Вы находитесь на странице: 1из 26

Advanced Physical Practical Report William Hodds.

Constant Volume Heat Capacity, CV, m, of SO2


Derived from Vibrational-Rotational IR Spectra,
and Temperature Determination of CO Through
Statistical Thermodynamics
William Hodds
Abstract
In this report a statistical thermodynamics approach to calculating the constant volume molar heat
capacity is demonstrated for SO2 at 298 K, where summation of all excited degrees of freedom yields

the total constant volume molar heat capacity, ,


. , was calculated to be 31.37 1 1
using this approach. Degrees of freedom considered were vibrational, rotational and translational,
and were found to contribute 12.47 1 1, 12.47 1 1 and 6.43 1 1
respectively. These values were calculated using equipartition theory for the translational and
rotational degrees of freedom, where << kBT, and statistical thermodynamics for the vibrational
degrees of freedom, where the 3 fundamental vibrational modes, 1, 2, and 3 were considered to
contribute at 298 K. The electronic contribution to heat capacity was assumed to be zero.
Also, an attempt to calculate the temperature at which an IR spectrum of CO was recorded at was
made, however inaccurate values were obtained due to the large contributions of centrifugal
distortion and vibrational-rotational coupling at high and , to the rotational level transition
energy, and the rotational constant, , at which were relevant in the second overtone which
was studied.

Introduction
Spectroscopy and emission finds applications in a large number of academic and industrial uses,
including the detection of significant functional groups in a newly synthesised chemical by presence
of characteristic absorptions in the infra-red spectrum of the chemical, and even the probing of the
radiative emissions of deep space objects remotely, as has been done via mapping the emissions of
transient SO2 in the circumstellar envelope of VY Canis Majoris[1], one of the largest and most
luminous red hypergiant stars known[2]. By the detection of S-containing molecules in various
regions of red hypergiants like VY Canis Majoris, models of the flow of gases and escape of heavier
gases from the core can be tested[3]. In addition to this, information of what chemistry occurs within
the peripheral layers of stars can be gained from spectroscopic measurements in the IR range, for
example, the 3 mode of SO2 was detected at 7.3 in the spectra of the stars UX Cyg, o Cet and T
Cep[4], and has been speculated to have formed due to pulsation shocks, gravity induced
compressions within the star[5]. Among other techniques, rovibrational spectroscopy is the means of
which to monitor these mass-loss events remotely.
Rotational-vibrational spectroscopy is used to provide an insight into the occupation of energy states
of a sample of molecules, by measuring the emission or absorption of infra-red radiation (range
1

Advanced Physical Practical Report William Hodds.


14300 cm-1 to 500 cm-1), distinct vibrational transitions can be detected, co-occurring with a range
of rotational transitions, from the presence of bands in the observed spectra. Due to the
quantisation of energy levels, specific vibrational transitions occur at consistent and distinct
energies, of which, for the simple fundamental transitions (0 1 ) can even be predicted using ab
intio software such as Gaussian 09. The dominant model of considering vibrational states is that of
energy levels, where an absorption is an excitation of a molecule from a lower vibrational state to a
higher vibrational state, governed by selection rules, derived from solving various Schrdinger
equations, dependent on the type of description the molecule is treated with.
For a diatomic which is considered a harmonic oscillator (energy of each vibrational level is given by
1
2

= ( + ) , where is the wavenumber of the fundamental transition 0 1 and is the


vibrational quantum number) the specific selection rule for vibrational transitions is:
= 1, 2, 3

(1.1)

Another approximation to the diatomic molecule that can be made is that of the rigid diatomic rotor,
a molecule which experiences no centrifugal forces when rotating, and hence has a fixed bond
length. In such diatomics, the following specific selection rule is enforced on rotational transitions:
(1.2)

= 1, 2, 3

The harmonic oscillator and rigid rotor approximations are often made together to get an
approximate description of the vibrational and rotational energy levels in a diatomic, which can then
be carried onto predicting IR spectra, with aid of potential energy calculations. However, when
overtones (for example, 0 2 ) are observed, and high rotational energy levels are studied more
closely, the predictions of the approximate models begin to diverge from experimental fact. The
deviations observed are finite and are due to three phenomena, one anharmonicity, where the
1

energy of a given vibrational level is not given by ( + 2) but instead is given by:
1
1 2
= ( + ) ( + )
2
2

(1.3)

where is the anharmonicity constant for the mode, and is a few magnitudes smaller than . At
1 2

higher , anharmonicity becomes more significant (the ( + 2) term becomes larger, causing
the vibrational levels to converge, ultimately leading to = 0, which freely allows dissociation of
the molecule to occur.
The other factors not accounted for in the harmonic rigid rotor model are centrifugal distortion, and
rotational-vibrational coupling, and these are described in the results and discussion section to
follow.

It is also practically very useful to know the constant volume molar heat capacity of a gas, ,
, as in
engineering it is essential to be able to compensate for the possible heat-expansion of continuousflow gases within fixed-volume pipes and vessels. This is especially relevant in chemical plants,
where any splitting of pipes will lead to the expulsion of often toxic and environmentally-damaging
gases into the immediate environment/atmosphere.

Advanced Physical Practical Report William Hodds.

Results and Discussion


Fig 2.1. The IR spectrum of gaseous SO2 at a pressure of 12 mmHg (16.0 mbar) at 298 K, showing 3

(1 + 3)

prominent absorption bands at 1149 cm-1 (A =8.5 x 10-3), 1363 cm-1 (A = 6.7 x 10-2) and 2501 cm-1 (A =
1 x 10-4), attributed to 1, 3 and combinational mode (1 + 3), respectively.
In the above IR spectrum, the band at 1149 cm-1 is attributed to the fundamental 1 mode
(symmetric S-O stretch), the band at 1373 cm-1 to the fundamental 3 mode (asymmetric S-O stretch)
and the band at 2501 cm-1 to the combinatorial (1 + 3) mode (a mix of asymmetric and symmetric
stretching modes). These assignments are in agreement with several literature assignments [6, 7].

Fig 2.2. Fundamental modes of vibration for SO2 with wavenumber assignments and symmetries.The
above assignments also agree with the simulated spectrum of optimal geometry SO2 in the ab initio
software Gaussian 09 (see appendix 1).
SO2, as a non-linear triatomic molecule, has 3 6 fundamental vibrational modes, which works
out to be 3 modes. In addition, SO2 is a molecule with C2V point group symmetry, so all 3
fundamental modes should be IR active. These modes are denoted as the symmetric stretch (1), the
asymmetric stretch (3) and the bend (2). By consideration of the magnitude of the resultant dipole
moment upon vibration, a relative wavenumber ordering of 2 < 1 < 3 is obtained, where the (in
cm-1) for the stretching modes 1 and 3 are similar, and 2 is much lower, as bends are typically
lower energy modes than stretches.
3

Advanced Physical Practical Report William Hodds.

Equipped with this knowledge, the mistake of assigning the absorption at 2501 cm-1 to a
fundamental mode is avoided, as this band does not have a comparable to any other band
observed, so cannot be a stretch, and cannot be a bend by virtue of the high wavenumber it occurs
at relative to all other observed bands. Therefore, a proposed assignment for the band at 2501 cm-1
is to the combinatorial mode (1 + 3), as summing the observed bands for the 1 + 3 gives
2514 cm-1, which is appreciably close enough to 2501 cm-1 for this assignment to be reasonable. In
addition, by noting that the point symmetries of the modes 1 and 3 are A1 and B2 respectively, the
symmetry of the combinatorial mode (1 + 3) can be simply determined as the product of A1 and B2
in the C2V character table:
Table 2.3. The symmetry, and therefore IR activity, of the combinatorial mode (1 + 3), determined
via a point group symmetry treatment of the individual components of the combinatorial mode.

A1
B2
A1 x B2 (= B2)

C2

v (xz)

v (yz)

1
1
1

1
-1
-1

1
-1
-1

1
1
1

By taking this product, the symmetry of the combinatorial mode (1 + 3) is found to be B2. In the C2V
character table, B2 symmetry has a y component (perpendicular to principle C2 axis), so gives a ycomponent dipole change upon vibration, and the (1 + 3) mode is therefore IR active, fulfilling the
requirement that a mode must give a non-zero change in dipole moment upon vibration to be IR
active. Further to this, as a combinatorial mode, the intensity should be lower than a standard
fundamental mode band, and this indeed is the case here. All these points support the proposed
assignment of the 2501 cm-1 band to the combinatorial (1 + 3) mode.
The assignment of the bands at 1149 cm-1 and 1373 cm-1 is straightforward, as they are both known
to be fundamental stretch modes by virtue of their relatively high intensities. The higher
wavenumber band, at 1373 cm-1, is attributed to the asymmetric stretch mode, 3, due to being a
higher energy transition than the other stretch at 1149 cm-1, assigned to the symmetric stretch
mode, 1, by process of elimination.
The literature values of the fundamental SO2 bend mode, 2, from Briggs and from Isaacson et al. are
around 518 cm-1 [6, 7], so it is not surprising that no band is observable for this mode in the spectrum
presented here, as the lower resolution limit of the infra-red spectrometer used is seen to be ~700
cm-1, where the spectrum becomes increasingly noisy, presumably due to the IR-cell beginning to
absorb. It is not unreasonable therefore to propose that the bending band is masked by this low
wavenumber noise. In the high pressure (760 mmHg) IR spectrum of SO2, an potential vibrational
mode absorption is seen around 518 cm-1, however it is merged with the low wavenumber noise, so
this assignment is ambiguous.
Lastly, no significant overtones are expected to be observed at 298 K in this low pressure (12 mmHg)
spectrum of SO2, as for a harmonic oscillator, vibrational transitions of = + 2, + 3 + n (where n
>1) are selection rule forbidden, so should not be common transitions within a sample. Previously,
SO2 has been found to have only a small degree of anharmonicity in studied modes, shown by small
magnitude anharmonic force constants found for SO2 vibrational modes [8], therefore it is expected
that this selection rule is well obeyed, and any overtones that may be populated at 298 K are only
4

Advanced Physical Practical Report William Hodds.

(3 + 1)

(2 + 3)

23
(3 + 21)
OR H2O

21

populated in minor proportions, so will not be observed in the low pressure spectrum, if they are
populated at all.
Fig 2.4. SO2 IR spectrum at a pressure of 760 mmHg (1013 mbar, atmospheric pressure at sea level in
midlands UK [9]). In addition to the 3, 1 and (3 + 1) modes seen in the low pressure spectrum,
lower population modes are also observed at 1873 cm-1, 2295 cm-1 and potentially 2707 cm-1 and
3638 cm-1. Arguably, there is also an absorption at 531 cm-1 arising from the bending mode 2.
This higher pressure SO2 spectrum presents many more bands than in the low pressure spectrum fig.
2.1., all of which must either be n-overtones and/or combinational modes, owing to their relatively
low intensity.
To aid in assignment of the bands at 1873 cm-1, 2295 cm-1, 2707 cm-1 and 3638 cm-1, it is helpful to
obtain a range of possible transitions and then fit observed bands to these known transitions. This
will also confirm any previous assignments of bands to fundamental modes in fig. 2.1.. From the
work of Carney et al. [10], a table of experimentally-determined SO2 vibrational energy levels can be
adapted into the energies of realistic transitions:
Table 2.5. Table adapted from the work of Carney et al. [10] which shows the absolute energies of
common vibrational levels, and the energies of transitions from the ground vibrational state, = 0,
to a higher n-th state (n = 1,2). Observed transitions are in shaded rows.

/ cm-1

0 - n /
cm-1

Transition

0
0
0
2
0
0
1
0
1
1

0
1
2
0
0
1
0
0
1
0

0
0
0
0
2
1
0
1
0
1

1530
2047
2557
3835
4241
3404
2686
2890
3201
4035

517
1027
2305
2711
1873
1156
1360
1673
2505

2
22
21
23
(2 + 3)
1
3
(1 + 2)
(1 + 3)

Advanced Physical Practical Report William Hodds.

From the above data, it can be seen that previous assignments of the fundamental modes 1, 2, and
3 are further reinforced by the data of the absolute energies of the involved vibrational levels for
each transition. For 1, a wavenumber of 1155 cm-1 is observed, where a of 1156 cm-1 is reported,
for 2, 531 cm-1 is observed where 517 cm-1 is proposed, and for 3, 1358 cm-1 is observed where
1360 cm-1 is reported, all of which closely agree within the normal error range of IR measurements
(+ 10 cm-1).
In addition, the bands at 1873 cm-1, 2295 cm-1, and 2707 cm-1 are readily assigned to the transitions
(2 + 3), 22 and 23, respectively. An identical symmetry treatment to that in table 2.3. shows that
the mode (2 + 3) has B2 symmetry, 22 has A1 symmetry and 23 has A1 symmetry, meaning that all
proposed modes are indeed IR active, as they all give rise to a change in the dipole moment of the
molecule upon vibration, either in the z or y axes.
This leaves the low intensity, high energy band at 3638 cm-1, which does not fit the profile of any
simple binary combination or overtone, so must be a trinary or greater mode. The band could simply
be H2O impurity in the SO2 sample, as the O-H asymmetric stretch is known to occur around this
wavenumber range [11], though if this were the case, a band at approximately 1500 cm-1 would be
observed from the symmetric stretch fundamental mode, and this does not immediately stand out
in the spectrum, though what appears to be some noise at a comparable intensity to the band at
3638 cm-1 may well be this mode manifesting itself. Due to this possibility of H2O in the sample, the
assignment of the band at 3638 cm-1 to the trinary mode (3 + 21), which has a calculated
wavenumber of 3666 cm-1, is not completely unambiguous.
In summary, the assignments of all bands are shown in the table below:
Table 2.6. All bands in the high pressure (760 mmHg) IR spectrum of SO2 with proposed assignments
to vibrational modes, or impurities.

Observed / cm-1

Transition assignment

531
1155
1359
1873
2295
2515
2707
3638

2
1
3
(2 + 3)
21
(1 + 3)
23
(3 + 21) OR H2O

Calculation of the specific heat capacity, CV, of SO2


It is desirable to calculate the specific heat capacity of a molecule for both the practical usage of
such data in engineering as well as the fundamental insight into energy distribution amongst the
systems various energy levels at a given temperature, which is inherently a quantum phenomenon.
In order to calculate the specific heat capacity of SO2, a number of (justified) approximations must
be evoked, the first of which is the Born-Oppenheimer approximation: The translational, rotational,
6

Advanced Physical Practical Report William Hodds.


vibrational and electronic energy levels are well separated so do not interact. Therefore, the total
energy of a molecule, or ensemble of molecules, can be represented as the sum of these
components:
= + + +

(2.1)

Using this logic, any state function of a molecule that is proportional to energy, or the molecules
total partition function, may be expressed as the sum of translational, rotational, vibrational and
electronic components. Specific heat capacity at constant volume, CV, is one such property of a
system:

, = ,

(2.2)

Where = , , .
Therefore, to calculate the specific heat capacity at constant volume of SO2, each contributing term
must be calculated in term, then summed.
The electronic contribution to heat capacity at 298 K is, for most molecules, zero, as the population
of higher electronic energy levels at 298 K is approximately zero. This deduction is drawn from
noting that the energy gap between = 0 and = 1 levels is too large to be traversed by any
molecule at the relatively low temperature of 298 K. Quantitatively, this concept is shown by the one
of the key foundations of statistical thermodynamics, the Boltzmann distribution:

(2.3)

Where Pi is the probability of transition i occurring, is the energy gap between two states in a
transition, kB is the Boltzmann constant and T is the temperature in K.
It can be seen in equation 3 that the probability of a transition occurring (or a state being occupied)
is directly proportional to the energy gap between the two states, . If kBT is , 0. A ready
intuitive comparison between and T can be made if is converted into temperature units,
appropriately named rotational temperature for rotational energy gaps, translational temperature
for translational energy gaps, and so on. If this is done between the ground electronic state and first
excited state of SO2, which has been shown to be excited at around 300 nm (33333.3 cm-1) [12] , the
following result is obtained:
=

= 47 966.6

(2.4)

where c is the speed of light in cm s-1, h is planks constant and kB is the Boltzmann constant.
Immediately it is clear that 298 K does not rival 47966.6 K in magnitude, and it can be safely
assumed that the first excited state of SO2 is not occupied significantly at 298 K, therefore no energy
can be distributed amongst the electronic energy levels, and the heat capacity electronic

contribution is zero, ,
= 0. In other cases, however, where the between electronic energy
7

Advanced Physical Practical Report William Hodds.


levels is comparable in magnitude to T, this contribution would be required to obtain an accurate
heat capacity value. An example of such a system would be a metal, or metal cluster [13], where
electrons are readily excited and even ionised by UV radiation, via the well known photoelectron
effect [14].
For the rotational and translational contributions, the classical equipartition of energy
approximation can be made to greatly simplify calculation of the heat capacity contributions. The
equipartition of energy approximation assumes classical distribution of states, which are not
considered to be quantised but instead on a continuum, with a = 0 no particular state has
preference, and each state is occupied equally, giving an equal partition of a molecules energy
amongst all states. In this approximation, an average energy of kBT can be assigned to any degree
of freedom, or R for a mole of molecules.
Therefore, as a non-linear molecule, SO2 has 3 degrees of translational freedom and 3 degrees of
rotational freedom:
(2.5)
1
3

, = , = (3 ) =
2
2
Where R = the universal gas constant, 8.314 J K-1 mol-1.
From this, the heat capacity contribution for both rotational and translational modes of freedom, is
in total, 24.94 1 1.

Advanced Physical Practical Report William Hodds.


Applying the equipartition theorem to the translational and rotational energy levels is justified by
the vanishingly small between translational and rotational energy levels:
1

0, ,
( )
2

(2.6)

Translational modes of freedom are as classical as they come, they are indeed the modes which first
invoked the development of classical mechanics, which was first developed to model translational
motion. An entity in translation may adopt any given position in the x, y or z axis, as = 0, it can
exist in a functional continuum of positions/states at nearly any temperature. Although it is known
that even space is quantised [15], due to the infinitesimally small magnitude of the between
translational states, the extent of which space quantisation manifests in chemical systems is
effectively non-existent. Because of these factors, it not useful to attempt to calculate the
translational temperature, T, of an SO2 translational transition, except to say it is conceptually very
large.
The same concept applies to the 3 rotational degrees of freedom, which although 0, 0. A
rotational temperature, R, can be calculated for a given rotational transition in SO2, by using the
rotational fine structure observed in the low pressure IR spectrum of SO2 for the 3 mode, the
spacing between the maximum intensity peaks of the P and R branches, (max), will give a R that
can be considered conceptually to show rot << kBT, and thus that the system obeys the
equipartition theorem well for rotational modes:

(max)

Fig 2.7. Expanded 1445 1320 cm-1 region of the low pressure IR spectrum of SO2, showing
rotational fine structure for the 3 mode, and that (max) = 24.9 cm-1.
Noting that (max) = 24.9 cm-1, T = 298 K, and 0 = 0.344174 cm-1 [16], R may be calculated by
equation 6:
(

8
02
) 0.455
(max)2

(2.7)

Immediately it can be seen that << T, and therefore total occupation of rotational energy levels in
SO2 is expected, and that applying the equipartition assumption (that 0) is justified and a good
approximation to the real system of SO2.
9

Advanced Physical Practical Report William Hodds.

For the vibrational contribution to heat capacity, the energy levels are not closely spaced and
T, therefore a quantum treatment is appropriate, and only the fundamental vibrational modes that
are excited at 298 K must be considered. For this, the low pressure IR spectrum (fig 2.1.) of SO2
yields the wavenumbers of the = 0 = 1 transitions for all 3 fundamental modes, 1, 2 and 3,
discussed in detail earlier in this section, these wavenumbers are summarised in table 2.6..
Firstly, the vibrational temperature is calculated for each respective mode using equation 3, this
results in:
1 =

1 (6.626 1034 2.998 106 1 1155 1 )


=

1.381 1023 1

= 1662.10
2 =

2 (6.626 1034 2.998 106 1 531 1 )


=

1.381 1023 1

= 764.11
3 =

3 (6.626 1034 2.998 106 1 1359 1 )


=

1.381 1023 1

= 1955.60
Using the calculated vibrational temperature for each mode ( 1 , 2, 3 ), the vibrational
contribution to the constant volume heat capacity, from each mode, may be calculated, using the
following relation:

= [

(
)]

(2.8)

Where is the vibrational temperature of mode n, R is the universal gas constant, and T is the
temperature at which the constant volume heat capacity is desired.
Using equation 7, and the vibrational temperatures, the contribution from each mode is:

1 = 0.986 1 1

2 = 4.940 1 1

3 = 0.507 1 1

10

Advanced Physical Practical Report William Hodds.


and the overall vibrational contribution to constant volume heat capacity, at 298 K, is the sum of all
occupied vibrational modes at this temperature, here 1, 2, and 3:

= 6.43 1 1

(2.9)

2 key observations can be made here:

1. As < 3, energy amongst vibrational levels is not evenly dispersed (equipartition is not
observed), therefore the previous assumption, that vibrational energy level spacing is significant
enough to warrant a non-classical treatment of the modes, is justified at 298 K. For any of the
modes, is not >> 298 K, therefore this result is to be expected, as higher levels are not accessible.

2. 2 = 4.940 1 1 , which is 76.8% of the total vibrational contribution. This means that at
298 K, mode 2 (the bending mode) is most excited, and this is due to the low energy requirement of
531 cm-1 to cause a = 0 v = 1 transition, which allows higher order transitions to accessible,
thereby storing more energy. 1 and 3 contribute only minorly, 7.9% and 15.3% , again reflecting the
relatively high requirement for the fundamental excitation, due to the larger energy gaps.
Therefore, as:

() 0,

Further to point 1, by plotting

against T for each mode, the degree of equipartition in the

vibrational mode, at a given temperature, can be judged by the position of the temperature on the
resulting curve, which takes on a sigmoidal shape:

0.9
0.8

CV/R / no units

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
100 298

600

1100

1600

2100

2600

T/K

Fig 2.8. Plot of CV/R vs T for vibrational mode 1 of SO2 (symmetric stretch), at 298 K,
11 equipartition is not observed.

Advanced Physical Practical Report William Hodds.

1
0.9
0.8

CV/R / no units

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
100 298

600

1100

1600

2100

2600

T/K

Fig 2.9. CV/R vs T for vibrational mode 2 of SO2 (bend). At 298 K, the equipartition value of 1 is
approached, but is not achieved.
1
0.9
0.8

CV /R / no units

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
100 298

600

1100

1600

2100

2600

T/K

Fig 2.10. CV/R vs T for vibrational mode 3 of SO2 (asymmetric stretch), at 298 K, equipartition is not
observed.
From these plots it can be deduced that no vibrational mode for SO2, at 298 K, behaves classically.
This is simply because is not << kBT when T = 298 K.

12

Advanced Physical Practical Report William Hodds.


For modes 1 and 3, CV R only at T > 3000 K, only at these very high temperatures will the
population of all vibrational energy levels be equal, much like the populations of translational and
vibrational levels in SO2, at 298 K.
At T 100 K, as CV/R 0, it can be deduced that only the = 0 vibrational level is occupied and no
higher order vibrational levels are occupied, much like the electronic energy levels of SO2 at 298 K.
For mode 2, however, CV R at only ~1600 K, which is relatively low compared to the other two
vibrational modes. As such, at 298 K, a significant fraction of the SO2 population will exist in higher
order vibrational levels of the 2 mode, and therefore more thermal energy may be stored in this
mode, a reflection of this property being the relatively high contribution to the total vibrational
constant volume heat capacity of 2 when compared to 1 and 3.
With this knowledge, an appropriate schematic representation of different energy levels in SO2 is
shown below:

Electronic

Rotational

Translational

Mode 1, 1

Mode 2, 2

Mode 3, 3

Fig 2.11. Schematic representation of each category of energy level in SO2, showing relative
scales of magnitude of between the vibrational modes 1,2 and 3. It can be seen that of 3
is most like that of the electronic energy levels, and 2 is most like that of rotational energy
levels.
13

Advanced Physical Practical Report William Hodds.


Finally, the total constant volume heat capacity is the sum of all contributions from all accessible
degrees of freedom in SO2 at 298 K, 12 mmHg (due to equation 1), therefore:

, = ,
= + + = (12.471 + 12.471 + 6.43) 1 1

= 31.372 1 1
This value is remarkably close to the literature value for CV, m of SO2 (at 298 K), which is 32.8 J K-1
mol-1 [17], despite the inherent assumption that SO2 is an ideal gas, despite this not truly being the
case. In order to calculate the value of , , SO2 molecules and their energy distributive behaviours
have been considered independent of other SO2 molecules, that is to say a hard sphere
approximation has been made, and any SO2-SO2 interactions have been ignored.
This is not a truly realistic approximation however, as SO2 adopts a bent optimised geometry (point
group C2V), and there is a finite difference in electronegativities between S and O
(S = 2.44, o = 3.50 (Allred-Rochow [18] )), therefore the SO2 molecule must possess a dipole,
allowing the possibility of permanent dipole-permanent dipole intereactions within a sample of SO2.
These additional interactions would serve to increase the heat capacity of SO2, as additional thermal
energy input would be necessary to increase the temperature of partially-bound, interacting SO2
molecules.

Fig 2.12. Conceptually possible SO2-SO2 permanent dipole-permanent dipole interactions which
would tend the behaviour of SO2 away from ideality, and increase CV, m relative to an ideal SO2
sample.
For an ideal gas, the Van der Waals a constant is ~ 0, as the a constant is a measure of gas-gas
interactions in a sample, however for SO2, a = 6.803 L2 bar mol-2 (for comparison, He, a = 0.035 L2 bar
mol-2) [19]. This shows quantitatively that there is a non-zero attraction between SO2 molecules, but
perhaps the low pressure of the sample (12 mmHg) deems these interactions as negligible.

14

Advanced Physical Practical Report William Hodds.


Determination of the temperature at which a vibrational-rotational IR spectrum was recorded, for
CO, and direct observation of the Boltzmann distribution in rotational fine structure
In this sub-section an approach to finding the temperature at which a given (high resolution) IR
vibrational-rotational spectrum was recorded, is shown in practice, using a second overtone of CO as
an example.
An initial estimate of the T can be obtained readily from the spacing between the maximum intensity
peaks in the rotational fine structure of the P and R branches, respectively, of any vibrational
transition, if the rotational constant, 0 , and the rotational temperature, R, is known, using
equation 6, but rearranged so that T is the subject:
(

(max)2
)
802

(2.10)

(max)

Fig 2.12. The expanded rotational fine structure of a second overtone in the IR spectrum of CO, at
unknown T. (max) is shown to be 51.14 cm-1.

Here, (max) = 51.14 cm-1, 0 = 1.915 cm-1, and R = 2.76 K. Therefore, the approximate T at
which the spectrum was recorded is:

= 246.04
This value is not likely to be correct, especially as the boiling point of SO2 at 760 mmHg is 263 K[23].
In order to obtain a more accurate value for the temperature, a series of calculations can be
conducted which evoke the concept of the Boltzmann distribution, which the P and R branches
15

Advanced Physical Practical Report William Hodds.


conceptually have the form of, though it must be noted that the P and R branches are not
symmetrical to eachother, the R branch peak-peak spacing ((+1 )) decreases with increasing
whereas the P branch peak-peak spacing ((+1 )) increases with increasing . This effect is,
later in this section, rationalised as the result of rotational-vibrational coupling and centrifugal
distortion in CO.
In order to predict the temperatures, the intensities of each rotational transition must be measured
quantitatively, as well as assigned to a specific transition. In the spectrum studied here, all
peaks (until very large /) are well resolved. transitions have been expressed as values of
m, where ( + 1) = , ( + 1) = , for P and R branches respectively. For the R branch, >
0 and for the P branch, < 0.
Assignments for both the P and R branch were made up to = 21, at which point resolution is
too poor to differentiate baseline noise from actual rotational transitions. All assignments and
manually measured intensities are shown in appendix 2.
To account for the different values of in the P and R branches, the g value is introduced, which
standardises values by the following definition:
= 2, for P branch
= 2, for R branch
thereby accounting for the different signs of in the P and R branches.
Next, a modified form the Boltzmann distribution can be applied to this system. The intensities of
rotational fine structure peaks in simple heterodiatomic molecules, like CO, are given by the
following equation [21] :

= ( + + 1)

( +1)

(2.11)

However, by noting that:


= ( + 1), ( + 1)
= 2, 2

(2.12)
(2.13)

Equation 10 can be written in terms of and :

(1)

(2.14)

At this point, equation 13 has the form of a Boltzmann distribution model, with the major variables
I, R, and T featuring in analogous ways to which P, , and T feature in the original relation:

(2.15)

Where P is the probability of a state being occupied, A is the pre-exponential factor, i is the energy
gap between ground and the level concerned, kB is the Boltzmann constant and T is the
16

Advanced Physical Practical Report William Hodds.


temperature.

By taking the natural log of both sides of the equation 13, it takes on a format that can be analysed

by simple linear regression, plotting ln () against ( 1) will give as the gradient of the
equation of the line:
(( 1) )

ln ( ) = (
)

(2.16)

Table 2.13. Intensities of assigned m (-( + 1), ( + 1)) rotational peaks in the second overtone
mode of CO, used for the linear regression plot in figure 2.14..

17

R
m(m-1)R / K
0
5.52
16.56
33.12
55.2
82.8
115.92
154.56
198.72
248.4
303.6
364.32
430.56
502.32
579.6
662.4
750.72
844.56
943.92
1048.8
1159.2

ln(I/g)

P
m(m-1)R / K

ln(I/g)

2.169
2.031
1.946
1.972
1.792
1.727
1.631
1.609
1.473
1.267
1.068
0.838
0.767
0.463
0.288
0.104
-0.159
-0.405
-0.804
-1.124
-1.598

5.52
16.56
33.12
55.20
82.80
115.92
154.56
198.72
248.40
303.60
364.32
430.56
502.32
579.60
662.40
750.72
844.56
943.92
1048.80
1159.20
1275.12

1.705
1.981
1.764
1.853
1.732
1.609
1.341
1.279
0.927
0.718
0.506
0.486
0.019
-0.055
-0.406
-1.023
-1.329
-1.567
-1.692
-2.079
-2.639

Advanced Physical Practical Report William Hodds.

2.3
R branch

P branch

1.8
1.3
0.8

ln(I/g)

0.3
-0.2 0

200

400

600

800

1000

1200

1400

-0.7
-1.2

y = -0.003x + 2.0352
R = 0.9963

-1.7

y = -0.0036x + 1.9168
R = 0.9921

-2.2
-2.7

m(m-1)R / K

Fig 2.14. Linear regression plot of ln(I/g) against m(m-1)R, for a second overtone mode of CO, for
both the P (triangle markers) and R (square markers) branches.
The differences in the P and R branches are manifest here as each regression gives a different final T:
1

1
0.0036 1

= = 277.7

(2.17)

For R branch: = 0.0030 1 = = 333.3

(2.18)

For P branch:

For both regressions, R2 is ~1, so a Boltzmann distribution is followed to an extent, however because
R2 1, it can be inferred that there is some small departure from a Boltzmann distribution, and this
effect is more pronounced for the P branch than for the R branch.
The observation that 0 cannot be physically correctly describing the CO system, a
macroscopic sample of gas cannot exist at two different temperatures, as temperature is an average
macroscopic property of a substance. Therefore, a behaviour of the rotating and vibrating CO
molecule is not being accounted for by the model used in this calculation, the rigid rotor
approximation of a diatomic molecule.
The rigid rotor model describes a fixed bond-length diatomic rotating about an axis perpendicular to
the internuclear axis. This model, therefore, does not account for any effect where bond length
changes during the rotation, so cannot account for rotation-vibration coupling/interactions or
centrifugal distortions. These two effects are the source of the discrepancy of peak-peak spacing
between the P and R branches, and are described in the following subsection.

18

Advanced Physical Practical Report William Hodds.


The vibrational-rotational coupling interaction and centrifugal distortion
In vibrational-rotational coupling, vibrational modes of the diatomic molecule interact with
rotational modes, thereby influencing the relative energies of the rotational energy levels, therefore
the assumption, that R and V are massively different in magnitude, is not made. Rotationalvibrational is not significant in fundamental transitions (0 1 ) but becomes increasingly
significant with increasing , therefore it is unlikely to be a negligible contribution for the second
overtone of CO, where =3.
Vibrational-rotational coupling can be understood by considering the effect of vibrations of the
chemical bond on the bond length, . When the diatomic molecule vibrates, by definition the
internuclear distance is elongated, thereby necessitating an increase in , at the beginning of a
vibrational cycle. When increases, the moment of inertia, , of the diatomic molecule increases by
the following relation:

= 2

),
+

where is the moment of inertia, = (

(2.19)

the reduced mass of CO, and is the bond length.

And when increases, there is a corresponding decrease in the rotational constant :

82

(2.20)

This decrease in causes a change in the energy of any given rotational level (and therefore the
spacing between peaks of rotational transitions in the IR spectrum) dependent on what type of
transition the rotational level is associated with (which branch the peak is observed in):

For the P branch: 1 = 0 2

(2.21)

( + 1)
For the R branch: +1 = 0 + 2

(2.22)

where 1 is the energy of a given transition, is Plancks constant, 0 is the initial vibrational
quantum number of the vibrational component of the band.
Therefore, from equations 19-22, for the P branch: as increases, decreases, therefore the entire
term in equation 21 decreases, leading to an increase in 1 . As this increase in 1
2
becomes larger with higher levels, this causes the spacing between peaks, 1 , to increase
accordingly with the decrease in .
Similarly, for the R branch, the same logic may be applied to find that the 2 ( + 1) term in
equation 22 decreases, however because this term in additive in equation 22 rather than
subtractive, like in equation 21, +1 actually decreases, resulting in a decrease in +1 ,
causing the rotational peaks becoming increasingly similar in energy as increases.

19

Advanced Physical Practical Report William Hodds.


To illustrate the change in across vibrational levels, the combination differences approach can be
taken. This method takes advantage of the fact many transitions in the P and R branches share a
common state (the term for combination of rotational and vibrational energy levels associated in a
transition), therefore, any differences between the observed for two transitions sharing an excited
or ground state will be dependent on the properties of the unshared state[22].
Therefore, to find information about the upper state, such as the value of 3 , the difference in
transition energies (in the P and R branches) involving a common ground state must be calculated.
From the observed rotational fine structure in the spectrum of the second overtone of CO (fig 2.12.),
the following assignments can be made:
Table 2.15. Assigned transitions with manually measured , for the second overtone band in CO,
fig 2.12., R branch transitions are in shaded columns and P branch transitions are in white columns.
Note that all rotational transitions are accompanied by a vibrational transition with = 3.
/ cm-1
6354
6359
6362
6365
6369
6371
6347
6344
6338
6335
6331
6326

transition
0 1
1 2
2 3
3 4
4 5
5 6
1 0
2 1
3 2
4 3
5 4
6 5

Using these measured wavenumbers, the difference in energies between common ground state
transitions ( ) can be determined, which can be shown to be equal to:
1
= 43 (0 + )
2

(2.23)

Therefore, a linear plot of against (0 + 2) will give a line with a gradient 43.
Table 2.16. The combination differences, , of R and P branch transitions up to 0 = 5, from
the assignments made in table 2.15, of the rotational fine structure of fig 2.12..
/ cm
7
15
24
30
38
45

20

-1

1
(0 + )
2
0.5
1.5
2.5
3.5
4.5
5.5

Advanced Physical Practical Report William Hodds.

50
45

y = 7.5714x + 3.7857
R = 0.9977

40

/ cm-1

35
30
25
20
15
10
5
0
0

(0 + 1/2)

Fig 2.17. Combination differences plot for rotational transitions up to 0 = 5, using the data in table
2.16.. A high R2 value can be seen, indicating both good quality data and the absence of significant
centrifugal forces in these transitions, where equation 2.23 would begin to break down. The slope of
the line is equal to 43 .
Therefore, 3 can be shown to equal:
3 =

7.5714 1
=
= 1.893 1
4
4

(2.24)

From this calculated value of 3 , it can be seen that 0 3 , however the difference is small at
only 0.022 1. This small, but finite change in the rotational constant is one of the factors
contributing to the departure of the observed spectroscopic data from the predictions of the
harmonic rigid rotor model, and therefore the inaccuracy of the predicted temperatures and .

21

Advanced Physical Practical Report William Hodds.

P branch
1

121213
2

21

42

63

84

0
1
8
9 10 121213
2 ( + 1) 20
41

10

105 126 147 168 189 2010

R branch
+1

62

3
83

104

125

146

Fig 1.15. An exaggerated schematic representation of the convergence property of the P branch
rotational transition peaks and the divergence property of the R branch rotational transition
peaks, showing a higher weight B term as J increases, with consequences on the peak-peak
spacing, .
In centrifugal distortion, the diatomic bond length increases as the rotational quantum number
increases, due to an increase in moment of inertia, as in equation 18. This concept originates
from classical mechanics, where it is observed that the rotational velocity of two spheres
connected by a spring is proportional to the length of the spring, thus an analogy with the
chemical bond was drawn.

Fig 1.16. Schematic representation of the centrifugal force in a simple diatomic such as CO. As the
angular velocity increases, increases, resulting in a decrease in the rotational constant B.
Taking into account centrifugal distortion, the energy of a rotational transition becomes:
() = ( + 1) 2 ( + 1)

(2.23)

Where the constant is the centrifugal distortion constant, given by:


43
2

Where is the vibrational wavenumber of the mode concerned.


=

22

(2.24)

Advanced Physical Practical Report William Hodds.


From equation 23 it can be seen that is several magnitudes smaller than , as the for CO in
figure 2.12. is in the range of 102 103. Only when is very large will the entire 2 ( + 1) term
will begin to significantly affect the spacing between rotational energy levels, this is analogous to
centrifugal forces becoming greater when the angular velocity of the two sphere-spring system
increases.
In summary, the spacing between rotational energy levels, , is not fixed at 2, nor is
constant, as the rigid diatomic rotor model predicts, but instead is dependent on , the
centrifugal distortion term, and , both of which are dependent on the levels associated in a given
rotational transition, and become more significant at higher levels. Therefore, the rotational
temperature used for the prediction of T in equation 15 is not valid for all levels, and will become
increasingly irrelevant to the true energy of rotational levels as increases. The fact that 0 1
and the existence of the 2 ( + 1) term was not accounted for in the previous analyses, leading
to both a discrepancy between and and unreliable values of and in their own right.

Conclusion
In this report the IR spectrum of SO2 is assigned in detail with aid from group symmetry theory and

Gaussian software. This was done to calculate the vibrational contribution, ,


, to the molar
constant volume heat capacity, , , at 298 K, for SO2 using purely data from SO2 IR spectra
collected at 760 mmHg and 12 mmHg, to a high accuracy of 95.65 % (calc. , =
31.372 1 1 , lit. , = 32.80 1 1). This was achieved using a statistical
thermodynamics approach, involving the sum of heat capacity contributions from all accessible
degrees of freedom, translational, rotational and vibrational. The electronic contribution was
deemed negligible due to , any excited electronic states were shown to be too high in
energy to be accessible at 298 K by pure thermal excitation, by comparison of the electronic and
ambient temperatures, and . Translational and rotational energy level spacing was deemed
either continuous ( = 0) or pseudo-continuous ( 0), so the respective heat capacity

contributions were treated using classical equipartition theory, where both ,


and ,
3

contribute 2 each. The high accuracy of the calculated , value suggests that all assumptions
made are largely correct, but also that any SO2-SO2 intermolecular interactions do not seem to be
significant enough to contribute to , in any major way, as a fairly accurate value can be obtained
without their consideration.
In addition, an attempt was made to calculate the temperature at which an IR spectrum of CO was
recorded using a rigid rotor approach and fitting the P and R branches to a modified form of the
Boltzmann distribution. This was not successful, as two different temperatures were obtained, one
from each branch, owing to the breakdown of the rigid rotor model of a diatomic molecule at high
and levels, of which the second overtone P and R branches that were studied were associated
with. This is discussed in some detail, explaining the tendency of the R branch rotational peak-peak
spacing ( ) to converge (or approach zero) and the P branch rotational peak-peak spacing ( )
to diverge (or approach ) at high rotational energy levels. These tendencies were attributed to the
change in , the rotational constant, due to both centrifugal distortion at high levels and
rotational-vibrational coupling at the high level of the coupled vibrational transition of the
overtone. The change in when = 3 for a transition was shown to be 0.022 1, a small but
finite change, determined via the combinational differences approach.
23

Advanced Physical Practical Report William Hodds.

References
[1] - R. Fu, A. Moullet, N. Patel, J. Biersteker, K. Derose and K. Young, The Astrophysical Journal, 2012, 746, 42.
[2] - M. Wittkowski, P. Hauschildt, B. Arroyo-Torres and J. Marcaide, Astronomy & Astrophysics, 2012, 540,
L12.

[3] - Scalo, J. M., & Slavsky, D. B. 1980, ApJ, 239, L73


[4] - O. Hashimoto, H. Izumiura, D. Kester and T. Bontekoe, Astrophysics and Space Science, 1995,
224, 477-478.
[5] - Tsuji, T., Ohnaka, K., Aoki, W., Yamamura, I., 1997, A&A 320.
[6, 8] - A. Isaacson, 2016. J. Chem. Phys. 75, 3017 (1981)
[7] - A. Briggs, Journal of Chemical Education, 1970, 47, 391.
[9] - Metoffice.gov.uk, 2016.
[12] - G. D. Carney and C. W. Kern, Int. J. Quantum Chem. Symp. 9, 317 (1975).
[11] - D. F. Coker, J. R. Reimers and R. O. Watts Aust. J. Phys., 1982, 35, 623-38).
[13] - Thornton, Steven T, and Rex. Andrew, Modern Physics for scientists and engineers, Saunders
College Publishing, 1993.
[14] - H. Sauceda, J. Pelayo, F. Salazar, L. Prez and I. Garzn, J. Phys. Chem. C, 2013, 117, 1139311398.
[15] - I. Rabi, Phys. Rev., 1937, 51, 652-654.
[16] - R. PRATT, A. RON and H. TSENG, Reviews of Modern Physics, 1973, 45, 273-325.
[17, 20] - Encyclopedia.airliquide.com, 2016.
[18] - Weast. R. C. (Ed.), Handbook of Chemistry and Physics (53rd Edn.), Cleveland:Chemical Rubber
Co., 1972.
[19] - A L Allred and E G Rochow, J. Inorg. Nucl. Chem., 5, 264 (1958).
[21] - Herzberg, G. Molecular Spectroscopy and Molecular Structure, I Spectra of Diatomic
molecules, 1st Edition (1950, 1972), 2nd Edition (1989) Chap 3(e) page 121.
[22] - 10.P. Atkins and J. De Paula, Physical chemistry, 10th edn., 2010.

24

Advanced Physical Practical Report William Hodds.

Appendix 1. Simulated IR spectrum of geometry optimised SO2. Ab initio software Gaussian 09 used
for the modelling of SO2, using method: Hartree-Fock (restricted), basis set: 3-21G.
Immediately it can be seen that the initial assignments of fundamental modes in fig 2.1. (low
pressure SO2 IR spectrum) are correct, however unfortunately binary (or higher) modes are not
within the predictive power of this method and basis set using Gaussian 09, so the assignments
made in fig. 2.4. (high pressure SO2 IR spectrum) cannot benefit from the same verification.

3
2

25

Advanced Physical Practical Report William Hodds.


Appendix 2. Unprocessed assignments of m to individual rotational fine structure peaks, as well as
the measured intensity for each peak, of the second overtone mode of CO. Both P and R branch
assignments are shown.
R branch

P branch

26

height (I) / mm

height (I) /mm

-1
-2
-3
-4
-5
-6
-7
-8
-9
-10
-11
-12
-13
-14
-15
-16
-17
-18
-19
-20
-21

11
29
35
51
56.5
60
53.5
57.5
45.5
41
36.5
39
26.5
26.5
20
11.5
9
7.5
7
5
3

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21

17.5
30.5
42
57.5
60
67.5
71.5
80
78.5
71
64
55.5
56
44.5
40
35.5
29
24
17
13
8.5

Вам также может понравиться