that steric effects usually wash
out the quantum structure, so that the rotational motions are essentially random.)
Fortunately, a degree of simplicity is afforded by the fact that usually only the lowest
energy and most widely spaced few states are involved in determining the most
important aspects of photophysical or photochemical behaviour, since these states
lie on the scale over which single-photon visible and ultraviolet photoexcitation
occurs.
The Jablonski diagram of figure 11.1 assumes the electronic structure of large
polyatomic organic compounds consists of two kinds of electronic states, singlets
and triplets. The singlet electronic states have closed shell electron configurations—
all electron spins are paired such that S = 0; thus we have (2S+1) = 1, and their S (for
singlet) designations are distinguished by subscripts, being labelled S0, S1, S2 etc.
When two electron spins are unpaired we find triplet states, (2S+1) = 3, thus labelled
Figure 11.1. Representative Jablonski diagram, its vertical scale denoting energy, showing singlet electronic
states on the left and triplets on the right, with vibrational levels represented as continua. The separation
between singlet and triplet states in this way are not meant to suggest they necessarily have different
geometries. In addition to labelled processes of absorption and fluorescence (spin-allowed) and phosphor-
escence (spin-forbidden), secondary absorption processes that can occur if the S1 or T1 excited states are
sufficiently long-lived are depicted by dotted blue wavy lines and arrows. The wavy horizontal lines indicate
optical input and output; dotted horizontal arrows denote intersystem crossing, and the wavy vertical lines
signify intramolecular vibrational relaxation.
11-3
Molecular Photophysics and Spectroscopy (Second Edition)
T1, T2, etc. The subscript index is ordered in terms of increasing minimum energy of
the electronic state. Although this is standard notation, one should remember that S
in the spin multiplicity is the total spin quantum number and not the label of a
singlet state. As there is only one electronic ground state, S0, the index for the triplet
states begins at n = 1; that is, all triplet states are excited states.
It needs to be appreciated how it is that the energy of Tn is less than the energy of
the corresponding Sn state. The reasons are two-fold. First, due to Hund’s Rule
discussed in chapter 2, the two electrons in a triplet state occupy spatially distinct
orbitals which reduces the repulsive Coulomb interaction between the electrons.
Second, there is an important quantum mechanical symmetry consideration that
must be accounted for in multielectron systems. As noted above, electrons are
fermions and their wavefunctions are antisymmetric with respect to interchange of
labels. In multielectron systems no two electrons can be distinguished from one
another, and therefore it follows that if two electrons had exactly the same
coordinates (spatial and spin) the required anti-symmetry would yield a wave-
function that is identically zero. This is the basis of the Pauli Exclusion Principle,
also encountered earlier, dictating that no two electrons can occupy the same region
of space with the same spin orientation. Quantum mechanical calculations of the
electron–electron energy of an antisymmetric wavefunction yields not only the
expected Coulomb repulsion, but also a new energy term called exchange which has
no classical analogue. Roughly speaking it can be thought of as the energy resulting
from ‘sharing’ electrons simultaneously in two different orbitals. For singlet states
the repulsive nature of this exchange raises the energy, while for triplet states it has
an energy-lowering, stabilising effect. Hence, we can understand why T1 states are
lower in energy than S1. (Molecular oxygen, although a diatomic, is a rare but
important exception: with two unpaired ½ spins in its ground state configuration,
the total spin of S = 1 dictates a threefold degeneracy; i.e., a triplet ground state.)
The primary criterion that determines the accessibility of higher energy states by
photoexcitation from the ground state is the electron spin relationship between the
states so connected. Specifically, the strongest transitions are those that are spin-
allowed ΔS = 0, (chapter 2). Indeed, absorption from the ground state S0 as
described by Beer’s Law is governed primarily by spin selection rules. Hence as
shown in figure 11.1, the strongest transitions are between S0 and S1, where the
population of S0 is depleted and S1 increased. While shorter wavelength input can
sometimes directly populate an S2 level, direct transitions to T1 are not indicated
because to first order they are spin-forbidden, even though T1 is lower in energy
than S1.
At a more granular level it is helpful to recall from statistical thermodynamics
that after photoexcitation, if equilibrium were attained, the most probable
distribution of energy within the molecule would be among the highest number
of rotational-vibrational-electronic degrees of freedom. Put another way, this
principle will disfavour conditions in which all or most of the energy is concentrated
in one particular mode of any relatively large molecule. The actual processes
involved in distributing energy within a photoexcited molecule and deactivation are
considered next.
11-4
Molecular Photophysics and Spectroscopy (Second Edition)
Absorption S0 → S1 ka[S0]
Fluorescence S1 → S0 + hν kf[S1]
IVR S1* → S1 kIVR[S1]
IC S1 → S0 kIC[S1]
ISC S1 → T 1 S
k ISC [S1]
Phosphorescence T1 → S0 + hν kp[T1]
ISC T 1 → S0 kISCT[T1]
Collisional quenching S1 + Q → S0 + Q* kQ[S1][Q]
11-5
Molecular Photophysics and Spectroscopy (Second Edition)
lower electronic state, at approximately the same rate as IVR. Consequently, the
processes of IVR and IC are sometimes not clearly distinguished from each other.
In principle, an excited molecule could, as a result of effective IVR and IC decay
channels, relax to its ground state without emitting any light at all.
Consider the absorption of the laser dye Rhodamine 6G, in aqueous solution
(figure 11.2). This is a molecule that exhibits strong fluorescence, and whose
geometry slightly differs in its ground and excited states. For reasons that have
been established above, the release of electronic energy by fluorescence generally
occurs from the lowest vibrational levels of the S1 electronic excited state, due to
IVR and IC. Accordingly, we expect that the transition will ‘land’ on higher
vibrational levels of the S0 potential energy surface at a point that lies vertically
below its starting point on the upper surface. This means that the photon energy
released in the process of fluorescent decay is usually smaller—and hence the
wavelength is usually longer—than for the initial absorption: this is the explanation
for the well-known Stokes shift.
In the case of almost any organic dye such as Rhodamine 6G, its solution
absorption spectrum partly overlaps its fluorescence spectrum—but it also covers a
range of wavelengths shorter those found in emission. Irradiation with broadband
radiation, such as the output of a conventional white or ultraviolet lamp, is capable
of exciting a dye solution across its entire absorption band. As a result, fluorescence
will also occur right across its emission spectrum. For this reason, dye lasers which
employ molecules like Rhodamine 6G as their active (lasing) medium incorporate a
dispersive element such as a grating to generate a tuneable monochromatic output.
Figure 11.2. Absorption spectrum (solid line) and fluorescence spectrum (dotted line) of the laser dye
Rhodamine 6G, in aqueous solution.
11-6
Molecular Photophysics and Spectroscopy (Second Edition)
Even if monochromatic laser radiation is to be used to excite the dye, any input
wavelength within the absorption band can be chosen; emission still occurs at a
range of wavelengths, but only those that are longer than the input—i.e., photons
with less energy than the input light.
Outside the laboratory, more familiar examples of the spectral shift associated
with fluorescence are afforded by the fluorescent pigments used in many paints,
marker pens etc, whose high visibility colours are a result of light absorbed across a
broad spectrum being re-emitted at longer wavelengths. The reason that orange and
yellow ‘Day-Glo’® colours appear brighter than any others is only in part due to
wavelength-variation in responsivity of the human eye; there is plenty of light with
shorter wavelengths in the daylight that produces the primary absorption—as
compared to fluorescent colours at the blue end of the spectrum which can capture
only a much smaller fraction of the energy in the daylight spectrum.
Returning to the mechanisms of excitation decay, in addition to fluorescence,
IVR and/or IC, after photoexcitation, the S1 population can often transfer into the
T1 triplet state by intersystem crossing (ISC). This is generally a slower process
(10−11−10−6 s range) compared to IVR and IC, due to the change in spin state
required—but it can still be competitive with fluorescence. The rate of ISC, kISC, is
governed by the Fermi’s Golden Rule, which enables it value to be estimated from the
following expression;
2π 2
kISC =
ℏ
∫ ψT⁎HˆSOψSdτ ρ(E T ) (11.4)
where ψS, ψT refer to the overlapping singlet and triplet states, ĤSO is the spin–orbit
coupling operator, and ρ(E T ) is the density of triplet states—i.e. the number of states
per unit energy interval. The magnitude of kISC increases when the vibrational levels
of S1 overlap those of T1, and also for molecules containing a ‘heavy’ (high atomic
number) atom such as Br or I, due to their stronger spin–orbit interactions that
weaken the distinction between singlet and triplet states.
Decay due to IVR relaxes the T1 state into its lowest vibrational levels—where
either non-radiative ISC to S0 could take place, or the molecule might emit light by a
spin-forbidden process called phosphorescence. The latter is relatively slow (10−4−10 s
or longer) because of the forbidden nature of the transition.
Finally, there is the possibility of an additional process called collisional quenching
(Q), which is decay due to interaction with a molecule of another type. For example
collisions between an excited molecule in state S1 and a quencher molecule Q, which
is usually in a singlet ground state, may lead to energy exchange, where the excited
molecule decays to S0 and Q acquires the spent energy in excitation to a typically
vibrationally excited state, labelled Q*. Q could be a solvent or solute molecule. This
deactivation pathway can be minimised by working at low temperatures or by using
viscous solutions. Notably, the presence of even minute traces of molecular oxygen,
O2, will often non-radiatively ‘quench’ (depopulate) T1 states, because their shared
triplet character allows them to strongly interact.
11-7
Molecular Photophysics and Spectroscopy (Second Edition)
11-8
IOP Publishing
Chapter 12
Molecular fluorescence
In the study of electronic excitation, by far the simplest direct measurement—to gain
insights into the photophysics—is molecular fluorescence. Even as a straightforward
measure of some prior absorption it has the virtue of presenting an essentially zero-
background signal. However, as we have seen, considerable intricacies are involved
in the processes of excited state decay and their interplay. Given the widely differing
time scales and rates associated with these excited state deactivation processes, it is
important to understand the kinetics involved more deeply and to establish their
influence on observing fluorescence after photoexcitation. In fact the fluorescence
spectrum, lifetime and decay, and polarisation all provide extremely valuable
information.
This, simply put, represents the fraction of [S1] giving rise to fluorescence. A
fluorescence yield that is large (approaching unity) suggests that the rate constants
for other deactivation processes are small; i.e. those processes are slow, relative to
fluorescence. Conversely, if the non-radiative processes have large rates relative to
fluorescence, the quantum efficiency will approach zero.
Collisional quenching fluorescence experiments allow kf to be determined. Here
we define the fluorescence lifetime, τf, as:
1 S
= kf + kIVR + kIC + k ISC + k Q[Q] (12.4)
τf
and kfτf = ϕf. Using equation (12.1), we therefore have:
d [S1] [S ]
= ka[S0] − 1 . (12.5)
dt τf
At steady-state the left-hand side of (12.5) vanishes, so that [S1] = ka[S0]τf. Then,
since the measured fluorescence intensity is If = kf[S1], it follows that:
If = ka[S0]kf τf . (12.6)
Inverting this equation, applying (12.4) now yields a result expressible as:
1 1 ⎛ S ⎞
kIVR + kIC + k ISC k Q[Q]
= ⎜1 + ⎟+ . (12.7)
If ka[S0] ⎝ kf ⎠ k ak f [S0 ]
In the absence of a quencher the second term in (12.7) disappears and the
fluorescence intensity secured under those conditions would be written as Ifo.
Thus, taking the ratio Ifo /If one obtains:
Ifo kQ
=1+ [Q]. (12.8)
If kf
A plot of the intensity ratio as a function of [Q] will yield a straight line with a slope
of kQ/kf and a y-axis intercept = 1. Such graphs are called Stern–Volmer plots. No
lifetime measurements are required for the adoption of this simple method.
By performing fluorescence lifetime measurements, the rate constants kQ and kf
can be separately determined to obtain τf, if fluorescence and collisional quenching
are the dominant processes taking place. Then, equation (12.4) simplifies to:
1
= kf + k Q[Q]. (12.9)
τf
12-2
Molecular Photophysics and Spectroscopy (Second Edition)
Here, a plot of the inverse excited state lifetimes as a function of [Q] yields a straight
line with slope kQ and y-intercept kf.
12-3
Molecular Photophysics and Spectroscopy (Second Edition)
12-4
Molecular Photophysics and Spectroscopy (Second Edition)
a useful indicator of such processes. One rare exception is the case θ = 54.7°, the so-
called magic angle, when the fluorescence is fully unpolarised (r = P = 0) even before
any rotational motion. (A similar phenomenon arises in a spin-decoupling method
in NMR, and it too is associated with a detachment of two coupling parameters.)
In certain cases, it might be considered potentially problematic to differentiate
fluorescence from Raman scattering, since each can produce radiation with a red-
shifted wavelength. However, there are several well-known ways to overcome this
problem; for example use can be made of the fact that the fluorescence signal will not
generally exhibit any shift in frequency when the excitation frequency is changed,
whereas the entire spectrum of Raman signals will shift by the same uniform
amount. A more sophisticated method is to use pulsed lasers and time-gating
techniques to separate the time-delayed fluorescence from any spontaneous Raman
emission.
One other feature that can play into the emission characteristics in a multi-
component system is a degree of energy transfer between molecules, prior to
fluorescence. Here too, not only the emission wavelengths but also the temporal
characteristics of the donor and acceptor fluorescence can usually be distinguished,
as we shall see in the next chapter.
Figure 12.1. (a) Chemical structure of Rhodamine 6G. (b) Solution of Rhodamine 6G in methanol excited by
a green laser (solid dye at the side). Adapted from https://exciton.luxottica.com/laser-dyes.html.
12-5
Molecular Photophysics and Spectroscopy (Second Edition)
12-6
Molecular Photophysics and Spectroscopy (Second Edition)
A material made up of nano-sized quantum dots has a larger surface area to volume
ratio compared to the same volume made up of bigger particles. These particles are
neither atoms nor molecules, but in several respects they do behave rather like
atoms.
It is instructive to briefly consider the electronic structure of semiconducting
solids, to better understand the optical properties of quantum dots. If a solid
material has a periodic atomic arrangement, the Pauli exclusion principle leads to
the formation of electronic energy bands; specifically, a lower lying valence band
composed of a continuous density of orbitals that contain the outer electrons from
each atom. This valence band is separated from a usually empty higher-lying
conduction band, and the energy separation between the two bands is called the
band gap, Eg. From a general point of view, if Eg is zero, the material is classified as
a metal, electrons can easily move under an applied voltage. If Eg is relatively small,
typically 1–4 eV, the material is classified as a semiconductor because heat, light or
electricity can readily promote electrons from the valence band to the conduction
band, where they can then form a current under an applied voltage. This ability to
switch semiconductor current states on and off is at the heart of how computer chips
operate. Lastly, materials having very large band gaps (~15 eV) are classified as
insulators.
Semiconducting quantum dots are sufficiently small that quantisation of the
valence and conduction band becomes important (figure 12.2), and the actual energy
of an optical transition (in both absorption and emission) will be affected by the
radius of the quantum dot in a manner analogous to a three-dimensional particle in
a box.
When an electron is promoted to the quantised conduction valence band from the
quantised valence band it leaves behind a hole which behaves like a positive charge.
In fact, the electron–hole pair, which is bound together Coulombically like an
Figure 12.2. Left: cartoon of the continuum conduction (lower) and valence (upper) bands of a regular
semiconductor (left), separated by an energy gap Eg: Right: quantised energy levels for the same material in a
quantum dot of nanoscale radius d.
12-7
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 12.3. Emission from cadmium selenide (CdSe) quantum dots in solutions with differing diameters: blue
(2.0 nm), green (2.5 nm), yellow (3.0 nm), orange (3.9 nm), and red (4.2 nm). Photo: Courtesy Dr Michael
Wong, Rice University.
12-8
IOP Publishing
Chapter 13
Fluorescence resonance energy transfer
Figure 13.1. Typical spectral discrimination between the fluorescent emission from donor and acceptor species
(notionally a cyan fluorescent protein donor and a yellow fluorescent protein acceptor): (a) the transmission
characteristics of a short-wavelength filter will ensure initial excitation of only the donor. Separating the
different wavelengths with a beam-splitter and another narrow emission filter ensures only the Stokes-shifted
fluorescence from the donor reaches a detector; (b) in the same system a longer-wavelength emission filter
ensures capture of only the acceptor fluorescence, following energy transfer.
where the asterisk denotes an electronic excited state and the dagger, vibrational
excitation. In systems where RET occurs, the donors and acceptors are usually
fluorophores, i.e. chromophores that with a capacity to exhibit significant fluores-
cence decay. Moreover, the transitions of donor decay and acceptor excitation are
generally electric dipole-allowed—although other possibilities occasionally arise.
Accordingly, the most common mechanism of energy transfer, for donor–acceptor
displacements beyond significant wavefunction overlap, is electrodynamical cou-
pling between transition dipoles. With an inverse cubic dependence on distance,
much like the form of dipole–dipole interaction between two static dipoles, the rate
of energy transfer depends on the square of this coupling, and hence an inverse sixth
power dependence on distance.
For any donor–acceptor separation, R, usually a distance substantially smaller than
the wavelengths of visible radiation—a distance regime known as the near-field—a
widely used Förster theory gives the following rate constant for RET:
9κ 2c′4
kRET =
64π 4τ A*R6
∫ FA(ν)σB(ν) dνν4 . (13.2)
Here, FA(ν) is the normalised fluorescence spectrum of the donor and τA* is the
associated radiative decay lifetime; σB(ν) is the absorption cross-section of the
acceptor, c′ the speed of light in the host material. The overall rate and efficiency of
RET thus involves a spectral overlap, the integrated frequency-weighted product of
donor emission and acceptor absorption profiles. The parameter κ2 is a numerical
orientation factor, its value often close to unity. As is apparent from the inverse sixth
power dependence on R, the likelihood for energy to transfer between chromophores
13-2
Molecular Photophysics and Spectroscopy (Second Edition)
is severely restricted by separation; any longer-range transfer therefore usually takes the
form of a series of small hops between near-neighbours, as indeed happens in
photosynthetic light-harvesting complexes. It is known that the distance dependence
actually changes beyond the near-field; a unified theory shows that it eventually acquires
the inverse square dependence associated with radiative transfer. However, in the most
common applications, distances are seldom more than a few nanometres—sufficiently
small for the Förster theory to be an extremely good approximation.
Figure 13.2. Resonance energy transfer from donor (left) to acceptor (right). The shaded boxes indicate the
vibrational continua of ground electronic (singlet) states S0 and relevant excited singlet states Sn—most often in
practice the first electronic excited state S1.
13-3
Molecular Photophysics and Spectroscopy (Second Edition)
Exercise
Consider a certain donor–acceptor pair separated by the Förster distance. It is found
that when the energy transfer acceptor moves away from the donor, to a separation
of 19 Å from the donor, the extent of its fluorescence due to FRET (fluorescence
resonance energy transfer) is diminished to about 0.1% of its occurrence at its
previous location. Find the Förster distance.
Answer
Acceptor fluorescence depends on (R0/R)6, so (R0/19 Å)6 = 10−3(R0/R0)6, hence
(R0/19 Å)6 = 10−3, thus (R0)6 = 10−3 × (19 Å)6 and so R0 = 0.316 × 19 Å = 6.00 Å.
In fact, the detailed distance dependence of RET provides a basis for the
identification of relative motions in molecules, or parts of molecules. There are
many examples in structural biology, such as the molecular traffic across a cell
membrane, and the mechanisms of protein folding. These and other such processes
can be registered by selectively exciting one chromophore, and observing the
generally longer-wavelength fluorescence from the other chromophore as it adopts
the role of acceptor—see figure 13.3. In cases where the components of interest do
Figure 13.3. Energy transfer for the detection of protein–protein interactions. Chemical interactions between
proteins result in the attached chromophores becoming closer in proximity, allowing transfer of energy
between them. Emission wavelengths differ when energy transfer is present.
13-4
Molecular Photophysics and Spectroscopy (Second Edition)
not have suitable overlapped energy levels, molecular tagging with site-specific
extrinsic (i.e. artificially attached) chromophores can solve the problem. Lanthanide
ions, with characteristically sharp absorption features, prove especially useful. In
some applications the actual distance between the chromophore groups is of specific
interest, and the scale of their separation can be quantitatively assessed by
comparing RET efficiencies. This technique is popularly known as a spectroscopic
ruler.
The polarisation features of FRET are also interesting to pursue. FRET that
occurs due to energy transfer between different units held at a fixed mutual
orientation within freely mobile molecules have polarisation features that conform
to the interpretations provided by equations (12.10) and (12.11) in the previous
chapter. If, however, the donor and acceptor are unattached and they can undergo
uncorrelated rotational motions, a degree of polarisation anisotropy can still persist
in the acceptor fluorescence, though its extent will be small and evident only at very
short times.
13-5
IOP Publishing
Chapter 14
Chiral phenomena and optical activity
Figure 14.1. Equivalence of parity inversion (top) to successive reflection and rotation (bottom).
14-2
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 14.2. In an optical transition of a (notional) chiral molecule, left, the magnetic dipole transition
moment (purple double-headed arrow) and its electric dipole counterpart (red arrow) may be co-aligned. The
former involves electronic angular momenta; the right-hand panel is a reminder that such rotational quantities
l = r x p are invariant to space inversion.
Figure 14.3. Depiction of the hexameric structure contained in a unit of human insulin protein, exhibiting the
threefold rotation symmetry. With no other symmetry elements present, this signifies C3 point group symmetry.
magnitude of the couplings due to the electric and magnetic terms in the above
expression is around the same value as the universal fine structure constant:
e2
α= (14.2)
4πε0ℏc
giving α ~ 1/137. In terms of contribution to an observed spectral line, the small
magnitude of the result means that magnetic coupling is usually unimportant, and
for most transitions in achiral (non-chiral) molecules it is not always allowed by the
selection rules, if the electric dipole coupling is allowed. In fact, due to its link with
charge circulation, transitions are only magnetic dipole allowed if the symmetry of
irreducible representation of the excited state is the same as that of a rotation, Rx, Ry
or Rz. But it turns out that for chiral molecules, x and Rx always appear in the same
representation, as do all other translation and rotation pairs: so for such molecules,
any electric dipole-allowed transition is also weakly magnetic dipole allowed.
An example of chirality for molecules in the pure rotation group C3, is the human
insulin hexamer shown in figure 14.3; the same symmetry class can also be found in
some inorganic complexes based on octahedral coordination. As is evident from the
corresponding character table, table 14.1, the rotations and translations behave in
14-3
Molecular Photophysics and Spectroscopy (Second Edition)
Table 14.1. Character table for the Schoenflies point group C3; ε = exp(2πi/3).
Linear
functions, Quadratic
C3 E C3 (C3)2 rotations functions Cubic functions
A +1 +1 +1 z, Rz x +y,z
2 2 2
z3, y(3x2 − y2), x(x2 − 3y2),
z(x2 + y2)
E +1 +ε +ε* x + iy; Rx + iRy (x2 − y2, xy) (yz, xz) (xz2,yz2)[xyz,z(x2 − y2)]
+1 +ε* +ε x − iy; Rx − iRy [x(x2 + y2), y(x2 + y2)]
precisely the same way under each of the allowed symmetry operations. Each of the
irreducible representations supports chirality.
We find that the interaction term for magnetic coupling with light carries an
opposite sign for molecules of opposite handedness. Moreover, for molecules of one
particular handedness, this form of coupling term takes a different sign for left- and
right-handed circular polarisations. As a result, the slight difference in absorption
strengths for left- and right-handed light can be quantified by what is known as the
dissymmetry factor, usually defined in terms of extinction coefficients (molar
absorption coefficients) ε:
ε L − εR
g= 1 L (14.3)
2
(ε + ε R)
in which the denominator represents the average absorption. Typical values are in
the range 10-3–10−4. The exact value usually varies across the absorption band, and
can be plotted as a circular dichroism spectrum. Such methods can be very useful for
the conformational analysis of organic compounds, with great importance for the
drug industry.
14-4
IOP Publishing
Chapter 15
Multiphoton absorption in molecules
Figure 15.1. (a) Two-photon absorption from a single input beam can populate a highly excited state that is
inaccessible by single-photon absorption, due to different section rules; the ensuing decay emission has a
wavelength shorter than the input; (b) consecutive absorptions of different wavelength photons also populates
a highly excited state, but the main channel of decay from the first excited state produces emission with a
wavelength longer than the input.
For two-photon absorption to occur, it is clearly necessary for two photons to pass
essentially simultaneously through the region of space occupied by one molecule (or
else the space occupied by a chromophore, in the case of localised absorption in a
large molecule). The likelihood of this depends on the intensity of light and the
volume, and can be estimated quite easily using the following formula:
n Iλ
= 2 (15.1)
V hc
where n is the mean number of photons in a volume V, given an irradiance I of light
with an optical wavelength λ. For any one molecule in a liquid the mean volume is
given by the molar volume divided by Avogadro’s number L, i.e. VM/L = M/Lρ,
where M is the molar mass, and ρ is the density. Hence, for the instantaneous mean
number of photons n̄ per molecular volume (which will be a number much less than
one) we have:
IMλ
n¯ = . (15.2)
Lhc 2ρ
Let us see, for example, how this formula might be applied to an experiment on a
sample of benzene, whose molar mass M = 7.811 × 10−2 kg mol−1, and STP
(standard temperature and pressure) density ρ = 8.765 × 102 kg m−3. Consider a
focused incoherent UV source delivering a continuous irradiance of 107 W m−2, at a
wavelength of 255.0 nm corresponding to the absorption maximum of the sample.
15-2
Molecular Photophysics and Spectroscopy (Second Edition)
Using equation (15.2) we find that the ‘photon density’ is approximately 6.3 × 10−12
photons per molecule, or one molecule in around 1.6 × 1011. With this level of
intensity, few molecules experience the transit of even a single photon at any instant
of time, though this is enough to produce a strong and easily measured absorption
signal: clearly the measured absorption could even be 100% if every photon were to
be absorbed by one of the molecules it encounters. However, the probability of two
photons simultaneously traversing a molecule, which depends on the square of the
irradiance, is approximately 4 × 10−23, corresponding to only one in 2.5 × 1022—
equivalent to only a couple of dozen molecules per mole. As experiment verifies,
there is no chance of observing any two-photon absorption at this level of intensity.
But consider a pulsed laser source operating at the same wavelength, and whose
beam is focused into the sample. Now, peak intensities of 1014 W m−2 can easily be
delivered to the sample. Whilst this increase by a factor of 107 results in a roughly
proportional increase in the probability of finding one photon in a molecular
volume, the probability of finding two photons increases by a factor of approx-
imately 1014, signifying approximately one molecule in around 2.5 × 108; easily
enough for a two-photon absorption process to be detected. This illustrates the fact
that with pulsed lasers, we may indeed expect to be able to detect two-photon
absorption in most media. Indeed results abound on systems ranging from single
atoms and inorganic crystals through to numerous complex biochemical structures.
Non-resonant two-photon absorption is related to the energy–time uncertainty
1
principle ΔE Δt ≥ ℏ. Consider again the case where two long-wavelength photons
4
together provide the energy for an electronic transition at the two-photon level.
Bearing in mind molecular dimensions, we must assume both photons are located at
the molecule within a brief interval Δt. This leads to a virtual state with a large ΔE
that is badly energy non-conserving when there is no real state near the single-
photon level. Accordingly, the molecule can exist within that virtual state only if a
second photon absorption restores energy conservation within a very short time, as
dictated by the time–energy uncertainty principle. This is why high light intensities
are needed to achieve an appreciable two-photon absorption.
15-3
Molecular Photophysics and Spectroscopy (Second Edition)
the exacting conditions for both photons to arrive at essentially the same instant. In
the limit where an exact one-photon resonance occurs and ΔE = 0, there is no longer
any temporal constraint, save for the decay lifetime of the resonant state, because
that state can be physically populated by single-photon absorption; however, doing
so will completely dominate and mask any two-photon absorption.
As with the original Kaiser and Garrett experiment, one observation that verifies
the population of an excited state through two-photon absorption is the subsequent
fluorescence at wavelengths much shorter than the input. To illustrate: two photons
of 700 nm visible laser light convey the same energy as one photon of 350 nm
ultraviolet radiation; fluorescence after IVR losses might for example occur at 450
nm, shorter than the input. With a filter used to cut out any fluorescence near to and
beyond the input wavelength, such substantially shorter wavelength fluorescence can
only result from multiphoton excitation.
Another sensitive detection method, particularly for gas-phase molecules, is to use
laser input of a wavelength such that the sum energy of three incident photons
exceeds the ionization limit of the molecule; one can then monitor the production of
electrons or molecular ions by non-resonant two-photon ionization. Detection of
ions using a mass spectrometer is particularly useful if there are more than one
absorbing species in the sample.
If a two-photon absorption populates an excited state, one can detect electrons or
ions in a process called resonance enhanced multiphoton ionization or REMPI. Two-
photon absorption followed by ionization is called one-colour (2 + 1) REMPI,
where the first number in the brackets indicates that two photons are required to
pump the molecule to the excited state, and the second number indicates that one
additional photon is required to ionize the molecule from the excited state. One-
colour simply means that all photons come from the same laser. In general, one can
have (n + m) REMPI depending on the lasers available and the molecule. Ionization
limits for typical organic molecules range from 7 to 13 eV. Therefore, one could use
a (1 + 1) or (2 + 1) scheme using UV radiation. On the other hand, a molecule such
as N2, having an ionization limit in excess of 15.5 eV would require (2 + 2) REMPI.
15-4
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 15.2. Two-photon absorption spectra of 131Xe132Xe with a Σ+g ground state: (top) plane polarised light;
(bottom) circularly polarised light. The bands which are reduced in intensity in the lower spectrum involve Σ+g
excited states. Adapted from Dimov S S et al 1995 Chem. Phys. Lett. 239 332–38, with permission of Elsevier,
Copyright 1995.
where Icp and Ilp are the two-photon signal intensities measured on excitation with
input light that is circularly polarised input (photon angular momentum = ±1), and
linearly polarised (photon mean angular momentum = 0), respectively. Detailed
analysis shows that the value of the ratio R0 is expected to be 1.5 for two-photon
transitions that take place between electronic states of different symmetry, but
significantly <1.5 for two-photon transitions between electronic states of the same
symmetry. This is yet again another manifestation of the conservation of angular
momentum. The benefit of polarisation studies is that insight into excited symme-
tries can be had in the absence of rotationally resolved vibronic spectra.
Since the rate of two-photon absorption depends quadratically on the irradiance,
but still linearly on the concentration [C ] of absorber, the associated rate of loss of
intensity of light passing through a distance ℓ in the absorbing medium along the
z-direction is given by a modified Beer’s Law equation:
dI
− = β(λ)I 2[C ] (15.4)
dz
15-5
Molecular Photophysics and Spectroscopy (Second Edition)
Hence the usual Beer’s Law does not hold for this process, nor do the concepts of
absorbance and molar absorption coefficient as normally defined.
One additional feature of interest feature is the possibility of Doppler-free
spectroscopy in the gas phase. If a mirror is placed such that the laser beam
traverses the sample in both directions, then each molecule of the sample is
effectively irradiated by two counterpropagating beams of wave-vectors k and –k,
as illustrated in figure 15.3. The two Doppler-shifted frequencies ‘seen’ by a sample
molecule travelling with velocity v and a component vk in the direction of the left-
hand beam have equal and opposite ‘red’ and ‘blue’ shifts, so that two-photon
transitions in which counterpropagating photons are absorbed provide an excitation
in which the Doppler shifts essentially cancel; ΔE = hνk′ + hν−′ k = 2hν . The result is
a readily resolved, sharp and symmetrical Doppler-free absorption line, on top of a
broader band due to the two-photon absorption of photons travelling in the same
direction, for which the usual Doppler shift occurs.
Figure 15.3. Two-photon absorption with counterpropagating beams, photons in each direction having
Doppler shifts of opposite sign.
15-6
Molecular Photophysics and Spectroscopy (Second Edition)
Exercise
Using the fact that the process of (non-resonant) three-photon absorption displays a
cubic dependence on the instantaneous intensity of an excitation beam, derive a
relationship for the dependence of intensity on the distance of beam travel within a
medium that exhibits three-photon absorption.
Answer
Using k for a constant of proportionality, I for the intensity, [C ] the concentration of
absorber and z the distance, we can write
dI
−
dz
= kI 3[C ] ⟹ − ∫ dI
I3
1
= ∫ k [C ]dz ⟹ 2 =
2I
k [C ]ℓ + const.
where ℓ is the pathlength through the absorber in the z-direction. At z = 0 the input
intensity can be defined as I0, so substituting in we find the constant is 1/2I02. Hence:
1 1 2kIo2[C ]ℓ + 1 Io
= k [C ]ℓ + = ⇒I=
2I 2 2Io2 2Io2 2kIo2[C ]ℓ +1
where the positive square root is selected since light intensity cannot be negative.
As noted above (REMPI) spectroscopy allows multiphoton absorptions to be
detected at significantly lower laser intensities, and that ion detection is particularly
sensitive. As the laser wavelength becomes shorter, the number of photons required
to promote molecules from their ground state to the ionisation continuum drops,
and the ion current increases accordingly. However, by far the largest signals are
obtained under resonance conditions when the energy of one, two or three photons
coincides with that of a bound state, as shown in figure 15.4. Fragmentation patterns
also typically change according to the excitation wavelength.
REMPI is not limited to the use of one laser. If two lasers are involved the process
is called two-colour REMPI. Here the notation is (n + m′) REMPI, where the first
number indicates the number of photons n from the laser used to excite the first
transition while the m′ is the number of photons from a second laser required to
ionize the molecule, following excitation. Typically, the first laser is tuned through
an electronic state to excite rovibronic transitions while the second laser has a fixed
wavelength sufficient to ionize the molecule, regardless of the wavelength range of
the first laser.
15-7
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 15.4. In this example of REMPI, as the laser frequency is tuned between νl and ν2, multiphoton
resonances (encircled) are observed at νA (three-photon), νB (one-photon) and νC (two-photon). The orange
circle identifies the threshold for thee-photon ionization. For example, at frequency νC, three photons are
responsible for the ionization process itself, but it is the two-photon resonance that results in a peak in the
spectrum; this is commonly represented as (2 + 1)-ionization.
15-8
Molecular Photophysics and Spectroscopy (Second Edition)
Figure 15.5. (a) Narrowing profile of intensity and its square, across a typical laser beam. (b) Three-
dimensional nanoscale model of a bull fabricated by Satoshi Kawata’s team at Osaka University in Japan,
based on two-photon photo-polymerisation of an acrylic gel at the programme-scanned focus of a laser beam.
The result won for the team a Guinness Book of Records prize for microfabrication.
Exercise
Consider a well-behaved laser beam whose radial distribution of intensity has a
2
Gaussian form, I (r ) = I (0)e−2(r /w) where 2w is defined as the beam waist (which is
the radial distance from the centre of the beam to where the intensity drops to 1/e2 of
its centre-beam value). For a focused beam with w = 720 nm, calculate by what
factor the rate of three-photon absorption is less at r = 400 nm, compared to the rate
at the beam centre.
Answer
The rate of three-photon absorption depends on the cube of intensity, i.e.
2 2
[I (0)e−2(r /w ) ]3 = I 3(0)e−6(r /w ) .
So, comparing three-photon at the beam centre and at a radial position r, the relative
rate is:
2 2 2
e−6(r /w ) / e−6(0/w ) = e−6(r /w ) /1.
15-9
Molecular Photophysics and Spectroscopy (Second Edition)
Hence with the given data we find e−6(1/1.8)2 = e−1.85 = 0.16. But even this
figure represents an upper limit for a process such as polymerisation triggered by
actual three-photon absorption.
The innovations in nanofabrication are typified by setups in which a laser beam is
focused into a photopolymerisable liquid such as an acrylic gel, and the focus is
scanned under pre-programmed computer control. In this way, a three-dimensional
solid of nanoscale dimensions can be sculpted, usually by two-photon activation that
is effective only at the moving focus. The reason this method works so well is that
there is an effective threshold intensity for photo-polymerisation to occur; no
excitation occurs below a cut-off intensity well within the transverse profile, because
the photon flux outside those regions is just not high enough to initiate a multi-
photon process. Here, there are emerging applications in the fabrication of nano-
scale motor parts for optofluidics and for lab-on-a-chip analytical applications.
There are other advantages to the use of multiphoton absorption—especially
three-photon absorption—in addition to its sharp focus. For in vivo applications
based on fluorescence imaging there are good clinical reasons to prefer using long-
wavelength light to produce the necessary photoexcitation, as opposed to potentially
damaging short-wavelength UV light; moreover, longer input wavelengths suffer
less scattering. Thus, with a scanning laser and imaging equipment, surface and
deeper sub-surface layers of skin can be scanned, and the ensuing molecular
fluorescence mapped out. This enables the imaging of local concentrations of key
fluorophores, including photodynamic dyes that are selectively absorbed by cancer-
ous tissue.
15-10
IOP Publishing
Bibliography
For further reading on the majority of the topics covered in this book, most of the
standard physical chemistry textbooks provide decent coverage, save for some of the
more advanced material in the later chapters of the present volume. The choice is
huge, and individual books have their own slant—but most are considerably more
mathematical in their approach. Beyond these, there are many more specialized
books on specific optical techniques and interactions, most of them aimed at a
graduate readership. The following list is therefore only an indicative selection of
titles worth investigating.
16-2
Index
I-2
Molecular Photophysics and Spectroscopy (Second Edition)
I-3
Molecular Photophysics and Spectroscopy (Second Edition)
I-4