i r Spectroscopy

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/362685177

Theory of IR Spectroscopy

Preprint · August 2022


DOI: 10.13140/RG.2.2.16495.92325

CITATIONS READS

0 3,707

1 author:

Md. Sajid Hasan


Ulsan National Institute of Science and Technology
6 PUBLICATIONS 3 CITATIONS

SEE PROFILE

All content following this page was uploaded by Md. Sajid Hasan on 14 August 2022.

The user has requested enhancement of the downloaded file.


Theory of IR Spectroscopy

Md. Sajid Hasan


PhD (Running), UNIST (South Korea)
MS (1st class 1st), MBSTU (Bangladesh)
BSc (1st class 2nd), MBSTU (Bangladesh)
A brief history of infrared spectroscopy
The journey of infrared spectroscopy begins with Sir Isaac Newton (1642-1727).
Although he became famous for his monumental work of the laws of mechanics,
Philosophiae Naturalis Principia Mathematica (Mathematical Principles of Natural
Philosophy), better known as Principia, published in Latin in 1687, the optics was
among his interests early on. In 1669, he delivered his first lecture at Cambridge
University about optics. It is very interesting to say that his first article published
in 1672 and his last book, Opticks, published in English in 1704, were studies about
the light.

Fig. Newton portrayed by Godfrey Kneller, Newton studying the refraction of light

The actual law of refraction was discovered in the early 1600s by Dutch
mathematician, Willebrord Snel van Royen, but Newton was the first to develop a
serie of elaborated and accurated experiments. Then he established a mathematical
rules for the experimental measurements of the refraction of light. Now one of the
experiments created by him, described here in brief as, the light of the sun, comes
into a dark room through a hole in the wall and passes through a prism. A screen
with several small holes allows the selection of short bands of light. One of these
bands of light passes through the hole and through a second prism. This prism does
not divides the light into new colors, just produces a selected color spot.
 Newton, Isaac. The Principia: Mathematical Principles of Natural Philosophy. University of California
Press, (1999)
 Newton, Isaac. The Optical Papers of Isaac Newton. Vol. 1: The Optical Lectures, 1670–1672, Cambridge
University Press (1984).
 http://www.quimica3d.com/ir/introduction.
Newton initiated the work about optics, but it was finally complemented by
Frederick William Herschel. It was him who discovered the infrared light. It is
very interesting to note that “great scientific discoveries can also be made by
ordinary people”. That was the case of William Herschel.
Friedrich William Herschel (1738-1822) was born in Hanover, Germany. He
played oboe in the military band in his country but in 1775 his regiment was
transferred to England. At that time, the crowns of England and Hanover were
united under the throne of George II.

Fig: Frederick William Herschel portrait by Lemuel Francis Abbott, The 40-foot
(12 m) telescope built in the backyard of Herschel
After that, this brief visit caused him a good impression and in the following year
he left the regiment and went to London. Then quickly he learned English and
changed his name to Frederick William Herschel. When he was 19 years old
devoted himself professionally to music and composed numerous musical pieces.
His interest, however, was not confined to music i.e., in his spare time Herschel
devoted himself to mathematics and astronomy. He built several telescopes in his
backyard and in 1781 discovered the planet Uranus. His ingenious knowledge on
mathematics and optics, allowed him to complement the work initiated by Newton.
When Herschel was working on sunlight, he conceive the existence of other
components of the sun light, outside the visible region.
 “Herschel discover infrared light” Cool Cosmos. Archived from the original on 25 February 2012. Retrieved
6 June 2018.
 Herschel, William (1800). “Experiments on the refrangibility of the invisible rays of the Sun”.
Philosophical Transactions of the Royal Society of London. 90: 284–292.
Since this region is invisible to the human eye, Herschel created an experiment to
detect it. By placing a blackened thermometer bulb in the beam of the light below
the red color (a), he noted that the temperature increased up to the room
temperature (b). Herschel defined that temperature as “reference temperature” and
called infrared to that part of the light spectrum.

a b

c d

Fig: Herschel experiment


Herschel made another discovery in March 1800. In this experiment he placed a
sample in the path of the infrared light (c), and he observed that by changing the
part of the spectrum that passed through the sample, in some points the temperature
suddenly decreased (d). Then he elicited two important point i.e., (1) Temperature
decreased because the sample absorbed infrared light and, for this reason, he
defined infrared spectroscopy as the “measurement of light absorption in the
infrared”. (2) Absorption was proportional to the difference between the references

 Rowan-Robinson, Michael (2013). Night Vision: Exploring the infrared Universe. Cambridge University Press. P. 23.
 Augton, Peter (2011). The story of Astronomy. New York: Qercus. Retrieved 17 may 2018.
 http://www.quimica3d.com/ir/introduction.
temperature and the temperature at the points in which occurred absorption. With
this experiment, Herschel created the principle that would be used in the infrared
spectrometer.
In the early 20th century, development of electronics was very rapid, that enabled
creating devices to automatically generate the frequencies present in the spectrum
of infrared radiation. The fully automated infrared spectrometer was started from
1937 but the first commercial spectrometer, was released in the 1960s. This
spectrometer was known as Michelson interferometer.
Introduction to infrared Spectroscopy
Infrared radiation (IR), sometimes called infrared light, is a type of electro-
magnetic radiation (a wave with electricity). This radiation have wave-lengths
longer than those of visible light and shorter than microwaves. The word infrared
comes from the Latin word “infra” (meaning below) and the English word “red”,
i.e., infrared light has a frequency below the frequency of red light. Infrared
radiation cannot be seen by the eye but we can feel as heat. It was discovered in
1800 by astronomer Sir William Herschel.

Regions within the infrared


There are three well defined IR regions (near, mid and far). The boundaries
between them are not clearly defined and debate still persists, but in brief, they are
defined in term of wavelengths and wavenumbers as follows:

Regions Wavelengths Wavenumbers


Near Infrared 0.7 µm to 2.5 µm 14000-4000 cm-1
Mid Infrared 2.5 µm to 25 µm 4000-400 cm-1
Far Infrared 25 µm to 1 mm 400-10 cm-1

The first region (NIR) allows the study of overtones and harmonic or combination
vibrations. The MIR region is to study the fundamental vibrations and the rotation-
vibration structure of most organic molecules, whereas the FIR region is for the
low heavy atom vibrations (metal-ligand or the lattice vibrations).
 G.Herzberg, Atomic spectra and atomic structure, Dover Books, New York, Academic press, 1969, 472 p.
 Maas, J.H. van der (1972) Basic Infrared Spectroscopy.2nd edition. London: Heyden & Son Ltd. 105p.
 http://www.infrared.simple.wikipedia.org.
Absorption spectroscopy
A spectroscopic method that uses the wavelength dependent absorption
characteristics of materials to identify and quantify specific substances. Examples
are UV-Vis, IR, Microwave, etc., spectroscopy. Generally a molecule absorb
radiation by three main process e.g., electronic transition (by high energy UV-Vis
radiation), vibrational transition (by low energy IR radiation) and rotational
transition (by lower energy microwave radiation). Thus the energy level can be
denoted in the following order:

Rotational < Vibrational < Electronic.

Infrared spectroscopy (or vibrational spectroscopy) involves the interaction of


infrared radiation with matter. It covers a range of techniques, mostly based on
absorption spectroscopy. It is the most commonly used spectroscopic method.
There are a number of reasons for its great success and dissemination. The method
is rapid, sensitive, easy to handle and provides many different sampling techniques
for gases, liquids and solids. Important aspects are the convenient qualitative and
quantitative evaluation of the spectra.

Why Infrared spectroscopy is called Absorption spectroscopy?

This analytical technique involves absorption of specific energy of


electromagnetic (IR) radiation which correspond exactly in energy to specific
excitations within the molecule being examined.

Why Infrared spectroscopy is called vibrational spectroscopy?

The lower frequency infrared radiation cannot promote electronic excitation, but
is sufficiently energetic to cause bond to deform (vibrate). The energy required to
stretch or bend a bond depends upon its force constant and this is in turn dependent
upon the constituent atoms and whether they are singly or multiply bonded.
 Laurence M. Harwood and Timothy D. W. Claridge, Introduction to Organic Spectroscopy, Oxford
University press,1996
 http://www.infraredspectroscopy.wikipedia.org.
Introductory theory

Electromagnetic radiation
Electromagnetic radiation includes, in addition to what we commonly refer to as
‘light’, radiation of longer and shorter wavelengths. Examples are Radio waves,
microwaves, infrared radiation, visible light, ultraviolet light etc. As the name
implies it contains both an electric and a magnetic component, which are best
illustrated by the following figure. Figure-1 illustrates one photon of such radiation
travelling along the x axis, the electric component of the radiation along the y axis
and the magnetic component along the z axis.

Electric component

Wave propagation z Magnetic component

Fig: Propagation of alternating electric forces and the related magnetic fields.

Following this approach, the energy associated with regions of the electromagnetic
spectrum is related to wavelengths and frequency by the equation E = hν = hc/λ
where h is Planck’s constant, c is velocity of light, ν is frequency of light and λ is
the wavelength. The shorter the wavelength (or higher the frequency) of
electromagnetic radiation, the higher its energy and vice versa. Organic molecules
absorb different wavelengths of electromagnetic radiation and undergo energetic
transitions as a consequence of energy transfer. However, the classical treatment
does not give us the full picture of the energy transfer process.

 J. Michael Hollas, Modern Spectroscopy, 4th ed. John Wiley & Sons Ltd, England, 2004.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Laurence M. Harwood and Timothy D. W. Claridge, Introduction to Organic Spectroscopy, Oxford
University press, 1996.
 William Kemp, Organic Spectroscopy, 3rd ed. PALGRAVE, Hampshire, New York, 1991.
The quantization of energy
In the quantum mechanical description, the quantization of energy refers to the fact
that at subatomic levels, energy is best thought of as occurring in discreet “packets”
called photons. Resemble to paper money, photons come in different
denominations. Quantum physics describes photons are as a packages of energy
and correspond to different colors in the spectrum or different types of
electromagnetic radiation (radio waves, microwaves, x-rays, etc). A red photon has
specific energy value different from a blue photon. For examples, the red and blue
photons are therefore “quantized” just as dollar bill denominations are “quantized”.
Thus we conclude that each photon contains a unique amount of discreet energy.
More technically, the “quantization” of energy is related to Plank's constant, which
specifies “how quantized” energy can get. The energy of each photon is derived
by the following relationship:

E = hν

The above whole description, implies that transfer of energy does not occur over a
continuous range. Thus, we can say excitation of an organic molecule involves
absorption of specific quanta of electromagnetic radiation. This quantized energy
is transferred to the molecule and promoted it to a higher energy level, with the
exact nature of the excitation depending upon the amount of electromagnetic
energy absorbed. Howsoever, in any event, the photon of electromagnetic radiation
will only be absorbed by a molecule if its energy corresponds exactly to an energy
difference between two states of this molecule. Now the above relationship can be
described by a simple extension of the former equation as:

∆E = hν
where ∆E is the difference between the higher and lower states of the absorbing
species. This theory describes that, any photon corresponding to any possible
transition within the molecule might be absorbed and the vast majority of the
molecules being irradiated will exist initially in the unexcited ground state.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Laurence M. Harwood and Timothy D. W. Claridge, Introduction to Organic Spectroscopy, Oxford
University press, 1996.
On the other hand the easiest, most efficient absorption process will be that
involving the lowest energy input, elevating the molecule to its first excited state.
The above description indicates that, the strongest absorption will occur at an
energy corresponding to transitions from the molecule in its ground state to the
first excited state of the transition (fig-1.2).

continuum

Absorption intensity
E4 E8

E3

E2

1st excited
state

Ground Frequency
state
Fig: 1.2 an Idealized representation of excitation transitions and how these are
translated into an absorption spectrum

This idealized representation of excitation transitions provide two kinds of


information which can enable the electronic and structural characteristics of the
molecule. They are: (a) the absorption wavelength or frequency gives the energy
associated with a particular excitation which can be related to the functional group
responsible for absorption. (b) Absorption intensity reflects both the case of the
transition (which once again can provide information about the functionality
undergoing excitation) and the concentration of the absorbing species.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Laurence M. Harwood and Timothy D. W. Claridge, Introduction to Organic Spectroscopy, Oxford
University press, 1996.
 William Kemp, Organic Spectroscopy, 3rd ed. PALGRAVE, Hampshire, New York, 1991.
Molecular Energy

A molecule in space may possess different sorts of energy; e.g., (1) rotational
energy: by virtue of bodily rotation about its center of gravity, (2) vibrational
energy: due to the periodic displacement of its atoms from their equilibrium
positions, (3) electronic energy: since the electrons associated with each atom or
bond are in unceasing motion, etc.

Molecular vibration

Molecular vibration can be defined as periodic motion of the atoms of a molecule


relative to each other, where the center of mass of the molecule remains
unchanged. A molecular vibration is excited when the amount of energy absorbed
by the molecule corresponds to the vibrational frequency according to the relation
ΔE = hν
where h is Planck’s constant. Generally the vibrational states of a molecule can be
pierced in a variety of ways. Among them, the most direct and suitable way is
infrared spectroscopy, as vibrational transitions typically require an amount of
energy that corresponds to the infrared region of the spectrum.

The vibrating diatomic molecule

We know that, elasticity of chemical bonds led to anomalous results in the


rotational spectra of rapidly rotating molecules where the bonds are stretched under
centrifugal forces. Now, in this section we will consider another consequence of
this elasticity for a vibrating diatomic molecule. At first, we will consider the case
of a diatomic molecule and the spectrum which arises if it contains only vibrational
motion. Then we shall deal more precisely with the practical case of a diatomic
molecule undergoing vibration and rotation simultaneously. Finally, we shall
prolong this study to more complex molecules.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Atkins, P. W.; Paula, J. de (2006). Physical Chemistry, 8th ed. New York: W. H. Freeman. p. 460.
 http://www.molecularvibration.wikipedia.org.
The Energy of a Diatomic Molecule
In order to form a stable covalent molecule (e.g., HF, HCI gas), two atoms here
combine with each other and they are said to be undergone some internal electronic
rearrangement. Here, we will simply look on the phenomenon as a balancing of
forces that exists within this molecules. In a simple covalent molecule, there is two
type of forces, they are (a) repulsive force: between the positively charged nuclei
of both atoms, and between the negative electron ‘clouds’, (b) attractive force:
between the nucleus of one atom and the electrons of the other, and vice versa.
Then two atoms colonize at a mean internuclear distance such that these forces are
just balanced and the energy of the entire system is at a minimal.
Energy

Internuclear distance

Fig-1: Parabolic curve of energy plotted against the extension or compression of a spring
obeying Hooke’s law.
When we squeeze the atoms more closely together, then the repulsive force rises
rapidly, again attempt to pull them further apart, we are resisted by the attractive
force. In either case, if we attempt to distort the bond, requires an input of energy
(fig-1). At the minimal, consider the internuclear distance as the equilibrium
distance req., or simply, as the bond length. The compression and extension of a

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 J. Michael Hollas, Modern Spectroscopy, 4th ed. John Wiley & Sons Ltd, England, 2004.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
bond may be compared to the behavior of a spring and we may extend this analogy
by assuming that the bond, like a spring, obeys Hooke’s law. Now we may write

𝐹 = −𝑘(𝑟 − 𝑟eq.) (3.1)

here F is the restoring force, k the force constant, and r the internuclear distance.
In this situation the energy curve is parabolic and takes the form as follows:
1
𝐸 = 2 𝑘(𝑟 − 𝑟eq.)2 (3.2)

The above model of a vibrating diatomic molecule is so called simple harmonic


oscillator model. Though it is an approximation, for the discussion of vibrational
spectra it will be an excellent starting point.

The Simple Harmonic Oscillator


In the above Fig-1, we have plotted the energy according to the equation (3.2). At
the point r = req. the zero of curve and equation is found, and any energy in excess
of this, e.g., E1, arises due to the extension or compression of the bond within the
molecule. Fig-1 implies that if one atom A is considered to be stationary on the r
= 0 axis, the other will oscillate between B and C. Now, if we increase the energy
to E2, the oscillation will become more vigorous i.e., the degree of compression or
extension will be greater, but the vibrational frequency will remain unchanged.
Every elastic bond, like a spring, has a certain vibration frequency dependent upon
the mass of the system and the force constant but independent of the amount of
distortion. Classical mechanical treatment easily show that the oscillation
frequency is:
1 𝑘
𝜔𝑜𝑠𝑐. = 2𝜋 √𝜇 Hz (3.3)

Where is K the force constant that varies from one bond to another, μ is the reduced
𝑚1 𝑚2
mass (= ) of the system. Now we can convert this frequency to wave
𝑚1 +𝑚2

numbers as:

− 1 𝑘
𝜔𝑜𝑠𝑐. = 2𝜋𝑐 √𝜇 cm-1 (3.4)

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Like all other molecular energies, vibrational energies (EV) are also quantized, and
the allowed vibrational energies for any particular system we can calculate from
the Schrödinger equation. In case of simple harmonic oscillator it takes the form:

1
𝐸𝜈 = (𝜈 + 2) ℎ𝜔𝑜𝑠𝑐. Joules, (ν = 0, 1, 2, 3….) (3.5)

where ν is called the vibrational quantum number. Now converting this, to the
spectroscopic units, cm-1 we get:

𝐸 −1
𝜀𝜈 = ℎ𝑐𝜈 = (𝜈 + 2) 𝜔𝑜𝑠𝑐. cm-1 (3.6)

The only energies allowed to a simple harmonic vibrator are shown in Fig. In
particular we should notice that the lowest vibrational energy can be obtained by
putting ν = 0 in (3.5) or (3.6) equation and we have:

1 1 −
𝐸𝑂 = ℎ𝜔𝑜𝑠𝑐. or 𝜀𝑂 = 𝜔𝑜𝑠𝑐. (3.7)
2 2

The above equation implies that the diatomic molecule and, indeed, any molecule
can never have zero vibrational energy i.e., the atoms can never be completely at
1 1 −
rest relative to each other. The quantity ℎ𝜔𝑜𝑠𝑐. joules or 𝜔 cm-1 is known
2 2 𝑜𝑠𝑐.

as the zero-point energy.


This anticipation of zero-point energy is the basic difference between the wave
mechanical and classical approaches to molecular vibrations. Classical mechanics
cannot explain how a molecule possesses vibrational energy but wave mechanics
insists that it must always vibrate to some extent. Further use of the Schrödinger
equation leads to the simple selection rule for the harmonic oscillator undergoing
vibrational changes:

∆ν = ±1 (3.8)

Now of course, we must add the condition that vibrational energy changes will
only give rise to an observable spectrum if the vibration involves a change in the

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
dipole moment of this molecule. Thus vibrational spectra will be remarkable only
in heteronuclear diatomic molecules since homonuclear molecules have no dipole
moment. Now apply the selection rule and we get:

− 1 − 1
𝜀𝜈+1 →1 = (𝜈 + 1 + 2) 𝜔𝑜𝑠𝑐. ─ (𝜈 + 2) 𝜔𝑜𝑠𝑐.

= 𝜔osc. cm-1 [for emission] 9(a)

𝜀𝜈 →𝜈+1 = 𝜔osc. cm-1 [for absorption] 9(b)

This simple result is also clear from Fig-2. Since the vibrational levels are equally
spaced, now the transitions between any two neighboring states will give rise to
the same energy change.

Fig-2 the allowed vibrational energy levels and transitions between them foe a diatomic
molecule undergoing simple harmonic motion.
Besides, the difference between energy levels expressed in cm-1 gives directly the
wave number of the spectral line absorbed or emitted as:

𝜈Spectroscopic = ԑ = 𝜔osc. cm-1 (10)

It is also clear that, in absorption, the vibrating molecule will absorb energy only
from radiation with which it can coherently interact and this must be radiation of
its own oscillation frequency.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Now we can get another expression for Hooke’s Law by combining the equation-
3.4 and equation-10 as follows:

1 𝑘
𝜈 = 2𝜋𝑐 √𝜇 cm-1

𝑘
= 4.12√𝜇 cm-1

Calculations of stretching frequencies for different types of bonds:

 C=C bond, given that K = 10 × 105 dynes/cm

𝑚 𝑚 12×12
Here, 𝜇 = 𝑚 1+𝑚2 = =6
1 2 12+12

𝑘 10×105
𝜈 = 4.12√𝜇 = 4.12 √ = 1682 cm-1 (calculated)
6

= 1650 cm−1 (experimental)

 C─H bond, given that K = 10 × 105 dynes/cm

𝑚 𝑚 12×1
Here, 𝜇 = 𝑚 1+𝑚2 = = 0.923
1 2 12+1

𝑘 10×105
𝜈 = 4.12√
𝜇
= 4.12 √ = 3032 cm-1 (calculated)
0.923

= 3000 cm−1 (experimental)

 C─D bond, given that K = 10 × 105 dynes/cm

𝑚 𝑚 12×2
Here, 𝜇 = 𝑚 1+𝑚2 = = 1.71
1 2 12+2

𝑘 10×105
𝜈 = 4.12√
𝜇
= 4.12 √ = 2228 cm-1 (calculated)
1.71

= 2206 cm−1 (experimental)

However, experimental and calculated values may vary considerably owing to


resonance, hybridization, and other effects that operate in organic molecules.
 D. L. Pavia, G. M. Lampman, G. S. Kriz and J. A. Vyuvan, Introduction to Spectroscopy, 5 th ed. Cengage
Learning, Stamford, USA, 2015.
The Anharmonic oscillator
In a real molecule, bonds are elastic but not so homogeneous as to obey Hooke’s
law. As a results, they do not obey exactly the laws of simple harmonic motion. If
the bond between atoms is stretched, in this situation, there comes a point at which
it will break i.e., the molecule will dissociate into atoms. Thus for small
compressions and extensions the bond may be considered as perfectly elastic but,
for larger amplitudes (e. g., greater than 10 per cent of the bond length) a much
more complicated behavior (fig-3) will be assumed.

1.5 Deq

Deq
Energy

0.5 Deq Deq

0.5 1.0 1.5 2.0 2.5 A


o

req. Internuclear distance

Fig-3 the Morse curve: the energy of a diatomic molecule undergoing enharmonic
Extensions and compressions.
This Fig-3 shows, diagrammatically, the shape of the energy curve for a typical
diatomic molecule (red), together with (black) the ideal, simple harmonic parabola.
Now a purely empirical expression which fits this curve to a good approximation
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
was derived by an American physicist, P. M. Morse (administrator and pioneer of
operations research in World War II) and is called the Morse function as:
E = 𝐷𝑒𝑞 [1─ exp {a (𝑟𝑒𝑞. ─ 𝑟 )}]2 (11)
where a is a constant for a particular molecule and 𝐷𝑒𝑞 is the dissociation energy.
If equation (11) is used instead of Eq. (3.2) in the Schrödinger equation, the pattern
of the allowed vibrational energy levels will be as in the form:
1 1 2
𝜀𝜈 = (𝜈 + ) 𝜔𝑒 ─ (𝜈 + ) 𝜔𝑒 𝒳𝑒 cm-1, (ν = 0, 1, 2, 3….) (12)
2 2

where 𝜔𝑒 is an oscillation frequency (expressed in wavenumbers) which we shall

1.5 Deq

Deq
8
7
Energy

6
5
4

0.5 Deq 3 Do Deq

1
0

0.5 1.0 1.5 2.0 2.5 A


o

req Internuclear distance


.
Fig-4: the allowed vibrational energy levels and some transitions between them
for a diatomic molecule undergoing anharmonic oscillations.

discuss more closely below, and 𝒳𝑒 is the corresponding anharmonicity constant


which is always small and positive (≈ + 0.01), for stretching vibrations. Thus we
can see that the vibrational levels crowd more closely together with increasing ν.
Some of these levels are sketched in fig-4. It should be noted that equation (12),
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
like (11), is an approximation only. Now we can say that, more precise express-
ions for the energy levels require cubic, quartic, etc., terms in (ν + 12) with
anharmonicity constants ye, ze etc., rapidly diminishing in magnitude. These terms
are significant only at large values of 𝜈, and we shall ignore them. Hence we
rewrite Eq. (12), for the anharmonic oscillator, as:

1 1
𝜀𝜈 = 𝜔𝑒 {1─ 𝒳𝑒 (𝜈 + )} (𝜈 + ) (13)
2 2

Now compare this equation (13) with the energy levels of the harmonic oscillator
equation (3.6), we have:
1
𝜔osc.= 𝜔𝑒 {1─ 𝒳𝑒 (𝜈 + )} (14)
2

Now we can say that the anharmonic oscillator behaves like the harmonic oscillator
but with an oscillation frequency which decreases steadily with increasing 𝜈.
1
Again, if we consider the hypothetical energy state obtained by putting 𝜈 = − 2

(at which, according to equation 13, ε = 0) the molecule would be at the


equilibrium point with zero vibrational energy. Hence its oscillation frequency (in
cm-1) would be as in the form:
𝜔osc.= 𝜔𝑒

From the above discussion 𝜔𝑒 may be defined as the (hypothetical) equilibrium


oscillation frequency of the anharmonic system i.e., the frequency for infinitely
small vibrations about the equilibrium point. Now at this situation, for any real
state specified by a positive integral 𝜈 the oscillation frequency will be given by
equation (14). So, in the ground state, where 𝜈 = 0, we would have:

1 1 1
𝜔o = 𝜔𝑒 (1─ 2 𝒳𝑒 ) cm-1 and 𝜀 o = 2 𝜔𝑒 (1─ 2 𝒳𝑒 ) cm-1

Here we see that the zero point energy differs slightly from that for the harmonic
oscillator equation (3.7). Thus the selection rules for the anharmonic oscillator
would be as in the form:
∆ν = ± 1, ± 2, ± 3…
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
In this manner they are the same as for the harmonic oscillator, with the additional
possibility of larger jumps. These, however, are predicted by theory and observed
in practice to be of rapidly diminishing probability and in fact only the lines of, ∆ν
= ± 1, ± 2, and ± 3 at the most, have remarkable intensity.
Also, the spacing between the vibrational levels is, as we will soon see, of order
103 cm-1. Now at room temperature, if we use the Boltzmann distribution equation
Nupper
{N = exp (-∆E/kT)}, we get:
lower

𝑁𝜈=1 6.63×10−34 ×3×1010 ×103


= exp {─ }
𝑁𝜈=0 1.38×10−23 ×300

≈ exp (- 4.8) ≈ 0.008

Furthermore, the population of the ν = 1 state is nearly 0.01 or some 1% of the


ground state population. In this way, for better approximation, we may ignore all
transitions originating at ν = 1 or more and restrict ourselves within the three
transitions:
 ν = 0 to ν = 1, ∆ν = +1, with considerable intensity
∆ε = 𝜀𝜈=1 − 𝜀𝜈=0
1 1 2 1 1
= (1 + 2) 𝜔𝑒 ─ 𝒳𝑒 (1 + 2) 𝜔𝑒 ─ {2 𝜔𝑒 ─ (2)2 𝒳𝑒 𝜔𝑒 }

= 𝜔𝑒 (1 ─ 2𝒳𝑒 ) cm-1 (15a)

 ν = 0 to ν = 2, ∆ν = +2, with small intensity


∆ε = 𝜀𝜈=2 − 𝜀𝜈=0
1 1 2 1 1
= (2 + 2) 𝜔𝑒 ─ 𝒳𝑒 (2 + 2) 𝜔𝑒 ─ {2 𝜔𝑒 ─ (2)2 𝒳𝑒 𝜔𝑒 }

= 2 𝜔𝑒 (1 ─ 3𝒳𝑒 ) cm-1 (15b)

 ν = 0 to ν = 3, ∆ν = +3, with normally negligible intensity


1 1 1
∆ε = (3 + 2) 𝜔𝑒 ─ {2 𝜔𝑒 ─ (2)2 𝒳𝑒 𝜔𝑒 }

= 3 𝜔𝑒 (1 ─ 4𝒳𝑒 ) cm-1 (15c)

The above three transitions have already been shown in Fig-4. Since 𝒳𝑒 ≈ 0.01,
these three spectral lines lie very close to 𝜔𝑒 , 2 𝜔𝑒 and 3 𝜔𝑒 . Here the line near
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
𝜔𝑒 is called the fundamental absorption, and those near 2 𝜔𝑒 and 3 𝜔𝑒 , are called
the first and second overtones, respectively.

n=4
n=3
Fig: The energy levels of
n=2
an anharmonic oscillator
n=1

n=0
Fundamental Overtones
Even if, we have neglect the transitions from ν = 1 to higher states, we should note
that, if the temperature is raised or if the vibration has a particularly low frequency,
the population of the ν = 1 state may become appreciable. Thus at about 300°C,
𝑁ν=1 /𝑁ν=0 becomes exp (-2.4) or about 0.09, and transitions from ν = 1 to ν = 2
will be some 10% the intensity of those from ν = 0 to ν = 1. A similar increase in
the excited state population would arise when the vibrational frequency are taken
to 500 cm-1 instead of 1000 cm-1. Now we can calculate the wavenumber of this
transition as:
 ν = 1 to ν = 2, ∆ν = +1, normally very weak
∆ε = 𝜀𝜈=2 − 𝜀𝜈=1
1 1 1 1
= (2 × 2 𝜔𝑒 ) ─ (6 × 4 𝒳𝑒 𝜔𝑒 )─ {1 × 2 𝜔𝑒 ─ 2 × 4 𝒳𝑒 𝜔𝑒 }
= 𝜔𝑒 (1 ─ 4𝒳𝑒 ) cm-1 (15d)

This above equation implies that, this weak absorption may arise. It will be found
close to and at slightly lower wavenumber than the fundamental (since 𝒳𝑒 is small
and positive). This weak absorptions are generally called hot bands as they
occurred only at high temperature. Again their nature may be confirmed by raising
the temperature of the sample while a true hot band will increase in intensity.
The above whole discussion was about vibration in a diatomic molecule, now we
will look at the phenomenon happened when it undergoes vibration and rotation
simultaneously.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
The diatomic vibrating rotator
In quantum chemistry the Born-Oppenheimer (BO) approximation is the
assumption that the motion of atomic nuclei and electrons in a molecule can be
treated separately. In molecular spectroscopy, using the BO approximation means
considering molecular energy as a sum of the independent terms e.g.,

𝜀𝑡𝑜𝑡𝑎𝑙 = 𝜀𝑒𝑙𝑒𝑐 + 𝜀𝑣𝑖𝑏 + 𝜀𝑟𝑜𝑡 + 𝜀𝑛𝑢𝑐𝑙𝑒𝑎𝑟 𝑠𝑝𝑖𝑛

Here the nuclear spin is so small that it is often omitted. Literature reveals that a
typical diatomic molecule has rotational energy separations of (1-10) cm-1 and the
vibrational energy separations of HCl molecule is nearly 3000 cm-1. As the
energies of this two motions are not same, we may consider an approximation that
a diatomic molecule can execute vibrations and rotations quite independently. This
approximation is tantamount to the Born-Oppenheimer (BO) approximation. Now
for combined rotational and vibrational case it is simply the sum of the separate
energies:

𝜀𝑡𝑜𝑡𝑎𝑙 = 𝜀𝑟𝑜𝑡. + 𝜀𝑣𝑖𝑏. (cm-1) (16)

Besides, the rotational energy levels for a Non-Rigid Rotator, can be calculated
from the Schrödinger wave equation as follows:

𝜀𝐽 = 𝐵𝐽(𝐽 + 1) − 𝐷𝐽2 (𝐽 + 1)2 + 𝐻𝐽3 (𝐽 + 1)3 ----- cm-1 (16a)

Where,
J = Rotational quantum number, (J = 0, 1, 2….).
B = Rotational constant, is given by B = h/8π2IB c (cm-1), here
IB is the moment of inertia, c is the velocity of light.
D = Centrifugal distortion constant, is given by
D=h3/32π4I2r2kc (cm-1), here k is force constant.
H = Is a small constant depends upon the geometry of the
molecule. It is negligible compare with D.

Now combining the separate equation (16a) and (12) respectively, we get the
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 M. Born, J. Oppenheimer, Quantum theory of molecule, Annalen Physik, German, 1927, 398 (20), 457-484.
Rotation- vibration energy expression from the equation (16), as follows:

𝜀 𝐽, 𝜈 = 𝜀𝐽 + 𝜀𝑣

= { 𝐵𝐽(𝐽 + 1) − 𝐷𝐽2 (𝐽 + 1)2 + 𝐻𝐽3 (𝐽 + 1)3 -- } +


1 1 2
{ (𝜈 + ) 𝜔𝑒 ─ (𝜈 + ) 𝜔𝑒 𝒳𝑒 } cm-1 (17)
2 2

Now, we shall neglect the small centrifugal distortion constants D. H, etc., and
hence we may write the above equation as:

𝜀 𝑡𝑜𝑡𝑎𝑙 = 𝜀 𝐽, 𝜈

1 1 2
= 𝐵𝐽(𝐽 + 1) + (𝜈 + 2) 𝜔𝑒 ─ (𝜈 + 2) 𝜔𝑒 𝒳𝑒 (18)

An important point should be noted here, that it is not logical to neglect D since
we are treating the molecule as rigid, yet vibrating! In fact, its retention in the
above equation would have only a very minor effect on the spectrum.

Jˈ = 6
5
4
3
Energy

2
1 0 ν=1

Jˈ̍ = 6
5
4
3
2
1
0
ν=0

Internuclear distance
Fig-5: The rotational energy levels for two different vibrational states of a diatomic molecule
In the above figure, the rotational levels are sketched for the two lowest vibrational
levels e.g., ν = 0 and ν = 1. As the separation between neighboring J values is in

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
fact, only some 1/1000 of that between the ν values, so here is no attempt at scale
in this diagram. Now it should be noted that, since the rotational constant B in
equation (18) is taken to be the same for all J and ν, the difference between two
levels of J is the same in the ν = 0 and ν = 1 states. Hence the selection rules for
the combined motions will be as:

∆ν = ± 1, ± 2 etc. ∆J = ± 1 (19)

In actual fact, we may also have ∆ν = 0, but this corresponds to the purely rotational
transitions. Again a diatomic molecule, except under a rare condition may not have
∆J = 0. In other words we can say that, a vibrational change muse be accompanied
by a simultaneous rotational change. In the Fig-5, we have drawn some of the
relevant energy levels and transitions, entitled the rotational quantum numbers as
Jˈ̍ (double prime for lower state) in the ν = 0 state and Jˈ (single prime for upper
state) in the ν = 1 state. Again we have to bear in mind that the rotational levels
Jˈ̍ are filled to varying degrees in any molecular population, so the transitions
shown will occur with varying intensities. This will be indicated in the spectrum
at the foot of Fig-6.
Now we will get an analytical expression for the spectrum by applying the
selection rules (equation-19) to the energy levels (equation-18). Taking into
account only the ν = 0 to ν = 1 transition, we get:

∆𝜀 𝐽, 𝜈 = 𝜀 𝐽ˈ, 𝜈=1 ─ 𝜀 𝐽ˈˈ, 𝜈=0

1 1 1 1
= 𝐵𝐽ˈ(𝐽ˈ + 1) + 1 2 𝜔𝑒 ─ 2 4 𝜔𝑒 𝒳𝑒 ─ {𝐵𝐽ˈˈ(𝐽ˈˈ + 1) + 2 𝜔𝑒 ─ 4 𝜔𝑒 𝒳𝑒 }

= 𝜔0 + 𝐵(𝐽ˈ─𝐽ˈˈ)(𝐽ˈ + 𝐽ˈˈ + 1) cm-1 [here, in brief, 𝜔0 for 𝜔𝑒 (1-2𝒳𝑒 )]

When we consider B to be identical in the upper and lower vibrational states then
rotation is unaffected by vibrational changes i.e., it corresponds to the Born-
Oppenheimer approximation.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
In case of combine motion, applying the selection rule (∆J = ± 1), we get two type
of energy expression as:

When ∆J = + 1, i.e., 𝐽ˈ = 𝐽ˈˈ + 1 or 𝐽ˈ─ 𝐽ˈˈ = +1

∆𝜀 𝐽, 𝜈 = 𝜔0 + 2𝐵(𝐽ˈˈ + 1) cm-1 𝐽ˈˈ = 0, 1, 2 … .. (20a)

When ∆J = ─ 1, i.e., 𝐽ˈˈ = 𝐽ˈ + 1 or 𝐽ˈ─ 𝐽ˈˈ = ─1

∆𝜀 𝐽, 𝜈 = 𝜔0 ─ 2𝐵(𝐽ˈ + 1) cm-1 𝐽ˈ = 0, 1, 2 … .. (20b)

∆J = ─1 ∆J = +1

4 ν=1

3
2
1
𝐽ˈ = 0

4 ν=0

3
2
1
𝐽ˈˈ = 0

2B 4B 2B

P 6 P 5 P 4 P 3 P 2 P1 Ro R1 R2 R3 R4 R5
𝜔0
cm-1
Fig-6: some transitions between the rotational-vibrational energy levels of a diatomic molecule
together with the spectrum arising from them.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Now combining the above two expression and replacing (𝐽ˈˈ + 1) and (𝐽ˈ + 1) into
“m” we have the following equation:

∆𝜀 𝐽, 𝜈 = 𝜈spect. = 𝜔0 + 2𝐵𝑚 cm-1 m = ± 1, ± 2…. (20c)

Note that, when ∆J = + 1, m takes positive value and if ∆J = ─ 1, m takes negative


value. It should be remember that “m” cannot be zero. Again the frequency 𝜔0 is
normally called the band origin or band center. The equation (20c) denotes the
combined vibration-rotation spectrum where, 2B the spacing from the band origin
𝜔0 to equally spaced band. As m ≠ 0, we will not get any line appearing at 𝜔0 .
Hence, the Lines to the lower frequency side of 𝜔0 i.e., negative m or ∆J = ─ 1 are
called the P branch while those for higher frequency side positive m or ∆J = + 1
are called the R branch. Also we may get another type of branches called (O, Q,
S) arising from (∆J = ─ 2, 0, + 2) respectively. Again if we include the centrifugal
distortion constant D to the (20c) equation we get the following expression:

∆𝜀 𝐽, 𝜈 = 𝜈spect. = 𝜔0 + 2𝐵𝑚 ─ 4𝐷𝑚3 cm-1 m = ± 1, ± 2, ± 3 … (20d)

We know that B is some 10 cm-1 or less and D is only some 0.01% of B. On the
other hand a good infrared spectrometer has a resolving power of about 0.5 cm-1.
So we can ignore D for high degree of accuracy. [m, O, P, Q, R, S are conventional
notation for spectrum]

Vibration rotation spectra of diatomic molecules (CO)


In this section we will look at the fundamental vibration-rotation band of carbon
monoxide under high and low resolution. In fig-7a, we have indicated some lines
named P and R branches according to their 𝐽ˈˈ values. From the fig-7a we consider
wavenumbers of the first four lines in each branch.

Line 𝝂 Separation (∆𝝂) Line 𝝂 Separation (∆𝝂)


P1 2139.43 R0 2147.08
3.88 3.78
P2 2135.55 R2 2150.86
3.92 3.73
P3 2131.63 R3 2154.59
P4 2127.68 3.95 R4 2158.31 3.72

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
At first, consider the fundamental vibration-rotation band of a diatomic
molecule e.g., carbon monoxide under high resolution are as follows:

R branch
P branch P3
Absorption

R2

P2 R1
P1 R0
0

cm-1

Fig-7: the fundamental vibration-rotation band of CO under high resolution in


gas pressure 100 mm Hg in a 10 cm cell.
Now the salient features of the above spectrum of carbon monoxide together with
the table as here: (a) all the lines are labelled in accordance with their 𝐽ˈˈ values. As
the sample CO contains 1% of 13C isotope, P branches are complicated at the band
center about 2100cm-1. (b) Band center (𝜔0 ) is at about 2143 cm-1 and the average
line separation (2B) near the center is 3.83 cm-1. (c) Decreases in separation
between the rotational lines with increasing wavenumber. That can be understood
from the table and the close wings of the spectrum.
∆𝝂

2050 2100 2150 2200 2250 cm-1

Fig-8: the fundamental band of carbon monoxide (CO) at lower resolution,

In this fig-8, the rotational fine structure is blurred out and it shows only the
fundamental band of fig-7 with a typical P and R contour view.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
At this stage, maximum intensity of transition will be occurred at J values of
1
√𝑘𝑇/2𝐵ℎ𝑐 ─ 2 . Again, m = J+1, and so we get from equation (20c) as:

1
m = ± √𝑘𝑇/2𝐵ℎ𝑐 + 2
1
𝜈max.intensity. = 𝜔0 ± 2𝐵(± √𝑘𝑇/2𝐵ℎ𝑐 + 2)

Where + and ─ signs indicate the R and P branches, respectively. Now the
separation between the two maxima (∆𝜈) is:

1
∆𝜈 = 4𝐵(± √𝑘𝑇/2𝐵ℎ𝑐 + 2)

= √8𝑘𝑇𝐵/ℎ𝑐 + 2B

= √8𝑘𝑇𝐵/ℎ𝑐 [as b is small compared with ∆𝜈]

From the Figure-8 and the table we noticed that, the separation is about 55 cm-1,
the band origin is at the midpoint of P1 and R0 i.e., 2143 cm-1. Then the first
overtone at 𝜔𝑒 = 2169.74 cm-1 and 𝒳𝑒 = 0.0061 will be as: 2 𝜔𝑒 (1 ─ 3𝒳𝑒 ) = 4260
cm-1

Break down of Born-Oppenheimer Approximation

This Approximation is valid only when vibration and rotation proceeds quite
independently of each other. We know during a single rotation, a molecule vibrates
1000 times. And so the bond length, moment of inertia, and constant B also
changes during this rotation.
For simple harmonic motion, the mean bond length or equilibrium bond length
does not changes with vibrational energy (fig-3.1). Howsoever, for non-rigid
rotator
1
(B ∞ )
𝑟2

but the average value of this quantity is not same as the 1/r2eq where req is
equilibrium bond length. As the vibrational amplitude increases with vibrational
energy, the value of B will rely on the quantum number.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
On the other hand, for anharmonic vibration, mean bond length increases with
vibrational energy and hence the rotational constant B also varies significantly. As
rav increases with vibrational energy, B is smaller in upper vibrational state than in
the lower and takes the form:
1
𝐵𝜈 = 𝐵𝑒 ─ α (ν + 2) (21)

Where 𝐵𝜈 , 𝐵𝑒 are the rotational constant in vibrational (ν), and equilibrium level
and α is small positive constant for each molecule. Now we will just look at the
fundamental vibration (ν = 0 to ν = 1). Taking B values as B0 and B1 (B0 > B1),
we may have:

∆𝜀 = 𝜀 𝐽ˈ, 𝜈=1 ─ 𝜀 𝐽ˈˈ, 𝜈=0

= 𝜔0 + 𝐵1 𝐽ˈ(𝐽ˈ + 1) ─ 𝐵0 𝐽ˈˈ(𝐽ˈˈ + 1) cm-1

Again consider two cases (𝐽ˈ or 𝐽ˈˈ = 0, 1, 2...) for the energy expression as:

Case-1: For R branch, ∆J = + 1 or 𝑱ˈ = 𝑱ˈˈ + 𝟏

∆𝜀 = 𝜈𝑅 = 𝜔0 + (𝐵1 + 𝐵0 )(𝐽ˈˈ + 1) + (𝐵1 ─𝐵0 )(𝐽ˈˈ + 1)2 cm-1 (22a)

Case-2: For P branch, ∆J = ─ 1 or 𝑱ˈˈ = 𝑱ˈ + 𝟏

∆𝜀 = 𝜈𝑃 = 𝜔0 ─ (𝐵1 + 𝐵0 )(𝐽ˈ + 1) + (𝐵1 ─𝐵0 )(𝐽ˈ + 1)2 cm-1 (22b)


Where 𝜈𝑅 and 𝜈𝑃 are the wavenumbers for R and P branches. Now combine this
two equation we get the following form:

𝜈𝑃,𝑅 = 𝜔0 + (𝐵1 + 𝐵0 )𝑚 + (𝐵1 ─𝐵0 ) 𝑚2 cm-1 [m = ± 1, ± 2...] (23)

In equation (23), when m is positive it refers to R branch and negative to P branch.


From the above discussion, now we can conclude that:
Ignoring vibration- rotation interaction (B0 = B1), equation (23) simplifies to
equation (20c). As B0 > B1, the term (𝐵1 ─𝐵0 ) 𝑚2 is always negative and has some
effect on the spectrum i.e., (a) increasing positive m, rotational lines crowded more
closely on the R branch side, (b) increasing negative m, lines become more widely
spaced on the P branch side. This effect is marked only for high m values.
 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.
 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Theoretically the calculated values of B0, B1 and 𝐵𝑒 for CO molecule are 1.915,
1.898 and 1.924 cm-1 respectively. Now the bond length at equilibrium, at ν =0
and ν =1 states will be as req = 0.1130 nm, r0 = 0.1133 nm and r1 = 0.1136 nm.

Hints:
 I = h/8π2Bc  C = 12× 1.67343 × 10-27kg

 μ =m1× m2 /m1+m2  O = 16× 1.67343 × 10-27kg

 r = √( I/μ)  Absolute mass of H atom


1.67343 × 10-27kg

Finally we can say that Born-Oppenheimer Approximation will be invalidated


when there is simultaneous interaction of rotations and vibrations.

Vibrations in a polyatomic molecule


To understand the vibrational mode in a polyatomic molecule, first of all, we have
to learn the following topics:

 Fundamental vibrations and their symmetry.

 Overtone, Combination and difference bands

 Fermi resonance

Fundamental vibrations and their symmetry

Suppose, a molecule have N number of atoms. Now let the position of each atom
by the three Cartesian coordinates e.g., x, y and z. Hence, the total number of
coordinate value for the molecule is 3N and we may say the molecule has 3N
degree of freedom. If, once all the 3N coordinates have been fixed, the bond angle
and bond distance will also be fixed. But the molecule can be moved freely without
changing its shape in this three-dimensional space. First of all, we get the
translational movement that uses three of the 3N degree of freedom leaving 3N─3.
Secondly, the rotation of a non-linear molecule can be determined about three
perpendicular axes: here these axes requires three degree of freedom so the
molecule is left with 3N─6 degree of freedom. Thirdly, for a non- linear N atomic

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
molecule may have internal vibrations of 3N─6. Thus the fundamental vibration
in non-linear molecule is:
3N─6

In case of linear molecule, there is no rotation about the bond axis and so here only
have two degrees of rotational freedom. Now the fundamental vibration in a linear
molecule is:
3N─5

We know that an N atomic molecule (acyclic) has N─1 bonds so its bond stretching
motion is N─1 and bending motions are 2N─5 [non-linear], 2N─4 [linear].

Example-1: Consider a linear diatomic molecule (CO) where, N = 2, so, 3N─5 =


1 this implies that the molecule have only one fundamental vibration. This 3N─5
rule does not say anything about the presence, absence or intensity of overtone
vibration. But this fact can be explained by the anharmonicity.

C2 axis

Symmetric stretch (ν1) Symmetric bend ((ν2) Antisymmetric stretch ((ν3)

Fig-9: Three fundamental vibration of the water molecule

Example-2: Consider a non-linear triatomic molecule (H2O) where, N = 3, so,


3N─6 = 3 this means that the molecule have three fundamental or normal
vibration: where normal vibration can be defined as a molecular motion in which
all the atoms move in phase and with the same frequency.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Dipole change during vibration

Now we will discuss the answer of the most important question associated with the
IR spectroscopy is: condition for a molecule to be IR active. The answer is that,
there must be a change in dipole moment as a result of the vibration that occurs
when the IR radiation is absorbed. This dipole change may take place either
parallel or perpendicular to the line of the symmetry axis.

Vibrational Distorted Normal Distorted


mode molecule molecule molecule

O
O O
ν1
H H
H H
H H
dipole

O O O
ν2
H H H H
H H

dipole

O O O
ν3 H
H H
H H
H

dipole

Fig-10: change in the electric dipole moment produced by each vibration of H2O molecule.

The fig-10 shows the nature of dipole change for the three vibration of H2O
molecule. Here conventionally, the vibrational frequency decreases as ν3 > ν1 > ν2.

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 Struve, S. Walter, Fundamentals of Molecular Spectroscopy, John Wiley & Sons Ltd, Canada, 1989.
Example-3: Consider a linear triatomic molecule (CO2) where, N = 3, so, 3N─5
= 4, hence we can expect four vibrational mode e.g., two stretching (symmetric &
antisymmetric) and two bending (in the plane & out of the plane, these two motions
are identical and also called them degenerated). Again, we get two set of symmetry
axes (C2, right angle to bond direction & C∞, parallel to bond direction) for CO2
molecule.

C2 axis

C∞ axis

Symmetric stretch (ν1) Antisymmetric stretch (ν3) Bending (ν2)

Fig-9: The symmetry and fundamental vibrations of the CO2 molecule

Although the permanent dipole moment of CO2 is zero, it is IR active. This is


because, when it absorbs IR radiation, it’s bending vibration and Antisymmetric
stretching involves rhythmical change in the dipole moment. Thus it is IR active.

Now we will discuss a little bit more about the modes of vibrational motion for a
molecule. We know a bond within a molecule may vibrates with both stretching
and bending modes. Here the stretching mode can be defined as the vibration that
occurs along the line of the bond i.e., the distance between the two atoms increases
or decreases. Undergoing stretching vibration bond length changes within this
molecule. On the other hand, the bending mode is a type of vibration that occurs
along the bond angle i.e., here the distance between the two atoms remain constant
but the bond angle changes within the molecule. All diatomic molecule only
 D. L. Pavia, G. M. Lampman, G. S. Kriz and J. A. Vyuvan, Introduction to Spectroscopy, 5 th ed. Cengage
Learning, Stamford, USA, 2015.
 Paula Y. B. Organic chemistry, 8th ed. PEARSON, California, USA, 2016.
undergoes stretching vibration since they have no bond angles. This the stretching
vibrational mode can be symmetric or asymmetric. In symmetric stretching, the
two atoms simultaneously move toward and away from the central atom while in
asymmetric stretching one of the atoms move toward the central atom and the other
moves away from the central atom.

Symmetric stretching Scissoring Rocking

Asymmetric stretching Wagging Twisting

Figure-1: stretching and bending vibration of methane molecule

Again the bending modes are often referred to as scissoring, rocking, wagging, and
twisting. Now we will shortly discuss this four type of bending vibration as:
Scissoring: In-plane bending where both atoms simultaneously moves either
toward each other or away from each other.
Rocking: In-plane bending where two atoms swing back and forth, i.e., they
approach to each other.
Wagging: Out-of-plane bending where both atoms move up to below the plane
with respect to the central atom.
Twisting: out-of-plane, where one atom moves forward while the other moves
backward.
 D. L. Pavia, G. M. Lampman, G. S. Kriz and J. A. Vyuvan, Introduction to Spectroscopy, 5th ed. Cengage
Learning, Stamford, USA, 2015.
 Paula Y. B. Organic chemistry, 8th ed. PEARSON, California, USA, 2016.
Asymmetric stretching needs more energy than symmetric stretching, why?
This is due to the fact of the displacement of the central atom. As in asymmetric
vibration, one atom comes closer to the central atom and the other atom departs
from it. In this process the central atom displace from its original position. Now to
get back this central atom in its position some amount of extra energy is needed
which is not required in symmetric vibration. And so asymmetric stretching always
needs more energy than the symmetric stretching.
Stretching vibration has higher frequency than bending vibration, why?

We know in a stretching vibration, the atoms move in and out along the inter-
nuclear axis and here this nuclei moves apart against the attraction of the bonding
electrons between them. On the other hand, in a bending vibration bond length
does not change instead, the bond angle changes. So it is easier to bend a bond than
to stretch because the nuclei are not moving against the attraction of the bonding
electrons. So Stretching vibration nessds frequency than bending vibration.

Overtone and Combination bands


To understand the overtone once again it is necessary to denote the fundamental
vibration. We know the vibration that involves the absorption of IR radiation to
promote a molecule from the ground state to its first vibrationally excited state is
called fundamental vibration.

Now the Overtones that result from excitation from the ground state (ν =0) to
higher energy states (ν = 2, 3...). Overtone bands are integral multiples of the
frequency of the fundamental absorption e.g., infrared absorption of 600 cm-1 may
have weaker overtone near 1200 cm-1, 1800 cm-1 and 2400 cm-1.

Combination bands is a type of overtones that occurs when two vibrational


frequencies (𝜈1 & 𝜈2) in a molecule couple to give rise to a new vibrational
frequency within the molecule. The energy of a Combination band is the sum of
the two independent absorption i.e. 𝜈𝑐𝑜𝑚𝑏 = 𝜈1 + 𝜈2 .

 D. L. Pavia, G. M. Lampman, G. S. Kriz and J. A. Vyuvan, Introduction to Spectroscopy, 5 th ed. Cengage


Learning, Stamford, USA, 2015.
 Paula Y. B. Organic chemistry, 8th ed. PEARSON, California, USA, 2016.
Generally, overtone bands are weak but they are important for the characterization
of some classes of organic molecules (benzene derivatives). For example, saturated
aldehyde and benzaldehyde can be differentiate by the above two band i.e., for
saturated aldehyde there is no weak overtone or Combination bands band near
2000 cm-1 while benzaldehyde have.

Difference bands and Fermi resonance

Like combination bands Difference bands are also another type of bands where
the observed frequency results from the difference between the two interacting
bands i.e.
𝜈𝑑𝑖𝑓𝑓 = 𝜈1 ─ 𝜈2 .

Now the Fermi resonance, which was discovered by scientist E. Fermi. He


noticed that, a molecule can oscillate independently only when it has different
frequencies. But in some cases these frequency are similar and they can exchange
their energy with one another and this situation is called resonance. More clearly
we can define Fermi resonance as coupled vibration that arises when a fundamental
vibration couples with an overtone or combination bands. It is observed in some
carbonyl compounds.

In Fermi resonance there is a sharing of


intensity so here the overtone is seen as
strong band. For benzoyl chloride
(C6H5COCl, we get two bands in the
C=O region at 1790 and 1745 cm-1. The
bands at 1790 cm-1 is for C=O stretching
and at 1745 cm-1 is for Fermi resonance
(overtone of CH bend coupled with C=O
fundamental) Fig-12: The IR spectrum (partly region)
of benzoyl chloride (KBr plates).

 C. N Banwell, Fundamentals of Molecular Spectroscopy, 3rd ed. McGraw-Hill, England, 1994.


 D. L. Pavia, G. M. Lampman, G. S. Kriz and J. A. Vyuvan, Introduction to Spectroscopy, 5 th ed.
Cengage Learning, Stamford, USA, 2015.
Now, at a glance, we can summarize the basic principle of Infrared spectroscopy
as follows:

The lower frequency IR radiation does not have enough energy to induce electronic
transitions (seen in with UV-Vis) but is sufficiently energetic to cause bonds to
deform. Thus absorption of IR is confined to compounds with certain vibrational
and rotational states.
When energy in the form of IR radiation is applied to a molecule, it induces
vibration (stretching or bending) between the atoms within the molecule. As the
molecule vibrates, there is a fluctuations in the dipole moment of the molecule.
Then this situation causes a field that interact with the alternating electric field of
IR radiation. Those frequencies of IR radiation that match the natural vibrational
frequencies of the molecule, are absorbed. Hence, the amplitude of the vibrational
motion of the bonds within the molecule are increased. At this point it should be
remembered that all bonds in a molecule are not capable of absorbing infrared
energy, even though the frequency of the IR radiation exactly matches that of the
bond motion. Only those bonds that are associated with change in a dipole moment
are capable of absorbing infrared radiation.
Thus increasing the rate of vibration leads to the intense absorption of IR radiation
and vice versa. Finally, it give rises to closely packed absorption band called IR
absorption spectrum.

View publication stats

You might also like