Вы находитесь на странице: 1из 20

31 Molecular Universe, HS 2009, D.

Fluri, ETH Zurich

3 Molecular Spectroscopy
Already the first studies on spectra revealed that, in addition to the continuous and line spectra
typical for atoms, there are spectra with an entirely different structure: in low spectral
resolution they appear not as single, sharp lines but more or less as broad wavelength features
(bands). We therefore speak of band spectra. The sources of band spectra are the molecules.
High resolution spectroscopy has shown that bands actually result from a very large number
of strongly overlapping individual molecular lines.

Figure 3.1: Bands of the molecules C2 and CN in low spectral resolution. Band positions are specified by
vertical lines above and below the spectrum. Note that every band actually consists of many individual, partly
overlapping spectral lines. Bands connected to the same horizontal line belong to one band system. From
Herzberg (1950).

The goal of this chapter is to understand the formation of molecular spectra with the
typical band structure. Fundamentally, there is no principle difference in the description of
atomic or molecular states and transitions. It is the same quantum mechanical principles that
apply to both atoms and molecules. However, molecules exhibit additional modes of motion,
namely rotations and vibrations, not present in atoms. This gives rise to additional quantum
numbers, which makes the structure of molecular states much more complex.
We restrict ourselves to the description of diatomic molecules, i.e. molecules consisting
only of two atoms. This is sufficient to understand the principles governing the structure of
molecules and the features of molecular spectra. Polyatomic molecules have additional
degrees of freedom for vibrational and rotational motions, which make their structure even
more complex but not fundamentally different.

3.1 Overview of molecular spectra

3.1.1 Born-Oppenheimer approximation

A molecule is an assembly of positively charged nuclei and negatively charged electrons that
form a stable entity through the electrostatic forces which hold it all together. Since all the
particles which make up the molecule are moving relative to each other, a full quantum
3 Molecular Spectroscopy 32

mechanical description of the molecule is very complicated and can only be obtained
approximately. Fortunately, the overall motion of the molecule can be broken down into
various types of motion, namely, translational, rotational, vibrational, and electronic. To a
good approximation, the so called Born-Oppenheimer approximation, each of these motions
can be considered on its own (Born & Oppenheimer 1927).
Their main objective was the separation of electronic and nuclear motions in a molecule.
The physical basis of this separation is quite simple. Both electrons and nuclei experience
similar forces in a molecular system, since they arise from a mutual electrostatic interaction.
However, the mass of the electron, me, is about four orders of magnitude smaller than the
mass of the nucleus, Mn. Consequently, the electrons are accelerated at a much greater rate
and move much more quickly than the nuclei. Therefore, as an approximation, we can regard
the dynamics of the electrons and nuclei as largely independent. When describing the
electrons, the nuclei can be considered as being fixed in space. On longer timescales, the
electrons can immediately follow the much slower nuclear motions.
To a very good approximation, the total energy of a molecule is given as the sum of three
components, corresponding to the three modes of motion mentioned above, namely

E = Ee + Evib + Erot , (3.1)

where the three terms on the right hand side represent the electronic, vibrational, and
rotational energy. We disregard overall translational energy of the molecule because it leads
mainly to collisions and, thus, has only an indirect influence on the internal structure of
molecules. Furthermore, we have neglected in Eq. (3.1) the interaction between the three
different types of motion. These interaction terms have to be considered in a consistent
treatment, and we will come back to them in the course of this chapter, but it is more
illustrative to neglect them as a first approximation.
In molecular spectroscopy, Eq. (3.1) is often expressed in units of the wavenumber
= /c , using the notation

T = Te + G + F , (3.2)

with

Te = Ee ( hc ) ,
G = Evib ( hc ) , (3.3)
F = Erot ( hc ) .

The energy difference between two states can then be denoted by

= Te + G + F , (3.4)

where the units of are cm1.


Since electrons move more rapidly than the nuclei, and the vibrations of nuclei are more
rapid than the rotation of molecules, the following relation holds:
33 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

Ee Evib Erot . (3.5)

We have already argued that the electronic energy is largest. The relative energies of
vibrations and rotations follow from simple order of magnitude estimates.
The valence electrons spread over the whole molecule (of size a 1 ) and thus have
typical energies

2
Ee 1 10 eV . (3.6)
me a 2

In this estimate we have used the Rydberg energy but with a instead of the Bohr radius. The
energy scale of vibrations about the equilibrium separation req of two nuclei in a molecule is
defined by . To find we consider the potential energy of the vibrational modes given
by Mn2(r req)2, cf. Eqs. (3.33) and (3.34). We have req ~ a. For vibrations with (r req) ~ a
the electron configuration would be modified strongly and costs roughly one electronic
energy Ee. From Mn2(r req)2 ~ 2 me a 2 we find

me 2
Evib 0.1 eV . (3.7)
M n me a 2

Rotational energies are estimated by J2/I (cf. Eq. (3.13)) where J is the angular momentum
and I is the moment of inertia. With J2 ~ 2 j ( j +1) 2 and I ~ Mna2 we get

J 2 me 2
Erot 1 meV . (3.8)
I M n me a 2

The corresponding length scales are e ~ 1001000 nm (i.e. in the order of visible
wavelengths), vib ~ 10 m (i.e. in the infrared region), and rot ~ 1 mm. Therefore, spectral
lines of molecules are observed in the visible (actually near UV to about 1 m) in the case of
electronic transitions, in the infrared for vibrational transitions, and in the sub-mm to about 1
mm region (microwaves) for rotational transitions.

3.1.2 Types of molecular spectra

As a result of Eq. (3.5) and our order of magnitude estimates of typical energy scales, we can
distinguish three types of spectra (Figure 3.2):

Rotational spectra are transitions between the rotational levels of a given vibrational
level in a particular electronic state. Only the rotational quantum number J changes in
these transitions. These spectra lie in the region of microwaves or in the far infrared and
are treated in Sect. 3.2. They consist typically of a large number of closely spaced, nearly
equidistant spectral lines.
3 Molecular Spectroscopy 34

Rotational-vibrational spectra consist


of transitions from the rotational levels
of a particular vibrational level to the
rotational levels of another vibrational
level in the same electronic state. The
electronic excitation state thus remains
the same. The quantum numbers J and
change, where characterizes the
quantized vibrational levels. These
spectra lie in the infrared spectral region
and are treated in Sect. 3.3. They consist
of a number of bands, i.e. groups of
closely-spaced lines.
Electronic spectra consist of transitions
between the rotational levels of the
various vibrational levels of one
electronic state and the rotational and
vibrational levels of a different electronic
state. The corresponding spectrum forms
a so called band system. It contains all
the vibrational bands of the electronic Figure 3.2: Vibrational levels (quantization number
transition being observed, each one with ) and rotational levels (quantum numbers J) of two
its rotational structure. In general, all electronic states of a molecule, denoted by I and II.
The three arrows refer to (from left to right) transi-
quantum numbers change in these
tions in the rotational, the rotational-vibrational, and
transitions, i.e. J, , and those describing in the electronic spectrum of the molecule. From
the electronic state. These spectra lie in Haken & Wolf (2006).
the near infrared, the visible, or the near
ultraviolet regions. Electronic spectra are discussed in Sect. 3.4. The band systems of all
the allowed electronic transitions of a molecule together make up the band spectrum
proper of the molecule.

Figure 3.3: An overview of the spectral positions of the absorption spectra of a small molecule. The numerical
values are approximately correct for HCl. From Haken & Wolf (2006).

3.2 Rotational Spectra


We first introduce the simple model of a rigid rotor to show the main characteristics of the
rotational spectrum. Subsequently, the model is extended to a non-rigid rotor to take care of
some necessary corrections.
35 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

3.2.1 Rigid Rotor

The simplest model of a rotating diatomic


molecule is the so called dumbbell model
(Figure 3.4). Wee assume that the molecule
consists of two atoms with masses
masse m1 and m2
connected by a rigid massless rod of length r. Figure 3.4: Rigid rotor or model (dumbbell),
illustrating the rotation of a diatomic
dia molecule
That is, we neglect the finite extent of the atoms
about its center-of-gravity.
gravity. The rotation axis is
and any possible modification of the distance perpendicular to the internuclear axis.
between the two atoms resulting from the
rotation.
In classical mechanics the energy of rotation Erot of such a body is given by

1 2
Erot = I , (3.9)
2

where is the angular velocity of the rotation (about the center of mass and perpendicular to
), related to the frequency of rotation rot = 2rot. I is the moment of
the internuclear axis),
inertia, defined by

I = m1r12 + m2 r2 2 , (3.10)

where r1 and r2, the distances to the masses m1 and m2 from the center-of-gravity,
gravity, are given by

m2 m1
r1 = r, r2 = r. (3.11)
m1 + m2 m1 + m2

Substituting for r1 and r2 in Eq. (3.10) gives an important result for the moment of inertia

m1m2 2
I= r = r 2 . (3.12)
m1 + m2

Here is called the reduced mass of the molecule, and the above analysis shows that this
simple system is actually equivalent to the rotation of a single point of mass at a fixed
distance r from the axis of rotation. Such a system is called a rigid rotor (or rotator). We can
now rewrite Eq. (3.9) in terms of the angular momentum vector J:

J2 J2
Erot = = . (3.13)
2 I 2 r 2

This result leads us naturally to a quantum mechanical description of the system.


system The
dinger equation for a single point of mass is
Schrdinger

2 2 2 2
+ + = E (3.14)
2 x 2 y 2 z 2
3 Molecular Spectroscopy 36

In polar coordinates the Hamiltonian is

2 2 2 2
H = r sin sin + . (3.15)
2 r 2 r r 2 r 2 sin 2 2

Two remarks should be made about this expression. First, it has the same form that appears
also when calculating the hydrogen atom. Second, for the rigid rotor the first term in Eq.
(3.15) disappears and the Schrdinger equation becomes

2 1 1 2
sin sin + ( , ) = E ( , ) , (3.16)
2 r 2 sin 2 2

where the operator in the brackets is just the square of the (dimensionless) angular momentum
operator J in polar coordinates, so that the equation may be rewritten in the form

2
J 2 ( , ) = E ( , ) . (3.17)
2 r 2

We have already shown above that r2 is the moment of inertia I. The eigenfunctions are
spherical harmonics and the eigenvalues are given by

2
E= J ( J + 1) . (3.18)
2I

In wavenumber units (see Eqs. (3.2) and (3.3)) the energy levels of the rigid rotor are

F ( J ) = BJ ( J + 1) , (3.19)

with the rotational constant B (in cm1) defined by

2
B= . (3.20)
2 Ihc

The rotational quantum number J takes integral values, J = 0, 1, 2, . If J = 0,


corresponding to the non-rotating molecule, then F(J) = 0. In a rigid rotor the energy of the
levels increases as J2. Every level is (2J + 1)-fold degenerate.
The selection rules for electric dipole transitions of the rigid rotor obey the relations

J = 1 ,
(3.21)
m = 0, 1 .

These correspond to the rigorous selection rules derived in Sect. 2.3.5 except that J = 0
cannot appear since we are dealing effectively with a single particle system. Thus electric
dipole transitions occur between adjacent levels.
37 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

Furthermore, in the case of a real diatomic


molecule, electric dipole transitions within a
fixed electronic state are only allowed if it
possesses a permanent electric dipole moment.
We come back to this point later as it is crucial
for the most abundant molecule in the universe,
H2, which has no permanent dipole moment
since it is homonuclear.
According to Eq. (3.19), the frequencies of
lines in the rotational spectrum (in wavenumber
units) scale linearly in J

= F ( J + 1) F ( J ) = 2 B( J + 1) . (3.22)

Therefore, the spectrum of the rigid rotor con-


sists of a series of equidistant lines with a sepa-
ration 2B (Figure 3.5 and Figure 3.6). Using
observed rotational spectra the rotational con-
stant B can easily be determined empirically
from Eq. (3.22). In general, B depends not only Figure 3.5: Energy levels for the rotation of a
on the specific molecule but also on the vibra- rigid diatomic molecule. The selection rule J =
tional and electronic state of the molecule. 1 yields a spectrum consisting of equidistant
lines as illustrated at the bottom. From Haken &
We can now estimate more explicitly the
Wolf (2006).
wavelength of lines in the rotational spectrum of
molecules by employing Eqs. (3.12), (3.20) and
(3.22). We consider the HCl molecule as a typical example and set J = 0 in Eq. (3.22). Insert-
ing the masses of H and Cl, and their internuclear distance, known from gas-kinetic mea-
surements to have the value r = 1.28 108 cm, we find a line frequency = 6.3 1011 Hz or
= 0.48 mm. This estimate confirms the statement from Sect. 3.1.2 that rotational spectra lie
in the far-infrared to microwave region.
Rotation of the diatomic molecule about the internuclear axis can be neglected because
the corresponding moment of inertia is very small. This is not anymore the case for multi-
atomic molecules, though.

Figure 3.6: Rotational spectrum of HCl in absorption. From Haken & Wolf (2006).
3 Molecular Spectroscopy 38

3.2.2 Non-Rigid Rotor

When rotational spectra are analyzed with high


precision, it becomes apparent that the lines
lin are not
exactly equidistant. Instead, their spacing becomes
smaller as the quantum number J is increased. To
understand this, we have to employ a better model
for describing the rotations of a molecule,
molecule which
brings us to the non-rigid
rigid rotor (Figure
( 3.7).
As a result of centrifugal forces when the
molecule rotates, the internuclear distance and hence
the moment of inertia will increase with increasing Figure 3.7: Non-rigid
rigid rotor model of a
rotation. In the rotating molecule the internuclear
i diatomic molecule. The two atoms are
distance r is determined by the requirement that the connected with a spring of force constant k
centrifugal force Fc is balanced by the restoring that can be stretched upon rotation. In
force k(r req), k being the force constant of the principle, this model allows also for
simultaneous rotation and vibrations (along
chemical bond (whenwhen viewed as a spring) and req the axis of the spring) but
b we will consider
being the equilibrium internuclear distance at rest this combination only in Sect. 3.3.3.
(i.e. without rotation).
Classically, we can calculate the new internuclear distance r and thus the stretching of the
molecule by

r 2 = k ( r req ) . (3.23)

Since the absolute value of the angular momentum is given by J = I = r2 we find

r 2 J2 J2
r req = = . (3.24)
k r 3 k req3 k

The total rotational energy is now the sum of the kinetic and potential energies, given by

J2 1 J2
+ k ( r req ) = + k ( r req ) .
2 1 2
Erot = (3.25)
2I 2 2 r 2
2

If we expand the denominator in the first term of Eq. (3.25), then substitute r req using Eq.
(3.24),, and neglect cubic and higher powers of r req, we obtain

J2 J4
Erot = + ... . (3.26)
2 req2 2 2 req6 k

quantum mechanics and replace J2


If we now make the transition from classical mechanics to quantum
by 2J(J + 1), as usual, we get the rotational energy of the non-rigid
rigid rotor as a power series
(neglecting higher order terms though)
39 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

2 4
Erot = ( + ) J 2 ( J + 1) ,
2
J J 1 (3.27)
2 req
2
2 req k
2 6

or, expressed in wavenumber units,

F ( J ) = BJ ( J + 1) DJ 2 ( J + 1)
2
. (3.28)

Compared to the rigid rotor, we get


corrections for the centrifugal distortion
expressed by the parameter D (and in
principle also higher order terms), in
addition to the identical first term. The
rotational constants B and D (in cm1) are
given by

2
B= ,
2 req2 hc
(3.29)
4
D= .
2 2 req6 khc

The term scheme of the non-rigid


non rotor
can be found from that of the rigid rotor by
shifting the levels and lines as illustrated in
Figure 3.8. The frequencies of the lines in
the rotational spectrum (in in wavenumber Figure 3.8: The energy levels and the spectrum of a
units) may be found from Eq. (3.28) by non-rigid
rigid rotor compared to a rigid rotor. A value D =
applying the selection rule J = 1 for 103B has been assumed for the rotational constant D.
The levels of the rigid rotor, which are equidistant
electric dipole transitions:
with a spacing of 2B,, are shifted towards lower
energies in the non-rigid
rigid rotor, and the shift increases
= F ( J + 1) F ( J ) with increasing J.. This effect is exaggerated in the
(3.30)
= 2 B ( J + 1) 4 D ( J + 1)3 . figure. From Haken & Wolf (2006).

3.3 Vibrational Spectra


The next stage in our development of the internal dynamics of a diatomic molecule is to
consider vibrations. We employ the model of the non-rigid
non rigid dumbbell where the two atoms are
connected with a spring (Figure
Figure 3.9). For instructive
tive purposes, we assume first in Sects. 3.3.1
and 3.3.2 that the moleculele is non-rotating.
non rotating. Once we understand the main features of such a
vibrating, but non-rotating,
rotating, molecule, we add also rotational movements and discuss the
realistic model of a vibrating rotor in Sect. 3.3.3. Since the vibrational energy is much larger
than the rotational energy the combined treatment of the two modes modes of motion will be
straightforward.
3 Molecular Spectroscopy 310

3.3.1 The Harmonic Oscillator

The simplest model of a vibrating molecule


is the harmonic motion of atoms. As above
we can represent this motion by the
harmonic oscillation of a single particle with
reduced mass as defined by Eq. (3.12).
The force acting on the mass point is

d2 x Figure 3.9: Model of a vibrating diatomic molecule.


F = = kx , (3.31) The two atoms are connected by a massless spring.
dt 2

where x = r req is the distance from the equilibrium position. The solution to this differential
equation is

x = x0 sin(2 osct + ) , (3.32)

where the vibrational frequency is given by

1 k
osc = , (3.33)
2

and x0 is the amplitude of the vibration. The force F is the negative derivative of the potential
energy V. Therefore, it follows that

1 2
V= kx . (3.34)
2

The total vibrational energy is given by

p 2 kx 2
Evib = + . (3.35)
2 2

We can now convert the problem from classical mechanics to quantum mechanics by
employing the correspondence principle, and replace x x and p i(/x) with operators.
The Schrdinger equation for the harmonic oscillator becomes

2 d 2 1 2
+ kx = E . (3.36)
2 dx 2 2

The resulting energy levels are well known from basic quantum mechanics

1
E = h + , = 0, 1, 2, , (3.37)
2
311 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

where is the vibrational quantum number and


is the harmonic vibrational frequency given by
Eq. (3.33). The levels are equally spaced as
illustrated in Figure 3.10. In wavenumber units
the vibrational energy is written as

E 1
G ( ) = = e + , (3.38)
hc 2

which represents (a first approximation of) the


second term in Eq. (3.2). The vibrational
constant e = /c is the vibrational frequency
measured in cm1. The actual number of Figure 3.10: Energy levels of a harmonic
vibrations is c times as great. It is important to oscillator. The levels are labeled with the
note that for = 0, the lowest vibrational level, vibrational quantum number . The selection
there remains a zero-point energy. rule = 1 is strictly valid only for the
harmonic oscillator and does not apply anymore
In principle, electric dipole transitions obey
for the anharmonic oscillator.
the selection rule = 1. Since the energy
levels are equidistant, the spectrum of a harmonic oscillator consists of a single line.
However, a real molecule is better described by an anharmonic oscillator, for which no
selection rules apply. Nonetheless, vibrational transitions within a given electronic state
remain strongest for = 1.
We can again readily estimate that molecular vibrations lie in the infrared spectral region,
which we show by a simple estimate for the HCl molecule. We consider the restoring force of
a spring F = k(r req). Assuming a pure Coulomb force, the force constant k (in cgs units)
can be calculated by

dF 2e 2
k= = 3 . (3.39)
dr req

From Eqs. (3.33) and (3.39), we find, when inserting the masses of H and Cl and their
internuclear distance r = 1.28 108 cm, an estimated line frequency 6 1013 Hz or a
wavelength 5 m. The measured wavelength of the = 1 transition of HCl is 3.5
m, which is of the same order of magnitude as our estimate from the greatly simplified
model.

3.3.2 The Anharmonic Oscillator

The potential energy of the harmonic oscillator is represented by a parabola as indicated by


Eq. (3.34). Therefore, the potential energy and the restoring force increase indefinitely with
increasing distance from the equilibrium position. However, this cannot be a correct
description for a real molecule because when the atoms are an infinite distance apart, the
attractive force between them must be zero, and the potential energy then has a constant
value, corresponding to the total energy of the separated atoms.
3 Molecular Spectroscopy 312

Figure 3.11: The Morse potential curve for the HCl molecule. A harmonic potential (dashed) is shown for
comparison. The dissociation energy from the potential minimum is called De. From Haken & Wolf (2006).

If we take into account the interaction energy between vibrational and electronic motions,
we obtain a better model of the molecule: the anharmonic oscillator. This is achieved by
employing a more realistic potential V. An often-used approach which agrees well with
experience is the so called Morse potential (Figure 3.11)

a ( r re )
2
V = De 1 e , (3.40)

where De is the dissociation energy and a characterizes the molecule under consideration.
Apart from its simplicity, an advantage of the Morse potential is that it enables the
Schrdinger equation to be solved rigorously.
In the general case it is found that the term values of the vibrational levels of the
anharmonic oscillator are given by

2 3
1 1 1
G ( ) = e + e xe + + e ye + + , (3.41)
2 2 2

although for the Morse potential the series in Eq. (3.41) does not extend higher than the
quadratic term. As a consequence of the quadratic and higher terms the vibrational spacing
between levels decreases as increases, with the levels essentially converging to the
dissociation asymptote (Figure 3.12). The vibrational equilibrium constants e, exe, and eye
depend on the potential curve of the electronic state, i.e. on the electronic configuration of a
given molecule.
From the energy terms given by Eq. (3.41) we can derive the spectrum of an anharmonic
oscillator by applying the selection rules (Figure 3.13). The electric dipole selection rule =
1 for the harmonic oscillator must be modified in the case of the anharmonic oscillator. In
fact, there are no principle restrictions for anymore, i.e. we have

= 1, 2, 3, , (3.42)
313 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

Figure 3.12: The energy levels of an anharmonic oscillator. The three arrows correspond to the fundamental
frequency and the first two harmonics in the vibrational spectrum. One can readily recognize the increase of the
average internuclear distance with increasing vibrational quantum number . From Haken & Wolf (2006).

Figure 3.13: Coarse structure of the infrared spectrum of HCl (schematic). The vertical lines indicate transitions
between the = 0 level and excited vibrational levels within the electronic ground state of the HCl molecule.
The labels indicate the vibrational quantum number of the upper level. The intensity actually falls off five
times faster than indicated by the height of the vertical lines. From Herzberg (1950).

but the relative intensities decrease rapidly. The transition = 1 0 is called the
fundamental transition, while those with greater differences in are known as overtone
transitions or simply as harmonics.
Note that the energy differences of transitions with = 1 are no longer the same for all
values of in the case of the anharmonic oscillator. Instead they decrease with increasing .

3.3.3 Vibrating Rotor

In real molecules simultaneous rotation and vibration occurs. The schematic vibrational
spectrum shown in Figure 3.13 represents only a reasonable first approximation at very low
spectral resolution, as is apparent from a measured CO spectrum (Figure 3.14). Because of the
3 Molecular Spectroscopy 314

rotation every line of the anharmonic


oscillator actually exhibits a rotational
substructure consisting of many individual
lines as revealed with high spectral
resolution (Figure 3.15). This spectrum
arises from the fact that vibrating molecules
also rotate, and that vibration and rotation
are coupled. Thus, such a spectrum is called
rotational-vibrational spectrum.
Therefore, to reach a realistic de-
scription of simultaneous molecular vi- Figure 3.14: The vibrational spectrum of CO in the
bration and rotation we have to improve our gas phase obtained with poor spectral resolution. The
modeling once more by introducing the fundamental vibration is at 2143 cm1 and the first
vibrating rotor. Neglecting the interaction harmonic at 4260 cm1. From Haken & Wolf (2006).
between vibration and rotation, the energy
of the vibrating rotor is simply the sum of the vibrational energy of the anharmonic oscillator
given by Eq. (3.41) and the rotational energy of the non-rigid rotor see Eq. (3.28).
Therefore, a series of rotational levels of the type of Figure 3.8 would exist for each of the
vibrational levels in Figure 3.12. The resulting energy level diagram within a given electric
state is illustrated in Figure 3.16.

Figure 3.15: Rotational-vibrational spectrum of the CO molecule due to the fundamental vibrational transition,
measured at high spectral resolution. Left and right of the center (2143.28 cm1) are the P and R branches. From
Haken & Wolf (2006).

Interaction between Vibration and Rotation

In a more accurate treatment we must take into consideration the fact that during the vibration
the internuclear distance and consequently the moment of inertia and the rotational constants
are changing.
315 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

We have already seen that vibrations


of a molecule take place on a much
smaller time scale than its rotations.
During a single rotation, a molecule
vibrates several thousand times.
Therefore, it seems plausible to use
mean values of the rotational constants B
and D in the vibrational level considered,
averaged over many vibrations.
In the case of the anharmonic
oscillator, the average internuclear
distance increases with increasing
vibrational quantum number , i.e. with Figure 3.16: Energy levels within a given electronic state
resulting from the vibrating rotor model. For each of the
increasing vibrational excitation. The vibrational levels (labeled by the vibrational quantum
moment of inertia also increases and the number ) a number of rotational levels are drawn (short
rotational constant B becomes smaller. horizontal lines).
According to rather involved quantum
mechanical calculations to a first (usually satisfactory) approximation, the rotational constant
B in the vibrational state is given by

1
B = Be + . (3.43)
2

Here, Be means the rotational constant as given by Eq. (3.29), and is a molecule specific
positive constant with Be. Due to the zero-point energy of an oscillator there is a
correction to Be even in the = 0 state.
In a similar manner, a mean rotational constant D, representing the stretching of the
molecule by the centrifugal force, must be used. For the anharmonic oscillator D depends on
the vibrational state

1
D = De + , (3.44)
2

where De corresponds to the vibrationless case as expressed by Eq. (3.29), and represents a
correction factor De.
Taking into account the anharmonicity and the interaction between vibration and rotation
we obtain the energy levels of the vibrating rotor (neglecting terms of higher order)

2
1 1
= G ( ) + F ( J ) = e + e xe + + B J ( J + 1) D J 2 ( J + 1) .
2
T ,J (3.45)
2 2

Rotational-Vibrational Spectra

Since the eigenfunctions of the vibrating rotor are essentially products of oscillator and rotator
eigenfunctions, it is readily seen that the same selection rules for electric dipole transitions
3 Molecular Spectroscopy 316

hold as for these systems individually. That is, can change by any integral amount although
= 1 gives by far the most intense transitions, and J can only change by unity, i.e. J = 1.
Of course, = 0 is also allowed. However, this does not give rise to a rotational-vibrational
spectrum but rather to the pure rotational spectrum discussed in Sect. 3.2.
Let us now consider a particular vibrational transition from to . Primed quantities
refer to the upper level and double-primed quantities to the lower level. According to Eq.
(3.45) the wavenumbers of the resulting lines (neglecting the rotational constant D)

= 0 + B J ( J + 1) BJ ( J + 1) , (3.46)

where 0 = G( ) G( ) is the wavenumber of the pure vibrational transition without taking


account of rotation (J = J = 0). With J = J J = +1 and J = J J = 1,
respectively, we obtain from Eq. (3.46)

R = 0 + 2 B + ( 3B B ) J + ( B B ) J 2 ; J = 0, 1, 2, , (3.47)

P = 0 ( B + B ) J + ( B B ) J 2 ; J = 1, 2, 3, . (3.48)

Here, as is common in molecular spectroscopy, J has been replaced by J. Since J can take a
whole series of values, these formulae represent two series of lines, which are called the R
and P branches, respectively, defined by

R branch: J = J J = +1 ,
(3.49)
P branch: J = J J = 1 .

Since the lowest value of the rotational quantum number (J) in the upper and lower levels is
zero it is clear from Figure 3.17 that the smallest value of J ( J) in the R branch is 0 while
in the P branch it is 1. The resulting spectrum is indicated schematically at the bottom of
Figure 3.17. It is qualitatively in complete agreement with the observed rotational-vibrational
spectrum shown in Figure 3.15.
If we neglect, for a moment, the interaction between rotation and vibration, we have B =
B = B and Eqs. (3.47) and (3.48) simplify to

R =0 + 2 B + 2BJ , P =0 2 BJ , (3.50)

that is we have two series of equidistant lines. The R branch, given by R, extends from 0
toward shorter wavelengths and the P branch, given by P, is going toward longer
wavelengths. Owing to the above restrictions for the J values, there is no line at 0, i.e. we
have a zero gap. The spectrum calculated with according to Eq. (3.50) is shown in strip (b) at
the bottom of Figure 3.17. The observed slight convergence of the lines is to be traced back to
the influence of the interaction of rotation and vibration. In consequence of this interaction, B
differs from B , giving rise to the quadratic terms in Eqs. (3.47) and (3.48). This has the
effect that, when B < B , the lines in the R branch draw closer together and those in the P
branch draw farther apart.
317 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

Figure 3.17: Energy level diagram explaining the rotational structure of a rotational-vibrational
rotational vibrational band. In general,
the separation of the two vibrational levels is considerably larger compared to the spacing of the rotational levels
than shown in the figure. Every transition
nsition is labeled by the branch designation (R or P) and, in parentheses, by
the rotational quantum number J of the lower level. The schematic spectra (a) and (b) give the resulting spectra
with and without allowance for the interaction between rotation and and vibration. Note that in these spectra, short
wavelengths are at the left. From Herzberg (1950).

As may easily be verified, the two branches (3.47) and (3.48) can also be represented by a
single formula, namely

= 0 + ( B + B ) m + ( B B ) m2 , (3.51)

where m is an integral number which takes values 1, 2, 3, for the R branch (that is m = J +
1) and the values 1, 2, 3,
3, for the P branch (that is m = J)) (dont confuse this m with the
3 Molecular Spectroscopy 318

quantum number of the Zeeman sublevels!). Thus we can also say that we have a single series
of lines for which a line is missing at m = 0. The missing line, at = 0, is called the zero
line. It would correspond to the forbidden transition between the two rotationless states J = 0.
0 is also called band origin.
For sufficiently high J a reversal of the R branch occurs, corresponding to the vertex of
the parabola represented by Eq. (3.51). This reversal (band head formation) will be discussed
at greater length for electronic band spectra, since it is much more often observed there.
For completeness it should be noted here that there are cases when the zero line is
observed in rotational-vibrational spectra, or more precisely when a whole additional branch,
the so called Q branch (defined by J = 0) appears. The Q branch lines are much more closely
spaced and usually overlap so strongly that they appear as a single intense line at the band
origin. In diatomic molecules the Q branch is obtained within the quantum mechanical model
when considering the (tiny) moment of inertia about the internuclear axis and even then it is
only allowed if the angular momentum quantum number of the electrons > 0 (cf. Sect.
3.4.2). For many diatomic molecules, = 0 in the electronic ground state, so that the Q
branch is forbidden. Thus, our above treatment is satisfactory and typical for most cases of
diatomic molecules. In polyatomic molecules the Q branch is more common for some types
of vibrational modes, e.g. the bending mode of CO2, since in general all the moments of
inertia about all three principle axes are non-vanishing.

Intensity Variation in Rotational-Vibrational Bands

Within a rotational-vibrational band the intensities of the individual lines vary with a typical
distribution as shown in Figure 3.15. The intensities of lines scale with the level populations
and the transition strengths. In the case of rotational-vibrational bands the main characteristics
of line intensities can be obtained by the thermal occupation probabilities of the rotational
levels taking the degeneracy of the levels into account.
Assuming thermodynamic equilibrium the relative population of excited states at a given
temperature T is determined by the Boltzmann factor eE/kT, where E is the excitation energy.
In addition, we have to take into account that every rotational level is (2J + 1)-fold degenerate
and actually consists of (2J + 1) states. Therefore, the number of molecules NJ in the
rotational level J of the lowest vibrational state is proportional to (2J + 1) eF(J)hc/kT. For many
practical cases ( = 0, rigid rotor)

BJ ( J +1) hc

N J ( 2 J + 1) e kT
. (3.52)

This function is represented in Figure 3.19. Since the factor (2J + 1) increases linearly with J,
the number of molecules in the different rotational levels does not immediately decrease with
the rotational quantum number but first goes through a maximum lying at

kT 1
J max = , (3.53)
2 Bhc 2
319 Molecular Universe, HS 2009, D. Fluri, ETH Zurich

Figure 3.18: Intensity distribution in rotational-vibrational bands in absorption at different temperatures. (a) For
B = 10.44 cm1 (HCl). (b) B = 2 cm1. The y-axis, which is not explicitly given, is the same for all diagrams. The
lines are drawn with the separation that they would have if the constant B were the same in the upper and lower
levels. m is the running number of the lines (cf. Eq. (3.51)). Note that longer wavelengths are to the right in this
figure. From Herzberg (1950).

that is, at a value of J that increases


with decreasing B and increasing T
(Figure 3.18). It should be noted that
the number of molecules in the lowest
rotational level J = 0 is not zero. If
thermodynamic equilibrium applies,
the observation of a rotational-vibra-
tional band together with Eq. (3.53)
allows us to diagnose approximately
the temperature.
The position of the most intense
transition is given only approximately
by Eq. (3.53) because the intensity
distribution depends not only on the Figure 3.19: Thermal distribution of the rotational levels
occupation probabilities but also on the for T = 300 K and B = 10.44 cm1, i.e. HCl in the ground
state. The curve represents Eq. (3.52) as a function of J.
transition moments. A proper quantum The broken lines ordinates give the relative populations of
mechanical treatment yields the line the corresponding rotational level. From Herzberg (1950).
intensities within a rotational-vibra-
tional band

B J ( J +1) hc B J ( J +1) hc

I em S J e kT
, I abs S J e kT
, (3.54)

for emission and absorption lines, respectively. In the case of the rotor ( = 0) the line
strength SJ is given by SJ = J + 1 for the R branch (J = +1) and by SJ = J for the P branch (J
= 1). If 0 other formulae hold which will be discussed in Sect. 3.4.5 (Hnl-London
factors).
3 Molecular Spectroscopy 320

Need for a Permanent Electric Dipole Moment

Allowed transitions (i.e. electric dipole transitions) in a rotational or a rotational-vibrational


spectrum only occur if the molecule possesses a permanent electric dipole moment. This is in
contrast to electronic spectra involving a change in the electronic state, which can be allowed
in the absence of a permanent electric dipole moment. The latter situation is familiar from
atomic spectroscopy where we are only dealing with electronic transitions and from where an
allowed electronic transition requires a change in the electron configuration. Both rotational
and a rotational-vibrational spectra arise, however, from transitions within a fixed electronic
state, i.e. the electron configuration remains constant, leading to the requirement of a
permanent electric dipole moment.
According to classical electrodynamics, an intramolecular motion leads to radiation of
light only if a changing electric dipole moment is associated with it. Similarly, in quantum
mechanics, the transition amplitudes for pure rotational or rotational-vibrational allowed
transitions are non-vanishing only in presence of a permanent electric dipole moment.
This has serious consequences for astrophysics since H2, the most abundant molecule in
the universe and being a homonuclear molecule, has no permanent electric dipole moment in
the ground state. Rotational and rotational-vibrational spectra of H2 thus arise only due to
higher order multipole transitions, which are usually too weak to be observable.

3.4 Electronic Spectra


In Sect. 3.1.1 we described the Born-Oppenheimer approximation which allows us to separate
the electronic and nuclear motion. In the previous two sections we have already discussed the
nuclear motion in terms of rotations and vibrations. Here, in subsections 3.4.1 and 3.4.2, we
first consider the motion of the electrons in the field produced by fixed nuclear charges and
look at the resulting electronic states. Then, in Subsects. 3.4.3 and 3.4.4, we combine the
effects of rotational, vibrational, and electronic energies and discuss the features of electronic
spectra including the rotational and vibrational substructure.

3.4.1 Molecular Orbital Theory

The calculation of electronic states in molecules involves the solution of a many body
problem, already in simplest case of the H2+ molecule consisting of two protons and one
electron. Therefore, even after making the Born-Oppenheimer approximation, we must resort
to further approximations. There are a number of different approaches to the description of
molecular electronic states. We briefly discuss molecular orbital theory, which has been and
still is by far the most significant and popular approach to both the qualitative and quantitative
description of molecular electronic structure. We will restrict ourselves to a qualitative
treatment only.
Within molecular orbital theory, the most common and important approach to compute
the wavefunctions for a molecule is the linear combination of atomic orbitals (LCAO)
method. As the name implies, this method involves consideration of the interaction of the
atomic orbitals on the separated atoms to form molecular orbitals which accommodate the
available electrons. Thus, we can rely on and expand the atomic theory, from which we have

Вам также может понравиться