Sunteți pe pagina 1din 45

Giving the spherical cow its figure back: towards a

more realistic model of diatomic molecules

Gillian Kwan, Kimberly Edwards, and Leah Huiting

Department of Chemistry, Mount Holyoke College, South Hadley, MA 01075

∗ To whom correspondence should be addressed

1
Abstract

The simplified quantum mechanical models of vibrational and rotational motion in di-

atomic molecules were revised using perturbation theory, and the applicability of these new

models were evaluated by spectroscopic methods. Vibrational motion of diatomic molecules

has traditionally been modeled by quantum harmonic oscillator when introducing quantum

mechanics. This classically derived model is simple and didactic at first, but its validity is

harrowingly limited to modeling diatomic molecules with equilibrium geometries, rendering

it inapplicable in most realistic description of vibrating molecules. The vibronic spectroscopy

of iodine proves that molecular bonding is far from harmonic and only potentials that account

for anharmonicity and bond dissociation (e.g. Morse potential) can provide a satisfactory de-

scription of vibrating molecules. From the electronic spectrum of I2 , excited state iodine was

found to have a vibrational frequency ωe0 of 132.514 cm−1 , anharmonicity constant ωe0 χe0 of

0.987 cm−1 , dissociation energy D00 of 4316.268 cm−1 at vibrational quantum state v0 = 66,

and force constant k0 of 65.61 N · m−1 . Perturbation theory is applied to the rigid rotator model

of molecular rotation to account for the non-rigid reality of molecular bonding. Gas phase

infrared spectra were obtained for H35 Cl, H37 Cl, D35 Cl and D37 Cl. The constants directly

derived from the least squares fitting procedure of the vibrational-rotational spectra indicate

a pronounced isotope effect from isotopic substitution of hydrogen atom, but not for isotopic

substitution of chlorine atom (e.g. centrifugal distortion constant D ≈ 0.0007 cm−1 in HCl,

D ≈ 0.002 cm−1 in DCl). The parameters computed from these constants also display the

same trend (e.g. rotational constant in lowest state B0 ≈ 10.5 cm−1 for HCl, B0 ≈ 5.4 cm−1 in

DCl). Isotope effect is only observed upon deuteration of hydrogen because only a change in

the lighter atom of a heteronuclear diatom would cause an appreciable change in the dynamics

within the molecule.

Introduction

Milk production at a dairy farm was low so the farmer wrote to the local university,
asking help from academia. A multidisciplinary team of professors was assembled,

2
headed by a theoretical physicist, and two weeks of intensive on-site investigation
took place. The scholars then returned to the university, notebooks crammed with
data, where the task of writing the report was left to the team leader. Shortly thereafter
the farmer received the write-up, and opened it to read on the first line: “Consider a
spherical cow...". 1

The “spherical cow" shown above is an age-old joke poking fun at how scientists often study
complex real life scientific phenomena by using models so overly simplified, such as a spherical
cow in a vacuum, that they no longer reflect the reality. While this joke has greatly exaggerated the
simplification approach for comedic effect, it does underline the pillar of science: systematically
stripping away extraneous details to study the essentials. Reducing a problem into simpler form is
a necessary first step – it serves as the basis of all scientific education curricula. Without learning
about the simplest model, particle-in-a-box, solving the Schrödinger equation for systems with
potential energy operator in addition to kinetic operator would seem dauntingly overwhelming.
Without building a repertoire of quantum mechanical models, approximation methods such as
perturbation theory and variational principle would appear to be encrypted codes. As semester
progresses, we accumulate skills and knowledge so that we can update our understanding based on
the simpler models. This is the origin of the title, “Give the spherical cow its figure back" — with
updated knowledge, the spherical cow model will resemble reality more, or in spherical cow’s case,
more cow-like. The limitation of particle-in-a-box model of electronic transition has been explored
in previous experiments, in this paper we will revise the models of harmonic oscillator and rigid
rotator, and evaluate their validity using the molecular structure data extracted from spectroscopic
experiments on molecular iodine and hydrogen chloride.
Before delving into the two quantum mechanical models, let us take a step back: why do we
want a more accurate depiction of molecular motions and how is it related to spectroscopy? The
answer is straightforward: molecular motions give rises to energy, and upon irradiation in spec-
troscopy the energy in a molecule changes and is captured by a spectrophotometer as a spectrum.
Because spectrometers do not share our “spherical cow" view of molecules, we need to upgrade

3
our model to one that better reflects reality in order to understand the data obtained.

Origins of spectra

In classical thermodynamics, the total internal energy is partitioned into translational, rotational,
vibrational, and electronic energies. In quantum mechanics, spectroscopic studies of molecular
structures are largely facilitated by the Born-Oppenheimer approximation, which is invoked to
separate electronic and nuclear motions on the ground that the relative massiveness of nuclei puts
it on a different length scale and time scale than those of electrons, thus permitting separation
of electronic and nuclear energies. This separation of energies is further divided into four types
of energy levels much like in thermodynamics, except that the energy values are quantized as
illustrated in Figure 1. The spacing between each translational state is 1 × 10−37 J, 2 minuscule
enough to assume a continuous density of state distribution; for this reason translational energy
is considered a strictly classical concept, and it does not contribute to the quantum mechanical
discussion here. The energy level spacings in the other three energy types are large enough to be
qualified as fully quantized, thus they have discrete density of state distributions. The significance
of having a discrete distribution of density of states is that a state transition can only be induced by
energy larger or of the same magnitude as the spacings between energy levels. The quanta in each
energy types are matched by different regions in the electromagnetic radiation spectrum, and each
molecular process associated with these energies and radiation regions are summarized in Table I.

Table I: Electromagnetic (EM) radiation regions ranked by descending order in energy and
their relations

EM region molecular process information relayed


X-ray excitation of core electron
electronic structure and bond energy
Ultraviolet and Visible excitation of valence electron
Infrared bond vibration bond strength
Far infrared and Microwave bond rotation bond length
Radio nuclear spin change molecular structure

If a photon possesses energy hν greater or equal to the magnitude of energy level spacing in
a molecule, the molecule can absorb the photon energy and ascend to an excited state, then relax

4
back to the ground state via radiative or non-radiative decay. The interaction of molecule and
photon described forms the basis of spectroscopy. In addition to having frequency matching the
energy level spacing (∆E = hν), there are several requirements for the absorption and/or emission
of a photon by a molecule: (1) either the electric or magnetic field of the radiation must react with
a time dependent property of the molecule; (2) there must be molecules in the initial energetic
state for transition; (3) there must be fewer molecules in the final state than in the initial state. The
probability of transition from an initial state i to a final state j is proportional to R2 , the square of
transition moment integral R,
Z
R= ψi Ôψ j dτ (1)

where Ô is the appropriate operator for interaction between radiation field and a time dependent
property of the molecule, such as electric or magnetic dipole. In other words, whether a molecule
would absorb a certain type of radiation depends on whether it has a time dependent property that
would interact with the particular type of radiation. Since separations of rotational energy levels
correspond to the microwave region of the electromagnetic spectrum, a molecule must absorb mi-
crowave radiation to produce absorption peaks in pure rotational spectroscopy, and that prerequisite
time dependent property is a permanent dipole. The radiation used in vibrational spectroscopy is
infrared radiation, and a molecule can absorb infrared radiation only if its dipole moment changes
when there is a change in internuclear distance. These conditions are the gross selection rules for
pure rotational spectroscopy and vibrational spectroscopy. 3 The transition moment integral R gives
rises to specific selection rules, which are restrictions on the quantum numbers of the energy levels
available for transitions. For vibrational transition, the selection rule is ∆v = ±1, which means
that the initial state and the final state differs only by one quantum number after interaction with
a radiation field. Non-fundamental transitions such as ∆v = ±2, ±3 etc. may also be observed
because the vibrational selection rule is derived from approximations. The selection rule for rota-
tional transition ∆J = ±1 is derived directly from angular momentum and therefore more rigorous
than the vibration one. Selection rules narrow down the possible transition states, and in the case of
vibration and rotation, v ± 1 and J ± 1. There is no selection rule imposed on electronic transition,

5
and ∆n can be any integer.
The transition probability is actually the product of population difference between two quantum
states and the square of transition moment integral. 4 The relative population of states is specified
by Criteria (2) and (3) and it follows Boltzmann statistics

N j g j −∆E/kB T
= e (2)
Ni gi

in which g is the degeneracy of the state. At room temperature (about 300K), the ratio in the
exponent is approximately 1 for rotational transition, and much less than 1 for vibrational and
electronic transitions. This means that at room temperature most molecules are rotationally excited,
while being in vibrational and electronic ground states.
Having laid down the foundation for spectroscopy, we are now ready for detailed discussions
on the two quantum mechanical models. The next section is divided into two parts, each devoted
to a quantum mechanical model. The discourse on the theory of the models loosely follow the
progression from the spherical cow form, to revised model, to reconsidering the revised model for
spectroscopic purposes and methods for data analyses.

Diatomic molecules as oscillators

Molecular vibration is a process by which the distance between atoms varies periodically, or the
bond between atoms “stretches", while the translational and rotational motion of the molecule as a
whole remain relatively constant. Such nuclear motion can be modeled using a classical harmonic
oscillator, in which the masses (atoms) are expressed as reduced mass µ and they are connected
by a spring that obeys Hooke’s law ( f = −k(r − re )), where k the force constant corresponds
to bond strength in a molecule and r − re is the displacement of the spring from its undistorted
length re . This is a reasonable approximation as the net force acting on the atoms is zero at
equilibrium bond length. However unlike a classical spring, a chemical bond is not a concrete,
tangible object but a region of high electron density from overlapping atomic orbitals – it can be

6
stretched infinitely as there are no real tethers to the bonded atoms. Since the force acting on the
atoms is not directly proportional to the displacement bond length, vibrating diatomic molecules
do not exhibit harmonic behavior; rather they are anharmonic oscillators.
Recall that force is the negative derivative of potential energy (V ), integrating force with respect
to position r then yields an expression for V :

Z
V (r) = − f (r)dr + constant (3)

The harmonic oscillator potential is parabolic due to its defining function 21 k(r − re )2 . The
difference between a classical harmonic oscillator and a quantum mechanical harmonic oscillator
is that the energy in the latter is quantized, which is

 
1
Ev = hν v + v = 0, 1, 2, ... (4)
2

The quantization of energy manifests itself in intervals of hν (i.e. the spacing between each energy
level is hν), and it arises from the Schrödinger equation of one-dimensional harmonic oscillator

h̄2 d 2 ψ 1
 
2
Ĥψ = − + k(r − re ) ψ = Ev ψ (5)
2µ dr2 2

To obtain a potential energy function that accounts for anharmonicity in a vibrating molecule,
consider a Taylor expansion of V (r) about r = re :

1 d 2V 1 d 3V
     
dV 2
V (r) = V (re ) + (r − re ) + (r − re ) + (r − re )3 + ...
dr r=re 2! dr2 r=re 3! dr3 r=re
1 1 1
= kr2 + γ3 r3 + γ4 r4 ... (6)
2 6 24
   
d 2V d jV
where k = dr2
is the force constant and γ j = dr j
is the j-th anharmonic term. When
r=re r=re
the anharmonic terms are incorporated into the Hamiltonian operator for vibrational motion of a

7
diatomic molecule, the eigenvalue Ev after solving the Schrödinger equation is

1 2
   
1
Ev = ω e v + − ωe χe v + v = 0, 1, 2, ... (7)
2 2

where v is the vibrational quantum number, ωe is the fundamental vibrational frequency in wavenum-
ber form, and χe is the anharmonicity constant, and the subscript e denotes equilibrium state. (For
ease of vacillating between theoretical and experimental discussion, all equations henceforth are
expressed in wavenumber unit ( cm−1 ) unless noted otherwise.) The consequence of anharmonic-
ity is that the spacing between each energy level is no longer equal to hν; the loss of linearity in
Eq. (7) causes separation between energy levels to decrease as v increases.
The force constant ke defines the curvature of the potential curve at the minimum distance re
and it is the second partial derivative of potential energy: 5

∂ 2V
 
ke = = µ(2πcωe )2 (8)
∂ r2 re

From Eq. (8) above it can be seen that ke is directly proportional to ωe . A large ωe indicates a
strong stiff bond, which in turn implies small deviations from equilibrium (i.e. oscillation about
re is small, not anharmonic), therefore it can be deduced that ωe and the anharmonicity constant
χe are inversely related. The exception to this is hydrogen, whose tiny mass and relative lack of
electron density lead to anharmonic behavior even though the bond in a hydrogen bonded diatom is
generally stiff. The direct proportional relationship between k and ωe suggests that different types
of bond will result in different spacings between energy levels, as ωe dictates the magnitude of Ev .
It should be noted that while Eq. (6) accounts for anharmonicity, it does not predict bond break-
ing behavior as internuclear distance increases. A better potential model for vibrational motion in
diatomic molecules is the Morse potential, which includes both anharmonicity and bond dissocia-
tion effects. The Morse potential function is given by

 2
−β (r−re )
V (r) = De 1 − e (9)

8
where β originates from k = 2De β 2 . A typical Morse curve is shown in Figure 2; note that there
is a potential energy curve corresponding to each electronic state. The Born-Oppenheimer ap-
proximation separates electronic motions from nuclear motions, thereby reducing the Schrödinger
equations for electronic energies in molecules to a problem that can be solved at fixed nuclear
positions. Thus for each position of the nuclei, there correspond a Hamiltonian operator (and
Schrödinger equation) and an eigenvalue of electronic energy. The collection of these eigenvalues
is plotted against different nuclei positions, generating a molecular potential curve as a function of
internuclear distance. 6 Since electrons are promoted to from highest occupied molecular orbitals
(HOMOs) to lowest unoccupied molecular orbitals (LUMOs) upon excitation, the wavefunctions
describing excited states are different from those describing the previous electronic state, hence
there is a potential curve for each electronic state.
The position of the potential curve is governed by the Franck-Condon principle, which is also
built on the Born-Oppenheimer approximation. Because nuclei are monolithic relative to electrons,
redistribution of electrons due to electronic transition upon photon absorption proceeds so rapidly
that the internuclear distance appears to be frozen during transition; 8 graphically this is represented
by the vertical transition line in Figure 2. In light of the relative speed at which molecules rearrange
themselves after excitation, the Franck-Condon principle states that “the excited state is formed
with the same geometry as that of the ground state from which it derived." 7 Another factor that
determines the position of the excited state potential is the equilibrium bond length at each state.

Applying to UV-visible spectroscopy of iodine

The objective of this experiment was to characterize the excited state of iodine, assuming that
the molecular potential in gaseous I2 takes the form of Morse curve. Iodine Figure 3 shows a
Morse potential curve of iodine at the ground state X and the first excited state B, intersected by a
line A that signifies the energy of the ground state antibonding orbital. By convention, the lower
electronic state and upper electronic state are symbolized by double prime and single prime respec-
tively; consequently the parameters associated with the ground state and the first excited state are

9
Figure 1: Energy gap in each internal energy component at room temperature. Reproduced from
Reference 2.

Figure 2: A typical Morse curve reproduced from Reference 7.

10
Figure 3: Potential energy diagram of molecular iodine. (Originally from Rodney J. Sime, Physical
Chemistry, Saunders College Publishing, Philadelphia, 1990. P. 661)

11
each denoted by double prime and single prime quantities. The excited state parameters to be de-
termined from experimental spectrum are ωe0 , χe0 , v0d , D00 , and D0e . ωe0 and χe0 are defined previously
as the fundamental vibrational frequency and the first-order anharmonicity correction constant. D0e
is the depth of the excited state potential well, and it is measured from the bottom of the potential
to the top line where the series of vibrational levels converges due to anharmonicity effect. This
convergence corresponds to bond dissociation, and the quantum number v0 at dissociation is v0d .
The dissociation energy is actually D00 , the energy measured from v0 = 0 to v0d , instead of D0e be-
cause a molecule cannot have less than zero-point vibrational energy. D00 and D0e differ by one Ev ,
which is defined in Eq. (7).
The main method of spectral data analysis for this experiment is the construction of Birge-
Sponer plot. A Birge-Sponer plot plots the transition frequency between adjacent vibrational levels
against vibrational quantum numbers. The transition frequency (∆E in the unit of cm−1 ) between
adjacent vibrational levels v0 and v0 + 1 in the upper electronic state is simply

∆E(v00 → v0 ) − ∆E(v00 → v0 + 1) = ∆Eupper (v0 ) (10)

Since the majority of iodine molecules are in the ground vibrational state at room temperature, v00
is presumed to be 0. Then the energy difference can also be expressed as the difference between
the energy measured from the zero-point energy at v00 = 0 to v0 and the energy measured from
v00 = 0 to v0 + 1 (cf. Figure 3). The transitions from v00 = 0 to v0 encompass the electronic energy
of the excited state iodine molecule Eel0 = T , with Eel for the ground state arbitrarily set to zero.

12
The expression for ∆Eupper is

∆Eupper (v0 ) = (Eel0 − Eel0 ) + (Ev+1


0
− Ev0 )
1 2 1 2
       
0 0 1 0 0 0 0 0 1 0 0 0
= (T − T ) + ωe v + 1 + − ωe χe v + 1 + − ωe v + − ωe χe v +
2 2 2 2
= ωe0 − 2ωe0 χe0 (v0 + 1)

∆Eupper (v0 ) = ωe0 − 2ωe0 χe0 −2ωe0 χe0 v0 (11)


| {z } | {z }
intercept slope

Therefore the linear regression for the Birge-Sponer plot is Eq. (11) with a slope of −2ωe0 χe0 and
an intercept of ωe0 − 2ωe0 χe0 . In practice, each ∆Eupper is the difference between the wavelength
corresponding to adjacent spectral lines v0 . The collection of peaks are entered into a table called
Deslandres table for data analysis. 9
The dissociation energies D00 and D0e can also be determined from the Birge-Sponer plot. From
Figure 3, D00 is the sum of all spacings between adjacent vibrational levels from v0 = 0 to v0d .
Mathematically, this is

D00 = ∆Eupper (v0 = 0) + ∆Eupper (v0 = 1) + ∆Eupper (v0 = 2) + ... + ∆Eupper (v0 = v0d ) (12)

Assuming that the partitions within Riemann sums are infinitesimal (i.e. the energy gaps are very
small), the summation can be approximated by numerical integration:

Z v0
d
D00 = ∆Eupper (v0 )dv0 (13)
0

The value of D00 is therefore the area of the triangle under the Birge-Sponer plot evaluated from
v0 = 0 to v0 = v0d . v0d is the x-intercept because by definition v0d is the vibration quantum number at
which the energy spacing between adjacent vibrational levels converge, hence ∆Eupper (v0 ) = 0 at
convergence. The value of D0e can be determined by adding the spacing between the bottom of the
potential and the zero-point vibrational energy, Ev0 = E0 (cf. Eq. (7)), to D00 .

13
Additional calculations The Birge-Sponer method can only extract few details from the excited
state potential well of iodine. If the objective of this experiment were to expand to include ground
state potential characterization as well, an emission spectrum would be necessary. The multiple
linear regression method is able to compute both excited state and ground state quantities without
necessitating fluorescence spectroscopy. It should be noted that without resolving rotational fine
structures, the values regarding bond lengths cannot be determined. For this reason literature values
of re and re0 are used in constructing Morse potential curves for the ground state and first excited
state.

Diatomic molecules as rotators

In gas phase, molecules are far apart that their individual molecular motions can occur without
inducing interactions with the neighboring molecules. Therefore the gas phase infrared spectra of
simple linear diatomic molecules such as HCl and DCl can be explained by quantum mechanical
models.
The photons in the infrared region possess energy sufficient to induce vibrational excitations
in molecules, which are represented as absorption peaks in a gas phase infrared spectrum. If we
imagine a linear diatomic molecule as a harmonic oscillator, we can see that there is only one way
the bond length in a diatomic molecule can vary: stretching. Hence there is only one vibrational
mode available in diatomic molecules, and consequently only one absorption peak is observed in a
vibrational spectrum. However in each vibrational state there are rotational sub-states, just like in
each electronic state there are vibrational sub-levels; which is why rotational fine structures may
be observed in vibrational bands.
The Schrödinger equation of a rigid rotator is

L̂2
Ĥψ = − ψ = EJ ψ (14)
2I

where L̂ is the angular momentum operator, and I is the moment of inertia. [energy spacing] The

14
energy of a rigid rotator is given as
EJ = BJ(J + 1) (15)

where J is the rotational quantum number, and B, the rotational constant, is defined below.

h2
B= (16)
8π 2 cµr2

The term µr2 is the moment of inertia I, in which µ is the reduced mass and r is the interatomic
distance. A more accurate definition of moment of inertia is ∑i mi z2i , where mi is the mass of the
ith atom and zi is the distance from the center of mass of the molecule.
While the rigid rotor model is relatively accurate, it does not embody all aspects of a linear
rotating diatomic molecule; some perturbation is needed to correct for some oversights. Firstly,
the bond between two atoms is not fixed as in a classical rigid rotor; the electron clouds from each
atoms overlap and constitute a bond. When a linear diatomic molecule rotates, the entire molecule
is subjected to centrifugal force and the bond is stretched (i.e. zi is not constant as presumed).The
energy of a rigid rotor model that has taken the centrifugal distortion of the bond into account is
then

EJ = BJ(J + 1) − DJ 2 (J + 1)2 (17)

where D is the centrifugal distortion constant. Like Ev with anharmonicity correction, the loss
of linearity causes the spacings between adjacent rotational levels to be uneven. [more] [diagram]
The second aspect of a linear rotating diatomic molecule that deviates from a classical rigid
rotator is the zero-point motion in the ground vibrational state. The zero-point vibrational mo-
tion does not invalidate the rigid rotor assumption that the internuclear distance is constant, since
vibration is much faster than rotation that the atomic positions in rigid rotor can be given by aver-
age atomic positions in vibrational motion. However, the average atomic positions from vibration
depend on the vibrational states that the molecule is in. Consequently the correct energy equa-
tion would require a perturbation term involving a constant Bv corresponding to each vibrational

15
state. Bv can be expressed in terms of a rotational constant Be and a vibrational-rotational coupling
constant α:
 
1
Bv = Be − α v + (18)
2

where Be is the rotational constant corresponding to a system in which the atoms are fixed in their
equilibrium positions at equilibrium internuclear distance re . Rearranging Eq. (18) engenders an
expression for α:
2
α= (Be − Bv ) (19)
2v + 1

The complete model for the energy levels of a linear diatomic molecule in its ground electronic
state is:
1 2
   
1
Ev,J = ωe v + − ωe χe v + + Bv J(J + 1) − Dv J 2 (J + 1)2 (20)
2 2

Although Dv , the centrifugal distortion constant, may show a vibrational state dependence, this
effect is small and it is assumed to be the same in all vibrational states in this experiment.

Applying to infrared spectroscopy of hydrogen chloride

Evaluation of the anharmonic oscillating rotator model developed in the previous section requires
vibrational spectroscopy with rotational resolution. At room temperature, most molecules are in
the ground vibrational and electronic state, but the smaller increment between adjacent rotational
levels results in molecules distributed over a variety of rotational states. Assuming that Be =
10 cm−1 , ṽ0 = 2880 cm−1 and D = 0, at thermal equilibrium the energy difference between the
ground rotational state J = 0 and upper rotational state J is

∆EJ=0→J = v˜0 + Be J(J + 1) (21)

The relative populations of rotational levels at room temperature (298K) is then

NJ
= (2J + 1)e−∆EJ=0→J /kB T (22)
N0

16
Figure 4: Rotational-vibrational transitions. Image from <http://chemwiki.ucdavis.
edu/Physical_Chemistry/Spectroscopy/Rotational_Spectroscopy_of_
Diatomic_Molecules>.

17
Figure 5: Rotational transitions diagram from <http://www.ucl.ac.uk/~ucapphj/
lecture_26.htm>. Note that ν̃0 (or ω0 in the figure) is missing because ∆J = 0 transitions
are forbidden.

18
where Boltzmann constant kB in spectroscopic unit is 0.695 cm−1 · K−1 . From the plot of distribu-
tion of rotational levels shown in Figure 6, it appears that the majority of gaseous HCl molecules
reside in the rotational states J = 1 through J = 5, with J = 3 being the most populated. The density
of population of a rotational state is directly correlated to the absorption intensity in a spectrum.
Since infrared radiation is not energetic enough to induce electronic excitation, the single prime
and double prime notation is dropped for vibrational states. Instead the single and double prime
notation is implemented to differentiate rotational states nested within the lower and upper vi-
brational states, which happen to be v = 0 and v = 1 at room temperature. The transitions in
vibrational-rotational spectroscopy are depicted in Figure 4. The vibrational transition frequency
for v = 0 → v = 1 is
ν̃0 = E1vib − E0vib = ωe − 2ωe χe (23)

The energy difference between a rotational state in lower vibrational state and a rotational state in
upper vibrational state is then

∆E = ν̃0 + B1 J 0 (J 0 + 1) − DJ 02 (J 0 + 1)2 − B0 J 00 (J 00 + 1) − DJ 002 (J 00 + 1)2 (24)

Let the rotational quantum number in the lower rotational state J 00 be J, then the rotational state
in the upper vibrational state J 0 becomes J ± 1. There are two types of transitions possible as per
the selection rule: ∆J = +1 and ∆J = −1, and they are illustrated in Figure 5. From Eq. (24), it
can be seen the transitions involving ∆J = +1 occur at frequencies higher than the fundamental
vibrational frequency ν̃0 , while transitions involving ∆J = −1 occur at frequencies lower than
ν̃0 . These two kinds of transitions are reflected in a vibrational-rotational spectrum as two series
of peaks on either side of ν̃0 ; the peaks emerging from ∆J = +1 and ∆J = −1 transitions are
designated the R branch and the P branch respectively. The frequencies relating to the R and P

19
branches are

ν̃R = ν̃0 + B1 (J + 1)(J + 2) − B0 J(J + 1) − D (J + 1)2 (J + 2)2 − J 2 (J + 1)2


 
(25)

ν̃P = ν̃0 + B1 J(J − 1) − B0 J(J + 1) − D J 2 (J − 1)2 − J 2 (J + 1)2


 
(26)

For the R branch, J can have values of integers starting from 0, therefore the lowest R branch
transition frequency for J = 0 is ν̃0 + 2B1 − 4D. There is no corresponding transition in the P
branch because J must be greater than 0 in the upper state. The first P branch transition frequency
for J = 1 is ν̃0 − 2B0 + 4D. Substituting α = B0 − B1 (derived from Eq. (16)), Eq. (3) and Eq. (4)
become

ν̃R = ν̃0 + 2Be − 3α + (2Be − 4α)J − αJ 2 − 4D(J + 1)3 (27)

ν̃P = ν̃0 − (2Be − 2α)J − αJ 2 + 4DJ 3 (28)

From these two equations we know that the vibrational band diverges, one above and one below the
band origin ν̃0 . The spacing between adjacent lines is approximately 2B, then gradually changes
as the effect of vibrational-rotational coupling (α) and centrifugal distortion (D) become more
apparent. R and P branches are separated by width about 4B which is known is the band gap. The
two equations for R and P branches can be combined into one if the lines are labeled with a running
index, m, where m = J + 1 for the R branch and m = −J for the P branch. The combined equation
can be simplified to
ν̃ = ν̃0 + (2Be − 2α)m − αm2 − 4Dm3 (29)

In data analysis, a third order polynomial regression line is used to fit the plot of absorp-
tion wavenumbers versus running index m. The values of the centrifugal distortion constant D,
vibrational-coupling constant α, rotational constant at equilibrium Be and fundamental vibrational
frequency ṽ0 are facilely extracted from the coefficients of the best fit line equation. The rotational

20
constants B0 and B1 are calculated by solving the linear system


α = B0 − B1 

(30)
1 1 
Be − α = B0 + B1 
2 2

And bond lengths r can be solved using the rearranged expression of Be in Eq. (16):

s
h
r= (31)
8π 2 cµB

Goal

In summary, the goal of this two-part experiment is to acquire information about diatomic molecules
relayed from different regions of electromagnetic radiation and their eponymous spectroscopic
methods listed in Table I, thereby assessing the validity of the anharmonic oscillator model of
iodine and anharmonic oscillating rotator model of hydrogen chloride and its isotopes. The infor-
mation relayed from UV-visible electronic-vibrational spectroscopy is the bond energy and bond
strength of iodine in its excited state, and the constants to be extracted from spectral data analysis
are ωe0 , χe0 , v0d , D00 , and D0e . The information communicated from infrared vibrational-rotational
spectroscopy is bond strength and bond length, and the constants to be determined are ṽ0 , D, α,
Be , B0 , B1 , re , r0 , r1 and I.
By developing more sophisticated models of molecular motions, we can better explain the
information gained from spectroscopy that classical mechanics and “spherical cow" quantum me-
chanics could not.

21
Methods

Electronic spectrum of iodine

When a diatomic molecule such as iodine absorbs a photon of energy hν, not only does it cause
a change in electronic state, but it is coupled with vibrational transition as well. Since vibrational
fine structures are observed in the electronic spectrum, the spectrum generated from UV-visible
spectroscopy is also called vibronic spectrum, a portmanteau of vibrational and electronic transi-
tions. The advantage of vibronic spectroscopy to studying iodine is that it circumvents the problem
of infrared inactivity of I2 . Naturally rotational excitation would also be induced along with vi-
bronic transitions upon ultraviolet-visible photo-radiation , but the rotational fine structures were
not visible at the resolution set for this experiment.
The vibronic spectrum of iodine vapor was taken on a Varian Cary 2200 high resolution UV-
visible spectrophotometer. A rapid scan of I2 in the range of 500-650 nm was first obtained by
setting the scan rate to 1.0 nm/s and chart display to 10 nm/cm. This low resolution scan was
a overview of the electronic spectrum without resolving the vibrational fine structures. A high
resolution scan was next collected by setting the scan rate to 0.2 nm/s and chart display to 1 nm/cm.
After printing the spectra, absorption peaks were assigned with the literature values reproduced in
Table II serving as guiding points.

Table II: Numbering of band heads for iodine

v0 v00 wavelength (nm)


27 0 541.20
28 0 539.00
29 0 536.90
18 1 571.60
19 1 568.60
20 1 565.60
13 2 595.70
14 2 592.00
15 2 588.50

22
Vibrational-Rotational spectra of HCl and DCl

An evacuated 10 cm gas cell was placed in the sample holder inside Brüker Tensor 27 Fourier trans-
form infrared spectrometer for a high resolution (0.5 cm−1 ) background spectrum. Afterwards the
gas cell was removed from the spectrometer. Gaseous hydrogen chloride and deuterium chloride
mixture was introduced into to the gas cell via a gas sampling train system, whose apparatus as-
sembly is shown in Figure 7. The filled gas cell was loaded onto the spectrometer as swiftly as
possible to avoid water vapor and carbon dioxide contamination from human contact. The spectra
of the sample were then collected and printed for data analyses. The HCl spectrum was expected
in the range of 3100 − 2500 cm−1 and the DCl spectrum in the range of 2300 − 1900 cm−1 .
Each line in the spectra was composed of a closely spaced doublet. The higher intensity bands
are ascribed to the more abundant isotope of chlorine, 35Cl, where as the shorter peaks are at-
tributed to the less abundant isotope 37Cl. Each line was labeled with a running index m for each
isotope. For clarity, in additional to line index, the spectral lines were also categorized as R and P
transitions. The series of lines occurring at frequencies higher than the band gap in the middle of
the spectra were identified as the R branch, and those occurring at frequencies lower than the band
gap were labeled as the P branch. The absorption wavenumbers were entered onto a spreadsheet
program for third-order polynomial fit. The constants extracted from the fit lines were then used to
calculate bond lengths and rotational constants on a computer algebra system program.

Results

Ultraviolet-Visible Spectroscopy of I2

A low resolution spectrum of iodine is shown in Figure 8. The high resolution spectrum from
which data were collected was too large to be captured by resources available and thus is not
shown here. The absorption spectrum is actually comprised of overlapping bands - a cold band
and several hot bands. A cold band consists of absorption peaks due to transition from the lowest

23
Figure 6: Relative populations of rotational levels in rigid rotator HCl at room temperature (298K).

Figure 7: Gas sampling train system. Image taken from <http://www.epa.gov/ttn/emc/


ftir/reports/r03.html>.

24
vibrational state in the ground electronic state v00 = 0 to a vibrational state in the upper electronic
state v0 . Hot bands are composed of peaks caused by transitions from various v00 to v0 . These
span from 650 nm up to 550 nm and they overlap with the cold band in the 575 nm region. 10 For
this reason v00 = 0 → v0 transition peaks were identified and recorded starting from v0 = 18, the
vibrational state at which absorption peak began to emerge above the overwhelming signals from
the hot bands. The first few rows of the Deslandres table are shown in Table III; the cells contain
absorption wavelengths corresponding to each vibrational quantum number in the upper electronic
state converted to wavenumber units, and the italicized numbers are ∆Eupper (v0 → v0 + 1). Few
peaks were identified from the hot band regions: three from the v00 = 2 → v0 transitions, and 9 from
the v00 = 1 → v0 transitions. Although not shown in Table III, peaks corresponding to v00 = 0 → v0
transitions were successfully assigned for quantum number v0 = 18 up to v0 = 52, with two outliers
occurring at v0 = 48 and v0 = 49 omitted. The Birge-Sponer plot for vibrational transition from
v” = 0 to v0 is shown in Figure 9. The linear regression line is y = −1.974x + 130.54, from which
the coefficients −1.974 and 130.54 are −2ωe0 χe0 and ωe0 − 2ωe0 χe0 respectively. ωe0 and ωe0 χe0 were
calculated to be 132.514 cm−1 and 0.987 cm−1 after algebraic manipulation. v0d , the x-intercept,
was found to be 66 after rounding. D00 was solved by integrating ∆Eupper with respect to v0 from
0 to v0 = 66 and was equal to 4316.27 cm−1 . Adding Ev0 (cf. Eq. (7)) to D00 gave a value of
4349.83 cm−1 for D0e . Using the reduced mass of iodine (1.053 × 10−25 kg), the excited state force
constant ke0 was computed to be 65.61 N · m−1 . The excited state quantities extracted from the
Birge-Sponer plot were summarized in Table IV. The experimentally determined ωe0 and ωe0 χe0 were
plugged in Eq. (7) to generate a plot of vibrational energy in the upper electronic state displayed
in Figure 10. Note that the maximum vibrational energy only hovers above 4000 cm−1 without
electronic contribution.
The Morse potential function of the upper electronic state is

 r 2
k0
− (r−re )
2D0e
V 0 (r) = T + D0e 1 − e  (32)

25
Figure 8: Rapid scan spectrum of I2 .

Figure 9: Birge-Sponer plot of I2 .

26
Figure 10: Vibrational energy in the upper electronic state as a function of vibrational quantum
number v0 using experimentally determined constants.

Figure 11: Vibronic energy a a function of vibrational quantum number v0 plotted from experimen-
tal data.

27
Table III: First few entries of the Deslandres Table

v0 v00 = 0 v00 = 1 v00 = 2 Average Difference


13 16778.52
104.81 104.81
14 16883.34
106.13 106.13
15 16989.47

18 17702.25 17488.63
94.51 92.24 93.38
19 17796.76 17580.87
92.33 93.22 92.77
20 17889.09 17674.09
90.06 87.90 88.98
21 17979.14 17761.99
87.70 89.42 88.56
22 18066.85 17851.40
85.27 89.68 87.47
23 18152.11 17941.08
82.42 88.31 85.36
24 18234.53 18029.39
80.49 83.27 81.88
25 18315.02 18112.66
80.86 79.08 79.97
26 18395.88 18191.74
75.09 75.09

A full characterization of the excited state potential well can be achieved by inclusion of T and
re . T was extrapolated from the regression line for the plot of vibronic energy in Figure 11 and
was found to be 15679 cm−1 . re was solved by setting the left hand side of Eq. (32) to be the
frequency at which the maximum absorption intensity occurred. [Franck-Condon *cite iodine
maple*] At maximum absorbance, the internuclear distance r at the excited state becomes the
ground state equilibrium bond length. The most intense peak was that corresponding to v0 = 34 and
∆E(v0 = 34) was 18959.14 cm−1 . After solving the quadratic equation, re was found to be 2.98Å.
Having complete parameters of the excited state potential well, the Morse potential curve for the
upper electronic state and the correlating vibrational states were plotted and shown in Figure 12.
Data are meaningless without context. Comparison with the ground state quantities is the

28
Table IV: I2 Excited state parameters as determined by the Birge-Sponer method

ωe0 132.514 cm−1


ωe0 χe0 0.987 cm−1
D00 4316.268 cm−1
D0e 4349.827 cm−1
v0d 66
k0 65.61 N · m−1

best way to gain insight to the experimentally determined excited state data. As discussion in a
previous section, ground state information cannot be attained via the Birge-Sponer method without
procuring an emission spectrum, but it can be done with a multiple linear regression analysis. Using
the procedure outlined by Cooper, 11 the excited state and ground state spectroscopic constants
computed are summarized in Table V. A quick comparison of Table V and Table IV reveals that the
excited state values acquired from the Birge-Sponer method and multiple linear regression method
are primarily similar, so the excited state data do have an internal consistency to an extent. The
juxtaposition of the excited state properties with ground state properties in Table V perspicuously
demonstrates the energetic stability of the ground state. The anharmonicity is smaller in the ground
state, dissociation energies are substantially larger than those in the excited state, the vibrational
quantum number at dissociation is higher, the force constant k00 is almost a 100 units bigger than
k0 at excited state; all these values point to the superior bond strength in the ground state relative
to the excited state. A graphical representation of the energetic difference is in Figure 13, which
shows the Morse potential curve of both ground state and excited state iodine, constructed using
the data acquired from the multiple linear regression method.

Table V: Constants extracted from multiple linear regression method

Ground state Excited State


ωe00 208.06 cm−1 ωe0 132.59 cm−1
ωe00 χe00 1.0146 cm−1 ωe0 χe0 1.0220 cm−1
D0 00 12297.59 cm−1 D00 4234.54 cm−1
D00e 12401.37 cm−1 D0e 4301.09 cm−1
v00d 102 v0d 64
k00 161.74 N · m−1 k0 65.69 N · m−1
Eel 15703.28 cm−1

29
Infrared spectroscopy of HCl and its isotopes

The vibrational-rotational spectra of HCl and DCl are shown in Figure 14a and Figure 14b. In
HCl, the lines from the region of 2900 − 3100 cm−1 are of the R branch, while those from the
< 2900 − 2600 cm−1 region are of the P branch. In DCl, the absorption lines in the range of
∼ 2100 − 2250 cm−1 belong to the R branch, and those in the range of < 2100 − 1900 cm−1 are of
the P branch. Within the spectra of HCl and DCl, the spectral lines attributed to the molecules with
35Cl are differentiated from those of 37Cl by frequencies and line intensity. The absorption peaks
of H35 Cl and D35 Cl occur at the higher frequencies out of the doublets, and they show more intense
absorbance due to higher relative abundance. The peaks were assigned with a line index m then cat-
aloged into 4 sets of data, each associated with H35 Cl, H37 Cl, D35 Cl, and D37 Cl respectively. Each
set of data was plotted against running index m and fitted with a third-order polynomial regression
line. Each polynomial regression has a form coinciding with Eq. (6), so ν̃0 and α can be directly
extracted from the fit lines, and D and Be can be determined with minimal calculations. The con-
stants educed from the fitting procedure are compiled in Table VI. It appears that the isotope effect
of chlorine is not as prominent as that of deuterium, seeing that the constant values do not vary by
much within the H35 Cl-H37 Cl and D35 Cl-D37 Cl pairs, while significant differences are observed
when comparing across hydrogen chloride and deuterium chloride isotopes. Given that chlorine is
already the heavier constituent of the diatom, having two additional neutrons on the chlorine atom
does not alter the weight distribution in the molecule to a noticeable extent. Replacing hydrogen
with deuterium on the other hand produces a significant reduction in the constant values except D.
Since the mass of hydrogen is very small, doubling its mass upon isotopic substitution causes a
notable increase in reduced mass and subsequently moment of inertia. A diminished magnitude of
ν̃0 is found in the deuterated isotopes; this is most likely due to the fact that the heavier deuterium
impedes the speed of bond stretching, hence the fundamental vibrational frequency is lower in the
DCl isotopes. Similarly, because rotational constant Be is inversely related to moment of inertia,
a doubling of reduced mass upon deuteration causes Be to be halved. The centrifugal distortion
constant D sees a trend reverse of that in Be and ν̃0 . Although there is no mathematical quan-

30
tification of D in this paper, the increase of centrifugal distortion upon isotopic substitution can
be rationalized by a spring-rotator model. Under the Born-Oppenheimer approximation, the force
constant k is postulated to be independent of isotope substitution. One explanation proffered by
Molecular Orbital Theory is that since bonds are formed by electron density overlap, bond order
is independent of isotopic effect because isotopic substitution changes not the number of electrons
but that of neutrons. So assuming that the force constant k (thereby bond stiffness) in the four iso-
topes are identical, the bond displacement would be greater in the molecules with heavier atoms on
either ends. Consider four springs of equal stiffness, each with two spheres attached to its ends and
has the same asymmetric weight distribution found in the four isotopes. During rotation the spring
distortions experienced in the models with the same lighter atoms (i.e. H group and D group) are
on the same order of magnitude as the percent difference in moment of inertia between each pair is
less than 0.02%. Spring distortions are more extreme when the lighter end of the spring is replaced
by a heavier mass; the gravitational force is stronger when mass is increased, so the spring is com-
pressed more during rotation. This classical analogy accounts for the increase in D as a function
of reduced mass.
The bond lengths and rotational constants at each vibrational state and at equilibrium were
calculated using the values from Table VI. The results are listed in Table VII. Bar charts, shown
in Figure 16, were created as visualization tool to better elucidate the isotope effects on bond
lengths and rotational constants. Figure 16a compares re , r0 , and r1 among the four isotopes.
Across the three groups, there is a clear trend that bond lengths increase as the energy of the
state increases. re refers to the bond length in a hypothetical state in which atoms are frozen in
their equilibrium positions, so by definition re is in fact the bond length of a rigid rotator. In the
ground vibrational state the atoms are unfrozen from their equilibrium positions and are allowed to
vibrate, so naturally r0 are longer than re . At excited vibrational state, bonds are vibrating at higher
frequencies and are more energetic, therefore the values of r1 are even larger than those of r0 . It
needs to be clarified that even though the bar charts in Figure 16a explicate the subtle differences
among the isotopes, these within-group differences are in fact of no physical significance and

31
are only visible due to the scaling. The standard deviations for each bond length group are all
less than 0.004Å, proving that there is no significant difference within a bond length group, and
correspondingly no isotope effect detected for bond length. The independence of bond length from
isotopic substitution in a given state further corroborates the validity of the Born-Oppenheimer
approximation.
Unlike bond lengths, isotopic effect is very much present in rotational constants. The three
rotational constant groups show very discernible disparities in bar height in Figure 16b. A quick
standard deviation analysis indeed reveals that the differences within groups are significant, with
SD ≈ 3Å in each group. Based on the bar chart, the effect is only pronounced for isotopic sub-
stitution of hydrogen with deuterium, and no perceptible effect is found for the chlorine isotopes.
In all three states the rotational constants of H35 Cl and H37 Cl are approximately twice as large
as the rotational constants of D35 Cl and D37 Cl. This reflects the doubling of reduced mass in the
denominator of B. Across groups, the rotational constants decrease slightly as the energy of the
given state increases. Because bond length increases when energy is increased, and that rotational
constant B is inversely proportional to bond length, rotational constants are expected to decrease in
magnitude with increasing energy. However seeing that the bond length difference between each
state is on the scale of picometer, the change in bond length is hardly conveyed in the rotational
constant value.
To sum up, bond length is independent of isotopic effect but correlated with energy of a given
state, consistent with the behavior of a kinetic property. Conversely rotational constant is inde-
pendent of state but governed by isotope effect; this is a reasonable conclusion as it justifies the
eponym rotational constant, for B is an inherent property of a molecule.

Table VI: Constants extracted from polynomial fitting procedure

D( cm−1 ) α( cm−1 ) Be cm−1 ṽ0 ( cm−1 )


H35 Cl 0.000625 0.304 10.615 2886
H37 Cl 0.000675 0.3044 10.6134 2883.9
D35 Cl 0.0015 0.1122 5.4482 2091.1
D37 Cl 0.0015 0.1117 5.4322 2088.1

32
Table VII: Results

re (Å) r0 (Å) r1 (Å) Be ( cm−1 ) B0 ( cm−1 ) B1 ( cm−1 ) µ( kg) I( kg · m−2 )


H35 Cl 1.2733 1.2825 1.3015 10.6150 10.4630 10.1590 1.6267 × 10−27 2.6372 × 10−47
H37 Cl 1.2724 1.2816 1.3007 10.6134 10.4612 10.1568 1.6292 × 10−27 2.6376 × 10−47
D35 Cl 1.2747 1.2813 1.2948 5.4482 5.3921 5.2799 3.1624 × 10−27 5.1381 × 10−47
D37 Cl 1.2746 1.2813 1.2948 5.4322 5.3764 5.2647 3.1718 × 10−27 5.1532 × 10−47

Discussion

Anharmonic oscillator - I2

Calculations of spectroscopic constants would have been impossible if the quantum mechanical
model of molecular iodine were harmonic oscillator instead of anharmonic oscillator, considering
that it is anharmonicity that causes energy levels to be unevenly spaced and eventually converging
at very large v. The experimental excited state quantities gathered from Birge-Sponer method and
multiple linear regression method are in good agreement with each other, and comparison with
the ground state parameters extracted from multiple linear regression does confirm that the excited
state potential is more shallow and offset from the ground state potential (cf. Figure 13). The
comparability between the two sets of results proves that the experimentally determined excited
state parameters are precise (or at least in proximity to precision), but are they accurate? Table VIII
lists the experimentally determined excited state quantities of iodine along side literature values
reproduced from Reference 10. It appears that the experimental values tend to overestimate the
vibrational frequency, anharmonicity and force constant at the excited state, and consequently
underestimate dissociation energies and vibrational quantum state in which dissociation occurs.
Birge-Sponer method fares slightly better than multiple linear regression analysis in that its extent
of misestimation is smaller. The discrepancies between the experimental values and the literature
values could be due to instrumental error. The UV-visible spectrometer Varian Cary 2200 was not
only dated (hence less reliable than newer instruments), the mechanical arm holding the pen at the
recorder writing panel was broken. To ensure ink head was in continuous contact with the sheet
during printing, a person had to manually hold the sheet at an upward angle. Since the printing of

33
high resolution spectra was quite a lengthy process, the sheet feeding might not be steady and the
peak positions might have shifted in printing, thereby producing a misaligned spectrum. Another
probable reason for experimental data deviation from reported values is analytical error. Literature
values are often obtained from the most comprehensive, accurate computation possible; in this
case, the reported values are calculated using an anharmonic oscillator potential that includes many
more high order correction terms. Since only one anharmonic correction term is considered in this
experiment, the vibrational frequency (ωe0 ) contribution to Ev (v0 ) may have been exaggerated, and
this error becomes propagated in subsequent calculations.

Table VIII: Iodine excited state parameters

Experimental Literature 10
Birge-Sponer Multiple Linear Regression
ωe0 132.51 cm−1 132.59 cm−1 125.67 cm−1
ωe0 χe0 0.987 cm−1 1.022 cm−1 0.750 cm−1
D00 4316.3 cm−1 4234.5 cm−1 4318.6 cm−1
D0e 4349.8 cm −1 4301.1 cm−1 4381.2 cm−1
v0d 66 64 83
k0 65.61 N · m−1 65.69 N · m−1 59.01 N · m−1
Eel (15679.0 cm )−1 15703.3 cm−1 15769.1 cm−1

A fluorescence spectroscopy and a rovibronic (rotational-vibrational-electronic) spectroscopy


would be instrumental to completing the picture of diatomic iodine. Although ground state infor-
mation can be garnered from multiple linear regression analysis, Table VIII suggest that Birge-
Sponer may be a better technique, which necessitates an emission spectrum for extracting ground
state parameters. A rovibronic spectroscopy entails a spectrometer that is capable of resolving ro-
tational fine structures, and by performing the same analytical procedures outlined for vibrational-
rotational spectra of HCl and DCl, the properties associated with rotation can be determined. With
bond lengths obtained from rotational analysis, a complete experimentally determined Morse po-
tential curve can be constructed. All the information compiled from the two experiments would
result in a fairly comprehensive characterization of molecular iodine at room temperature.

34
Anharmonic oscillating rotator - HCl

The rigid rotator model is modified to account for centrifugal distortion and bond vibration during
rotation. Unlike the harmonic oscillator, the rigid rotator model is indeed relatively accurate. Al-
though bond length does increase upon vibrational excitation, as seen in Table VII, the anharmonic
oscillating factor is relatively minor because the percent difference between the shortest r and the
longest r across states and across isotopes is only 2.29%. Since the difference between re and r1 is
rather inconsequential, this give credence to that rotation and vibration transpire on a different time
scale; in other words, this justifies and proves that the rigid rotator model is a good approximation
to molecular rotation.

Conclusion

Continuing with the spherical cow analogy, the harmonic oscillator model of molecular vibration
is found to be mostly incompatible with reality and can be rightfully qualified as a spherical cow
model. However it is a simple and convenient concept as a starting point, which explains why it
is still a mainstay in academia. To lose the spherical quality and be more cow-like (i.e, realistic),
the harmonic oscillator model would need to take anharmonicity as well as bond dissociation into
account. The rigid rotator model on the other hand is found to be remarkably accurate for an
approximation model, so on the spherical cow scale, it is tipping towards the realistic, cow-like
side (excuse the puns!). The rigid rotator can be made more accurate by incorporating centrifugal
distortion and anharmonic oscillating effects. Spherical cow is an essential tool to understanding
any natural phenomena, and its significance is best recounted by Sir Arthur Eddington, a British
astrophysicist, in his book The Internal Constitution of the Stars.

I conceive that the chief aim of the physicist in discussing a theoretical problem is
to obtain "insight" – to see which of the numerous factors are particularly concerned
in any effect and how they work together to give it. For this purpose a legitimate
approximation is not just an unavoidable evil; it is a discernment that certain factors

35
– certain complications of the problem – do not contribute appreciably to the result.
We satisfy ourselves that they may be left aside; and the mechanism stands out more
clearly, freed from these irrelevancies. This discernment is only a continuation of a
task begun by the physicist before the mathematical premises of the problem could be
stated; for in any natural problem the actual conditions are of extreme complexity and
the first step is to select those which have an essential influence on the result – in short,
to get hold of the right end of the stick. The correct use of this insight, whether before
or after the mathematical problem has been formulated, is a faculty to be cultivated,
not a vicious propensity to be hidden from the public eye. 12

References

(1) Harte, J. Consider a Spherical Cow: a course in environmnetal problem solving; University
Science Books, 1988.

(2) Chen, W. CHEM-325 Atomic and Molecular Structure Lecture 37. 2013.

(3) Vallance, C. Molecular Energy Levels and Spectroscopy. University of Oxford, 2005.

(4) Severin, K. Raman and Infrared Spectrophotometry. Michigan State University, 2000.

(5) Garland, C.; Nibler, J.; Shoemaker, D. Experiments in Physical Chemistry; Mcgraw-Hill
College, 1995; pp 425–429.

(6) Ramachandran, B. R. Electronic Absorption Spectrum of Iodine. Louisiana Tech University,


Ruston, LA, 2001.

(7) Monk, P. M. S. Physical Chemistry; Understanding our Chemical World; Wiley, 2008.

(8) Loh, A. P. Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine. University of


Wisconsin La Crosse, 2004.

(9) McNaught, I. J. Journal of Chemical Education 1980, 57, 101.

36
(10) Williamson, J. C. Journal of Chemical Education 2007, 84, 1355.

(11) Cooper, P. D. Journal of Chemical Education 2010,

(12) Eddington, A. S. The Internal Constitution of the Stars; Cambridge University Press, 1926.

Appendix

Full derivation of Birge-Sponer fitting line

The selection rule for vibrational transition is ∆v = ±1. For vibrational transition at room temper-
ature,

1 1 1 1
∆E(v0 , v00 = 0) = T + ωe0 (v0 + ) − ωe0 χe0 (v0 + )2 − ωe0 + ωe00 χe00 (1)
2 2 2 4

Equation Eq. (1) is the energy difference between vibrational level v0 in the upper electronic
state and the vibrational level v00 = 0 in the ground electronic state.

∆E(v0 + 1, v00 = 0) − ∆E(v0 , v00 = 0) = ∆Eupper (v0 ) (2)

Equation Eq. (2) is the energy difference between adjacent vibrational levels v0 and v0 + 1 in
the upper electronic state. The first term ∆E(v0 + 1, v00 = 0) refers to the energy difference between
vibrational level v0 + 1 in the upper electronic state and vibrational level v00 = 0 in the ground
electronic state. The second term ∆E(v0 , v00 = 0) refers to the energy difference between vibrational
level v0 in the upper electronic state and the ground vibrational level in ground electronic state.
Deriving Equation Eq. (2) from Equation Eq. (1) is just subtracting Equation Eq. (1) from the
energy difference between v0 + 1 and v00 = 0, thus obtaining the difference between adjacent line.

1 1 1 1
∆E(v0 , v00 = 0) = T + ωe0 (v0 + ) − ωe0 χe0 (v0 + )2 − ωe0 + ωe00 χe00
2 2 2 4

37
The energy difference between v0 + 1 and v00 = 0 is

1 1 1 1
∆E(v0 + 1, v00 = 0) = T + ωe0 (v0 + 1 + ) − ωe0 χe0 (v0 + 1 + )2 − ωe0 + ωe00 χe00
2 2 2 4

T remains the same because v0 + 1 is still in the same electronic state as v0 .

1 1 1 1
∆E(v0 + 1, v00 = 0) − ∆E(v0 , v00 = 0) = T + ωe0 (v0 + 1 + ) − ωe0 χe0 (v0 + 1 + )2 − ωe0 + ωe00 χe00
 2 2 2 4
1 1 1 1
− T + ωe0 (v0 + ) − ωe0 χe0 (v0 + )2 − ωe0 + ωe00 χe00
2 2 2 4
1 1 1 1
= ωe0 (v0 + 1 + ) − ωe0 χe0 (v0 + 1 + )2 − ωe0 (v0 + ) + ωe0 χe0 (v0 + )2
2 2 2 2
= ωe0 − 2ωe0 χe0 (v0 + 1)

= ∆Eupper (v0 )

Full derivation of running line index equation for P and R branch frequencies

The equation for frequencies corresponding to the R branch is

ν̃R = ν̃0 + B1 (J + 1)(J + 2) − B0 J(J + 1) − D (J + 1)2 (J + 2)2 − J 2 (J + 1)2


 
(3)

ν̃R = ν̃0 + B1 J 2 + 3J + 2 − B0 J 2 + J − 4D(J + 1)3


 

= ν̃0 + (B1 − B0 )J 2 + 3B1 J + 2B1 − B0 J − 4D(J + 1)3

Using the definition of Bv from Eq. (18) and α from Eq. (19),

 
1 3
B1 = Be − α 1 + = Be − α
2 2
α
B0 = Be −
2
1 3
B0 − B1 = − α + α = α
2 2

38
Substituting these expressions into Eq. (3)

    
2 3 3 α
ν̃R = ν̃0 − αJ + 3 Be − α J + 2 Be − α − Be − J − 4D(J + 1)3
2 2 2
 
3 1 α
= ν̃0 + 3 Be − α − Be + J + 2Be − 3α − αJ 2 − 4D(J + 1)3
2 3 6
ν̃R = ν̃0 + 2Be − 3α + (2Be − 4α)J − αJ 2 − 4D(J + 1)3 (5)

Repeating the same procedure for ν̃P :

ν̃P = ν̃0 + B1 J(J − 1) − B0 J(J + 1) − D J 2 (J − 1)2 − J 2 (J + 1)2


 
(4)

= ν̃0 + B1 J 2 − B1 J − B0 J 2 − B0 J + 4DJ 3

= ν̃0 + (B1 − B0 )J 2 − B1 J − B0 J + 4DJ 3


   
3 1
= ν̃0 − αJ − Be − α J − Be − α J + 4DJ 3
2
2 2
ν̃P = ν̃0 − (2Be − 2α)J − αJ 2 + 4DJ 3 (5)

Letting the running line index m be m = J + 1 for the R branch and m = −J for the P branch:

ν̃ = ν̃0 + (2Be − 2α)m − αm2 − 4Dm3 (6)

39
Figure 12: Morse potential curve for the upper electronic state.

41
Figure 13: Complete Morse potential of iodine from multiple linear regression. Output gener-
ated from Maple worksheet provided from <http://pubs.acs.org/doi/abs/10.1021/
ed075p1192.3>.

42
(a) Vibrational-Rotational spectra of H35 Cl and H37 Cl

(b) Vibration-Rotational spectra of D35 Cl and D37 Cl

Figure 14: Infrared spectra of gas phase HCl, DCl and their isotopes.

43
Figure 15: Plots of absorption wavenumber versus running line index m in H35 Cl, H37 Cl D35 Cl
and D37 Cl.

44
(a) Bond lengths

(b) Rotational constants

Figure 16: Bar charts comparing experimentally derived bond lengths and rotational constants in
H35 Cl, H37 Cl D35 Cl and D37 Cl.

45

S-ar putea să vă placă și