Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

PHYS3040 Laboratory – Student Guide

Molecular Spectroscopy

IMPORTANT SAFETY INFORMATION


This experiment uses a Class III helium-neon laser. The output from these lasers can
be hazardous to the eyes.
Prior to turning on the laser for the first time, you must go through a safety induction
with your tutor and sign the safety register
The laser beam has been shielded to avoid exposure. Only lift the shield in the
presence of a tutor and after switching off the lasers.

1 Introduction
Atomic spectroscopy is the study and use of of the unique frequency (or wavelength) profile of the electromagnetic
spectrum absorbed or emitted by a substance. You will be familiar with atomic spectroscopy from earlier study.
In this practical, we focus on molecular spectroscopy, which is substantially more complicated (and interesting!)
than the atomic case, due to the extra degrees of freedom. In particular, you will investigate diatomic iodine,
one of the simplest molecules able to exhibit phenomena unique to molecular spectroscopy.

Learning Goals
• Appreciate the different spectral features arising from electronic, rotational and vibrational degrees of
freedom.
• Relate observed patterns to the mathematical description and the relevant quantum numbers.
• Use derived model and experimental data to predict known quantities of diatomic iodine; the vibrational
frequency ωe and anharmonicity constant ωe xe .
• Demonstrate laboratory procedures, including the use of lab books and record keeping, proper and justified
procedure, and error control and propagation.

Session 1 Aims
By the end of session 1, you should have
• Collected broad spectrum results for both the green and orange wavelengths.
• Attempted to observe a double peak
• Completed the mathematical model and have this checked over by a tutor. Between sessions 1 and 2, you
should map your observed spectrum to the predicted. Session 2 should be largely devoted to attempting
to filling in any gaps.
If you are nearing the end of your first session an have not completed these steps, please contact a tutor.

2 Experiment
Equipment
This experiment consists of a vial containing a sample of diatomic iodine, a helium-neon laser used to excite the
sample, and an Andor Solis spectrometer to record the resulting emission spectra. You are tasked to control
the spectrometer and optimise the emission spectra for at least two excitation wavelengths.

1
The default set up (as demonstrated in the video on Blackboard) is to use a tunable laser, capable of emitting
light over a range a frequencies. Two excitation wavelengths most useful for inducing fluorescence in the iodine
are 543.003 nm (green) and 611.971nm (orange). The frequency is controlled by the knob on the back of the
laser (see Figure 1).
The alternative set-up uses different fixed-frequency lasers to excite at different wavelengths. Please consult
your tutor for how to swap between the different lasers.

Figure 1: The back of the tunable Helium-Neon Laser

Most of the controls you need are in the Andos Solis software program, installed in the laboratory computer.
Figure 2 shows what you should see when you open the screen.
The various software controls in Figure 2 are:
1. This leads to save and export options. You should export the collected data as an ASCII file, for later
import to e.g. MATLAB.
2. This button starts the acquisition of data, which will be displayed on the screen. The data will not be
saved or recorded automatically. You must start an acquisition, stop an acquisition (see 11) and then
export the data you wish to use.
3. This button starts a run of acquisition according to acquisition parameters, as set by 5. This acquisition
will stop automatically, but will still need to be exported or saved.
4. Stops the acquisition.
5. Acquisition set-up. You can then specify all the parameters which will be used when running acquisition
(see 3). In particular, you can define the vertical binning range. The vertical size of the CCD matrix of
the photo detector is 256 pixels. By default, the spectrum is obtained by adding the signal in in all pixels
in the vertical column. This increases the number of detected photons but reduces the spectral resolution
because the optical aberration in the system causes distortion and the entrance slit is not imaged properly.
6. Re-scale for fit vertically in the screen.
7. Start/stop auto-scaling.
8. Click to set the temperature of the spectroscopy detector (CCD). The recommended temperature is -70◦ C.
You will know when it has reached the set temperature once it is coloured blue, as indicated in this image.
Once it has stabilised you can start acquisition.
9. The grey bar indicates the observed spectral region, and can be shifted up and down in frequency by
clicking and dragging. The width of spectrum is set by the choice of grating (see 11).
10. The dots indicates the range of sensitivity of the CCD sensor.
11. Displays the grating choice. You can select a different one by clicking on the corresponding coloured bar.
The available gratings are numbered from 1 to 4 clockwise:

2
Figure 2: The Andor-Solis Software

• Grating 1 (green bar): 300 lines/mm


• Grating 2 (red): 600 lines/mm
• Grating 3 (cyan): 1200 lines/mm .
• Grating 4 (violet): 1800 lines/mm
Use the coarse grating 1 to obtain a wide spectrum, then when you want to zoom into obtain more spectral
detail switch to the fine gratings 3 and 4. It is a good idea to stop acquisition (by pressing the red stop )
before changing any setting.
12. Choose the size of the input aperture (iris). This is not the entrance slit but an additional aperture which
limits the solid angle where the light is collected. Larger aperture sizes increase the intensity of the signal
but decrease the spectral resolution. There are also two additional hardware controls on the spectrometer
that allow you to further adjust the amount of light.
13. Select exposure time. You should aim to minimise noise without losing smaller signals. It may be beneficial
to view different parts of the spectrum with different exposure times.
If you do not see the menu on the bottom half of the screen, you can open this using the following sequence
illustrated in Figure 3.
When running the acquisition of data, ensure the scale of the output graph matches the scale of the grey bar
(see Figure 4), which represents the wavelength range. If the graph reads an x axis scale of 0 to 1000, it is
returning the scanner counts, and will need to be re-scaled. In such a case, halt the acquisition, close the graph
window, and recalibrate the axis using the steps outlined in Figure 5. Then re-start the acquisition.

3
Figure 3: To open the correct menu, choose the Spectography Setup option in the hardware menu, and then
make sure the settings match the picture above before clicking ’OK’

Figure 4: When correctly calibrated, the range of the grey bar will match the range displayed on the given
graph. This is important, as this represents how the exported data will be labelled.

For Investigation
We want to determine a model to predict the measured spectral lines, and then verify its ability to do so for a
simple diatomic gas, I2 . Consider the simple approximation of a diatomic molecule as two atoms connected by
a spring. The spectral lines of a molecule are must more complex than that of an atom due to the additional
degrees of freedom. Along with changes in atomic shell, diatomic molecules can also be subject to vibrations
and rotational energy changes. The total energy can thus be written:

T = T (n) + G(v) + F (J), (1)

which gives the total energy in terms of familiar electron excitation energy T (n), the vibrational energy G(v),
and the rotational energy F (J), labelled with their respective quantum numbers n, v, and J.
1. By modeling a diatomic molecule as two masses on a massless spring creating an anharmonic oscillator,
determine the energy levels dependant on quantum number v as

Eosc (v) = hνosc (v + 1/2) − hνosc xe (v + 1/2)2 (2)

and then divide by hc to determine an expression of energy in units of cm−1 :

G(v) = ωe (v + 1/2) − ωe xe (v/2)2 , (3)

where ωe is the frequency in units of cm−1 , and xe is a constant related to the anharmonicity. Support
for this task and that below are given in these notes.
2. By modelling the diatomic molecule as a rigid rotator, determine the rotational energy levels in terms
of rotational quantum number J. You should include a second order approximation to account for any

4
Figure 5: To calibrate the display and export data, halt all acquisitions, then under the Calibrate menu, choose
the X-Calibration by Spectrograph, and choose confirm ’Calibrate’ on the resulting pop-up window.

stretching in the molecule. As above, divide by hc to give an expression for energy in units cm−1 :

F (J) = Bv J(J + 1) + Dv J 2 (J + 1)2 (4)

where Bv and Dv are constants that depend on the vibrational state.


3. Model the viewed spectrum as a series of random decays from a single excited state, initiated by the laser.
Mathematically determine the excited state using the wavelength of the laser and the attached tables.
Justify, using a thermal distribution argument, the assumption that all molecules are excited to the same
state.
4. Selection rules governing the transition mean that the rotational energy level can only change by J = ±1.
Mathematically determine whether these transitions could be in the observed spectrum.
5. Using the attached tables, predict the expected observed transitions and their corresponding wavelengths
for the transition of the predicted excited state to the 20 available de-excited states for the green and
orange laser. Use this to label the vibrational quantum number for your observed data.
6. Perform a linear regression relating the vibrational quantum number and the energy of transition. Hint:
consider the change in transition energy between quantum number n and quantum number n + 1. Using
your data, you can compare the resulting vibrational frequency ωe and anharmonicity constant ωe xe to
the expected values for diatomic iodine.
7. You should attempt to view and export a double peak, a single photon count spike with a distinct double
peak when zoomed in upon. This should be discussed qualitatively. What might explain this phenomena?
Consider in particular the scale of separation of the peaks.
8. You should observe some transitions that occur at a shorter wavelength, or higher energy, than the
excitation of the laser. Explain this phenomena.

3 Further Background and Theory


3.1 Molecular energy levels
When we compare an atomic to a molecular spectrum, we see very quickly that molecular spectra are much
more ‘crowded’ with bands of peaks visible, rather than the well-defined lines of the atomic spectrum. This
difference arises from the additional degrees of freedom possessed by a molecule: it can both vibrate and rotate,
as well as undergoing electronic transitions. In this experiment, we focus on I2 , and so restrict our theory to
that of homonuclear diatomic molecules (i.e. molecule made up of two of the same atoms). (We base our theory
on Ref [1] chapter III. You can also find useful explanations in Ref [2] chapters 9 & 13.) The simplest model
is to consider the two atoms as two point-like masses joined by a spring. They can then vibrate in and out
along the joining spring and rotate around any axis perpendicular to the joining axis. Both of these processes
introduce extra energy levels on top of the electronic levels you know from atomic spectroscopy. Specifically,
each electronic level is split into many vibrational levels, which, in turn, are split into many rotational levels.
This structure is illustrated schematically in Fig 6.

5
Figure 6: Schematic diagram of molecular energy levels, showing vibrational and rotational structure.

Figure 7: Vibrational potential curves for different models: SHO (dashed line); actual potential curve (solid
line; drawn for the ground state of HCl); anharmonic oscillator (dotted line). (From Ref [1] p. 91)

The energy of a given rovibrational level is then the sum of its electronic, vibrational, and rotational components,
E = Eel + Eosc + Erot . We will examine each type of motion separately, before putting them together to obtain
the energy levels of a vibrating rotator.

3.2 Electronic energy


First, we assume that the nuclei of the molecule have a fixed positions in space. In this case, energy of the
electrons or electronic energy is quantized and has a set of defined values, similar to the case of an atom.
However, this energy depends on the distance between the nuclei and therefore instead of a set of discrete
values one gets a set or curves Ee (r). A typical curve in a stable molecule will have a minimum value at the
equilibrium position of the nuclei and it looks like the curve shown by solid line in Fig.7. This energy can be
seen as potential energy of the nuclei. The motion of the nuclei adds vibrational energy and rotational energy
to the electronic energy.

3.2.1 Vibrational motion


Turning first to vibrational motion, our simplest option is to model the potential energy as a simple harmonic
oscillator, where the force on the atoms is proportional to the change in their internuclear spacing from equi-
librium. Ideal harmonic oscillator has a parabolic potential and this approximation holds only if the position
of nuclei is very close to the equilibrium where the potential curve can b e approximated by a parabola. The
energy levels available are quantised, and we write them in terms of a vibrational quantum number v as
 
1
Eosc (v) = hνosc v + (5)
2

where νosc is the vibrational frequency calculated classically and v can take integer values 0,1,2,.... It is clear
from eqn 5 that this model leads to equally spaced vibrational energy levels.
This model is unrealistic for large internuclear spacings. In a real molecule, in contrast, there is no restoring

6
force experienced once the atoms are infinitely separated, which implies that the potential curve should level
off as the internuclear spacing increases (see Fig 7).
A better approximation to the real potential curve is to treat the molecule as an anharmonic oscillator, basically
by adding a cubic term on to the potential energy expression, to obtain
2 3
V = a (r − re ) − b (r − re ) (6)

where r is the internuclear spacing, re is it’s equilibrium value, and β ≪ α. The corresponding energy levels
are given by
   2
1 1
Evib (v) = hνosc v + − hνosc xe v + (7)
2 2
We divide through by hc to obtain what is known as the vibrational term value for a given vibrational energy
level    2
1 1
G (v) = ωe v + − ωe x e v + (8)
2 2
with ωe = νosc
c known as the vibrational frequency, and ωe xe known as the anharmonicity constant. Both ωe
and ωe xe have units of cm−1 ; as does therefore the term value G. (One of your main aims in this prac will be
to determine ωe and ωe xe for I2 .) From eqn 7 we see that the separation between the levels of an anharmonic
oscillator decreases as v increases.

Thermal distribution of vibrational levels Continuing to consider only vibrational levels of a molecule,
we now examine the probability that, at thermal equilibrium, a molecule will be in a particular vibrational
state. This comes to the same thing as asking about the number of molecules in each state when we observe a
large number of molecules (as is the case in many spectroscopy experiments).
We might expect molecules to tend to ‘fall to the bottom of the potential well’ (i.e. the bottom of the anharmonic
potential curve) and thus sit in the lowest vibrational state. Think of little beads in a bucket shaped like the
potential curve. At temperatures above absolute zero, however, the molecules will have some thermal energy;
this is equivalent to the beads ‘bouncing around’ inside the bucket and results in some beads being found higher
up in the bucket. In other words, above T = 0 K, some molecules will occupy higher vibrational states.
The probability of a molecule occupying a given state is governed by what is called the Boltzmann distribution.
(You will hear a lot more about this in your statistical
 mechanics course. For more information, see Ref [3] p. 220-
227) Specifically, this probability goes as exp −E
kT where E is the energy of the level, k is Boltzmann’s constant,
and T is the temperature in K. Hence N , the number of molecules in a vibrational state with a vibrational
quantum number v decreases exponentially
Evib
N ∝ exp − (9)
kB T
This is also shown for I2 in Fig 8.
The influence of temperature on vibrational level populations means that we can extract temperature infor-
mation from vibrational spectra by examining the relative intensities of different emission lines. Spectroscopic
measurements are often used to measure temperature in inaccessible samples, like hot gases.

3.2.2 Rotational motion


In the same way we built up our model of vibrational motion from the simplest assumption, we start here by
treating our molecule as a rigid rotator. In other words, we ignore that it also vibrates, which means that there
is no potential energy associated with the rotation. Solving the Schrodinger equation for this situation, we
obtain rotational energy levels in terms of the rotational quantum number J

h2
Erot (J) = J (J + 1) (10)
8π 2 I

where I is the moment of intertia, and J can take integer values 0,1,2,.... The spacing between energy levels
described by eqn 6 increases quadratically with J. Dividing eqn 6 through by hc again gives us the corresponding
term value, designated F for rotational energy levels
h
F (J) = J (J + 1) = BJ (J + 1) (11)
8π 2 cI

7
Figure 8: Thermal distribution of vibrational levels according to the Boltzmann factor. (Here, vibrational
energies correspond to the ground state of I2 at T = 300 K. From Ref [1] p. 122.)

where B is the rotational constant and is inversely proportional to the moment of inertia.
We know, however, that our molecule is not a rigid rotator: rather than comprising two point atoms joined
by a massless rod, we have modelled it as two points joined by a massless spring. Our first refinement to
eqn 11 consists of adding in the possibility that the spring could stretch during rotation. Indeed, this is exactly
what will occur as a result of centrifugal forces as the rotation increases. By equating the centrifugal force
to the restoring force that will result from the small stretching, we can find the new equilibrium internuclear
distance, and thus derive a correction to eqn 11 (see Ref [1] p 104-105 for details). This takes the form of adding
higher-order terms, as we did with the vibrational energy levels. We obtain
2
F (J) = BJ (J + 1) − DJ 2 (J + 1) + . . . (12)

where D is another rotational constant and is small compared to B. D gives a measure of the influence of
centrifugal force on the molecular energy. In terms of the energy-level spacings, the quadratic correction of
eqn 12 is mainly significant at higher rotational energy levels (or larger values of J).
Finally, we have to consider the effect of simultaneous vibration and rotation on our model, which will essentially
manifest in the values of the rotational constants B and D. Recall that B is inversely related to the rotator’s
moment of inertia. As the molecule vibrates, B will therefore vary, because the molecule’s moment of inertia
depends on its internuclear spacing. To a first approximation, B then takes the form
 
1
Bv = Be − αe v + + ... (13)
2

where we have used the subscripts v and e to distinguish the vibrating (v) rotator from its rigid, or equilibrium
(e), counterpart. αe is a constant that is small compared to Be . (We won’t go any further into how to derive
eqn 13 as Ref [1] describes the calculations as “rather involved”.) A similar relation obtains for Dv , which then
also depends on vibrational number v.
Substituting these vibration-corrected constants into our rotational term value expression of eqn 12, we obtain
its final form of
2
F (J) = Bv J (J + 1) − Dv J 2 (J + 1) + . . . (14)

The full term value for a rovibrational energy level is simply the sum of electronic (Te ), vibrational, and
rotational term values, or
   2
1 1 2
T = Te + ωe v + − ωe x e v + + Bv J (J + 1) − Dv J 2 (J + 1) (15)
2 2

where n is the electronic quantum number. Don’t forget that T has units of cm−1 .

8
3.2.3 Thermal distribution of rotational levels
Just like for the vibrational states, the Boltzmann factor also regulates the thermal distribution of molecules
in rotational states. However, we must also take into account that each rotational level is actually (2J + 1)
degenerate. In other words, there are actually multiple levels (to the number of (2J + 1)) with rotational
quantum number J, so all with the same rotational energy. (These levels split in the presence of an external
magnetic field, however, an effect known as Zeeman splitting.) This multiplicity of levels means we have to
multiply the Boltzmann factor by the number of levels (2J + 1) in order to obtain the distribution of molecules
in those states. You can think that more levels means there is a greater probability of a molecule being in that
state. The consequence for the thermal distribution is that it no longer declines exponentially with energy as
for the vibrational levels. Therefore there is a peak of population at some J, depending on the temperature
and the value of the rotational constants.

3.2.4 Thermal distribution of rovibrational levels


To obtain the thermal distribution of molecules into rovibrational levels, we simply combine the distributions
for vibrational and rotational levels considered differently. That is, the total population in all the rotational
levels of a given vibrational level will be set by the Boltzmann distribution of vibrational levels. Then within
that vibrational level, the rotational populations will be distributed as described in the previous section. If
v > 0, there will just be fewer molecules overall to be distributed into the rotational levels, but the shape of the
distribution does not change.

3.3 Transitions and molecular spectra


Just as in the familiar atomic case, molecular spectra appear as the result of transitions between different energy
levels. Electronic transitions form the coarsest structures in molecular spectra, corresponding to a wavelength
in the visible or ultraviolet regions. The scale of vibrational levels corresponds to infrared wavelengths, while
that of rotational levels is on the order of microwaves.
All transitions must obey selection rules which derive from quantum theory. For vibrational transitions in an
anharmonic potential, ∆v is unrestricted; however, rotational transitions may only occur for ∆J = ±1.
Transitions may occur either in absorption or emission (as with atomic systems). In this experiment, we study
emission spectroscopy, which entails using light of fixed frequency to excite the molecule of interest from the
ground state into an excited state. The excited molecule can then decay optically (i.e. via fluorescence) to a
range of lower levels, yielding many spectral lines. We record a spectrum consisting of the observed signal at
each detector wavelength.
Providing that the fluorescence also involves an electronic transition, there is no requirement that the final
vibrational and rotational quantum numbers be lower than those of the initial state. Similarly, exciting tran-
sitions with decreasing vibrational or rotational quantum numbers are permitted. Which transitions are most
likely to occur is governed by the Franck-Condon principle, which is based on electrons retaining their relative
position and velocity immediately before and after the ‘jump’ between vibrational levels of different electronic
states (see Ref [1] p. 194-198 for more details).
The energy change for a transition may be considered as the sum of electronic and (positive or negative) vibra-
tional and rotational energy differences. We usually calculate the wavenumber corresponding to the transition
as v = Tupper − Tlower , which has units of cm−1 . When you convert to wavelength, do not forget to correct for
the observed wavelength in air compared to that in a vacuum.

3.3.1 Nomenclature
By convention, we designate the upper rovibrational level in any transition by quantum numbers v ′ and J ′ ,
while the lower state is designated v ′′ and J ′′ . Therefore a given vibrational transition is written v ′ –v ′′ (eg.
26–0). This form does not change regardless of whether the transition involves absorption or emission.
As for the rotational part of transitions, we denote transitions with ∆J = −1 by the letter R, while transitions
with ∆J = +1 are termed the P branch. The difference in J is calculated here relative to the lower state.
Thus a transition denoted R(5) will have J ′ = 6 and J ′′ = 5, while P(5) will have J ′ = 4 and J ′′ = 5. This is
illustrated schematically in Fig 9.

9
Figure 9: Schematic illustration of molecular transition nomenclature. (Based on Ref [2] Fig. 13.36)

4 Tables
Note that in the tables (overpage), the electronic and vibrational terms are already combined into the value Ev ,
with Ev = Telec + G (v) − G (0). Tables 1 and 2 are also provided on Blackboard as Excel spreadsheets (only
v ≤ 28 for Table 1) to simplify importation into Matlab; just open the files with the uiimport command.

References
[1] Herzberg, G., Molecular Spectra and Molecular Structure. 2nd ed. Vol. 1: Spectra of Diatomic Molecules.
1950: Van Nostrand. [QC451.H463 v.1]
[2] Atkins, P. and J. de Paula, Atkins’ Physical Chemistry 2006: Oxford University Press. [QD453.A88]
[3] Schroeder, D. V., An Introduction to Thermal Physics. 2000, San Francisco: Addison Wesley Longman.
[QC311.15 .S32]
[4] Williamson, J. C., Molecular Iodine Fluorescence Spectra Generated with Helium-Neon Lasers for Spectrom-
eter Calibration. Applied Spectroscopy, 2010. 64(12): p. 1419-1422.
[5] Gerstenkorn, S. and P. Luc, Description of the Absorption Spectrum of Iodine Recorded by Means of Fourier-
Transform Spectroscopy – The (B-X) System. Journal de Physique, 1985. 46(6): p. 867-881.
[6] Lerner, J. M. and A. Thevenon, The Optics of Spectroscopy: A Tutorial. Horiba Scientific. [accessed 4 March
2014]; Available from: http://www.horiba.com/us/en/scientific/products/optics-tutorial/
[7] Paschotta, R., Spectrometers. Encyclopedia of Laser Physics and Technology [accessed 4 March 2014]; Avail-
able from: http://www.rp-photonics.com/spectrometers.html.
[8] Kramida, A., Ralchenko, Yu., Reader, J., and NIST ASD Team, NIST Atomic Spectra Database (ver. 5.1)
2013, Gaithersburg, MD: National Institute of Standards and Technology. [accessed 9 March 2014]; Available
from: http://physics.nist.gov/asd

Kim Hajek 2014


Taras Plakhotnik 2020
Cassandra Bowie, Joel Corney 2024

10
Table 1: Energies and rotational constants of the B-state in I2 . Note that Ev = Telec + G (v) − G (0). All values
in cm−1 . (Taken from Ref [5] Table V.)

11
Table 2: Energies and rotational constants of the X-state in I2 . Note that Ev = Telec + G (v) − G (0). All values
in cm−1 . (Taken from Ref [5] Table V.)

12

You might also like