Você está na página 1de 5

S-2 Investigating the Fine Structure of H and D using Fabry-Perot Interferometry

Austin Yu Liu
School of Applied and Engineering Physics, Cornell University
Ithaca, New York 14850, USA
(Dated: November 29, 2014)
The purpose of this experiment is to measure the fine-structure splitting in the H and D lines
in units of wavenumber, and in so doing obtain a value for the fine structure constant. The measurements are also used to determine the mass ratio of deuterium to hydrogen. The measurements
obtained give a value of the fine-structure splitting to be 0.316 0.016 cm1 for hydrogen, and to be
0.308 0.013 cm1 for deuterium. A value of = 7.04 7 103 from hydrogen fine-splitting and
a value of = 7.00 7 103 from the deuterium fine-splitting was obtained. In addition, a value
mD
= 1.85 0.38 for the mass ratio of deuterium to hydrogen was obtained. The uncertainties
of m
H
above are the statistical uncertainty from the measurements.
The experiment was performed using Fabry-Perot interferometry to obtain an image of the interference fringes from an ionized gas discharge. Subsequently, a charge-coupled device (CCD) camera
was used to obtain an image of the fringes for data analysis.

I.

INTRODUCTION

The purpose of this report is to outline the background


and rationale for investigating the fine-structure of the
H and D lines, as well as to discuss the relevant results obtained from my investigation of the fine structure.
The experiment seeks to determine the the fine-structure
splitting in the H and D lines in units of wavenumber.
The dimensionless fine-structure constant can then be
calculated from the value of the fine-structure splitting.
The interference pattern of the fringes are then used to
determine the mass ratio of deuterium to hydrogen.
The experiment uses a Fabry-Perot etalon for a high
spectral resolution of interference fringes. In the experiment, the light from a ionized gas discharge of mercury,
deuterium and hydrogen is passed through a slit, allowing
for interference fringes to be produced from the etalon.
The etalon consists of two thinly silvered plates that are
highly reflective. The various reflected rays then interfere, as shown in FIG. 2.
The fine structure arises from the relativistic correction
to the kinetic energy of the electron and the spin-orbit
coupling due to the interaction of the electrons spin with
the charge of the proton. Due to the substantial amount
of Doppler broadening in the setup, we do not expect to
observe the Lamb shift [1], which is roughly an order of
magnitude less than the shift predicted by the relativistic correction and spin-orbit coupling. The mass ratio
of deuterium to hydrogen is obtained by considering the
difference in the Doppler broadening of the fringes of the
hydrogen and deuterium spectra. However, the equipment used does not allow for the measurement of the
temperature or pressure in the Woods tube, which affects the Doppler broadening.

II.

do so was Michelson and Morley, who performed their


experiment with a Michelson interferometer in 1887 [2].
In 1913, Bohr obtained an expression for the allowed energies of the hydrogen atom in a somewhat
serendipituous[3] manner. This result was later put on
much firmer footing in 1924 after the Schrodinger equation was proposed. Solving the Schrodinger equation for
the hydrogen atom gives the same energies as those predicted by Bohr [3].
While this expression explained the large scale transitions where electrons transition between different principal quantum number states, it failed to explain the observed fine structure of atomic spectra. In 1928, the fine
structure was finally given a more consistent explanation
by Dirac [4]. The Dirac equation, put forth in [4], allows
for the explanation of spin as a consequence of requiring
quantum mechanics to be consistent with special relativity. The fine structure splitting in hydrogen and other
atoms can then be computed.
The fine structure of hydrogen can be analyzed using time-independent perturbation theory using nonrelativistic quantum mechanics by treating the relativistic correction to the kinetic energy and the spin-orbit
coupling as a perturbation. A standard treatment is
found in, for example, Chapter 6.3 of [3]. Alternatively,
it is possible to solve for the energies of the electron in a
hydrogen atom more directly using the Dirac equation,
which applies to spin-1/2 particles [4] [5]. This accounts
for the relativistic kinetic energy and the spin-orbit coupling.
Solving the Dirac equation for the case of an electron
in a Columbic potential gives the following expression for
the total relativistic energy eigenvalues:

THEORETICAL BACKGROUND

As early as the 19th century, the doublet structure of


the H lines had been noticed [2]. The first group to

0
Enj

2
= mc
1+ 

nj

1
2

1/2

2

q
2 (1)

12

2
+ j+ 2

2
2

e
Here is the fine-structure constant given by = ~c
in CGS units. m above is the mass of the electron, but
should be replaced with the reduced mass of the electronnucleus system to accurately determine the difference in
energies between deuterium and hydrogen. The rest mass
energy of the electron has to be subtracted from this expression to give the observed energy levels of the hydrogen atom. n is the principal quantum number which
defines the principal energy levels and takes on nonnegative integer values, and j is the total angular momentum number, which takes on values 1/2, 3/2. For
a given n, the angular momentum quantum number l
ranges from 0 to n 1. Each electron has a spin quantum number ms which can be 1/2 or 1/2. A detailed
overview of how the Dirac equation is solved for hydrogen
can be found in, for instance, [6]. A Taylor expansion of
the equation above, followed by subtracting the rest mass
energy of the electron, then gives the following expression
(ignoring terms of higher order).

FIG. 1. Experimental Setup

FIG. 2. Fabry-Perot etalon [9]

Enj

mc2 2
=
2


1
2
n
+ 2
n2
n j+

1
2


(2)
III.

The line we seek to observe involves a transition from


n = 3 to n = 2. Ignoring the second term in the expansion, i.e. assuming there is no fine structure, we get the
energy of a transition from n = 3 to n = 2 to be 656.3 nm
for hydrogen and 656.1 nm for deuterium. These are the
H and D lines respectively. The slight difference in
wavelength is due to the difference in the reduced masses
of the two systems. For the n = 2 state, l = 0, 1 and
j = 1/2, 3/2, giving rise to the 2s1/2 , 2p1/2 and 2p1/2
states. Solving the Dirac equation gives a degeneracy in
energy between the 2s1/2 , 2p1/2 states. However, these
two states are in fact not degenerate, with the shift being termed the Lamb shift [1]. We do not expect to
observe this as it is a much smaller effect, roughly an
order of magnitude less than the splitting due to the relativistic kinetic energy and spin-orbit coupling that the
Dirac equation accounts for [1]. The energy difference
between the 2s1/2 /2p1/2 and 2s3/2 states is found to be
2 4
E = mc32 . A summary of the relevant computational
steps can be found at [7]. The corresponding wavenumber is therefore

1
mc4
E =
hc
32h

(3)

From which we can obtain



=

32h

mc

1/4
(4)

DESCRIPTION OF APPARATUS AND


EXPERIMENT
A.

Experimental Setup

As shown in FIG. 1, the experimental setup consists of


a Woods Tube, a vacuum pump, and heavy water(D2 O)
and water (H2 O) samples to generate the discharge. The
Woods tube contains the electrodes of a high voltage
power supply, which is used to ionize the water molecules
in the tube, allowing for electron transitions between the
different energy levels which give rise to the spectrum of
hydrogen.
The mercury arc lamp is used to calibrate the FabryPerot etalon and also to determine the fringe broadening
due to the instrument. The dominant contribution to
fringe broadening at the eyepiece will be assumed to be
the Doppler effect. A detailed discussion of this effect
and other fringe broadening effects can be found in [8].
A brief discussion of the importance of this effect to the
experiment will be discussed later.
The optical setup can be seen on the right side of FIG.
1 The Fabry-Perot etalon is shown in FIG. 2, which is
reprinted from page 274 of [9]. The etalon consists of
two thinly silvered planar plates and produces fringes
as a result of interference between light rays undergoing
multiple reflections between the two plates. The prism
shown in the diagram is used to provide spectral resolution as light of different wavelengths is bent at different
angles by the prism. The Fabry-Perot etalon then provides an even higher spectral resolution.
The image of the fringes can be viewed at the eyepiece,
where the image can be brought into focus by adjusting
the focusing lens accordingly. At the focal plane, the eye-

3
piece may be replaced by a charge-coupled device (CCD)
camera. The CCD camera allows for the image of the
fringes to be captured at a high resolution, allowing us
to generate a plot of the intensity as a function of position.

B.

Methodology and Data

The Fabry-Perot etalon was first aligned by viewing


the fringes produced by the discharge from the mercury
green lamp and checking that the image captured by the
CCD was in focus. Once the alignment of the optical
setup was done, a discharge from the Woods tube was
obtained. This was done as follows: First the Woods
tube was sealed off by closing the air valve and clamping the water and heavy water valves. Next, the vacuum
pump is activated until a sufficiently good vacuum inside the Woods tube is obtained. Then, the clamp for
the water is loosened to allow water vapor to enter the
tube. The corresponding valve is used to regulate the
flow of water vapor in the Woods tube. The high voltage is then applied across the Woods tube to ionize the
gas inside for a discharge. At this point, the main valve
(the one leading to the vacuum pump) was also adjusted
to attain an optimal discharge. After the discharge was
obtained, the fringes were viewed from the eyepiece. As
the position of the mercury green line was known, along
with the relative position of the mercury yellow line, the
position of the H line at the eyepiece could be deduced
and the H line could be viewed accordingly through the
eyepiece. Following this, the CCD camera was used to
check that each fringe of the doublet had approximately
the same intensity.
To switch over to the deuterium discharge and view the
D line, I loosened the heavy water clamp and opened
the heavy water valve before closing off the water valve
and tightening the water clamp. This was to ensure that
the Woods tube remained in discharge throughout the
switching process. Finally, the experiment was closed
off by first unplugging the high voltage source. Next,
I checked that all the water and heavy water clamps
and valves were closed before the vacuum pump was unplugged. The air valve was then opened until the tube
was repressurized. Finally, the air valve was closed off.
The .tif images captured by the CCD camera are shown
in FIG. 3.

IV.

DATA ANALYSIS

From the images captured by the CCD camera, we


obtained a plot of relative intensity against position for
the various gas discharges. The following intensity plots
were obtained by using the NumPy sumleft routine on
the various .tif images captured by the CCD camera, then
removing the baseline noise. This baseline noise was determined from an intensity plot of the mercury fringes

FIG. 3. Images captured by the CCD camera for H, D and


Hg discharge respectively

FIG. 4. Intensity plot of fringes from deuterium discharge


with fitted curves

by assuming the mercury fringes to be relatively widely


spaced Gaussian functions, and taking the value of the
intensity between the Gaussians to be the noise.
Following this, the plots in FIG. 4, FIG. 5 and FIG.
6 were fitted to a linear sum of Gaussian functions that
were displaced along the position coordinate. All numerical manipulations were handled using the NumPy and
SciPy libraries in Python.
For constructive interference between the rays at the
etalon, the rays must satisfy

TABLE I. Table of results for deuterium


Splitting (cm1 )
0.323
0.310
0.291

Distribution SD
16.1
16.8
15.7
17.2
14.4
16.4

FIG. 5. Intensity plot of fringes from hydrogen discharge with


fitted curves

FIG. 6. Intensity plot of fringes from mercury discharge with


fitted curves

2dcos() 2t =

(5)

where we assume that is a sufficiently small angle and


t is the distance between the plates of the etalon. Here,

TABLE II. Table of results for hydrogen


Splitting (cm1 )
0.338
0.313
0.300

Distribution SD
22.0
19.4
19.4
18.8
18.7
18.7

is the wavenumber, which serves as a unit of energy.


For our etalon, d = 0.7000 cm. Hence we obtain a value
1
0.714 for the wavenumber between orders.
of 2d
To obtain the fractional order shift, the intensity plots
for the hydrogen and deuterium discharges were fitted to
Gaussian curves for the three central doublets in FIG. 5
and FIG. 4. From this, the relative positions of the intensity maxima and the relative separation between the
doublets could be obtained. The fractional order shift
was found by finding the ratio of half the distance between two fringes of the same doublet to the distance
between doublets. A summary of the relevant data can
be found in Tables I and II, where SD is an abbreviation for standard deviation. Consequently, for hydrogen, I obtained a fractional order shift in the H line of
0.222 0.011 for the shifts measured from FIG. 5. For
deuterium, I obtained a fractional order shift in the D
line of 0.216 0.009 over the shifts measured from FIG.
4. The error bars are due to statistical uncertainty. Multiplying the value of the wavenumber between orders, ,
by twice the fractional order shift directly gives the fine
structure splitting. For hydrogen, this is found to be
0.316 0.016 cm1 , and for deuterium, this is found to
be 0.308 0.013 cm1 . These values are in agreement
with those obtained by R.C. Williams [2], who measured
a splitting of 0.319 cm1 for hydrogen and 0.321 cm1
for deuterium. The theoretical value of the splitting is
0.328 cm1 [2].
To determine the fine-structure constant, we use the
expression outlined above in equation (4). Hence we obtain a value of = 7.04 7 103 from the splitting in
hydrogen and a value of = 7.00 7 103 from the
splitting in deuterium. This is within 5 percent of the
currently accepted value of = 7.29 103 , where is
in CGS units.
In order to find the mass ratio between deuterium and
hydrogen, we considered the fringe broadening as a result
of the Doppler effect. Qualitatively, the Doppler effect
causes fringe broadening because the atoms have some
velocity relative to the optical setup when they emit a
photon. As the velocities are distributed about a mean
velocity of zero, the frequencies emitted are similarly distributed. As a result, the intensity is also distributed
about an expected frequency corresponding to the original, unshifted frequency. This is seen as a broadening of
the fringe.
Quantitatively, if the velocities follow a MaxwellBoltzmann distribution about the mean velocity, the
probability distribution function for the velocity in a particular direction is given by
r
dw
u2
=
e
(6)
du

where u is the component of the velocity in the direction incident on the optical setup, = 2km
, and m is
BT
the mass of the particle. The Doppler frequency shift,
u
, is given by
0 = c . Hence the relative intensity as
a function of the frequency is found to be

5
V.

I() e

c2 (0 )2
2

(7)

This implies that


p the full width at half maximum
is
proportional
to
1/ and is hence proportional to
p
1/m. Hence the mass ratio of deuterium to hydrogen can be determined by the ratio of the widths of
the respective intensity curves. The full width at half
maximum of a Gaussian distribution is directly proportional to its standard deviation, so comparing the ratio
of the standard deviations for the deuterium and hydrogen fringes will give the same result. Assuming that the
fringes are approximately Gaussian, in the relative position units above, the standard deviation of the hydrogen
fringes is found to be 19.5 1.2 and the standard deviation of the deuterium fringes is found to be 16.1 0.9.
However, the above two values do not account for the
broadening introduced by the optical setup. Using the
data from the mercury fringes, we can approximately account for this. As the atomic mass ratio of hydrogen to
mercury is roughly 200, the broadening in the mercury
fringe lines can be attributed almost entirely to the optical setup. First we obtain a value of 10.8 1.0 for the
standard deviation of the mercury fringes. The true standard deviation of the hydrogen and deuterium fringes can
be found by considering:
2
2
2
setup
+ true
= observed

CONCLUSION

The following is a summary of the relevant results. The


fine structure splitting is found to be 0.316 0.016 cm1
for hydrogen, and found to be 0.308 0.013 cm1 for
deuterium, from which a value of = 7.04 7 103
from the splitting in hydrogen and a value of = 7.00
7 103 from the splitting in deuterium was obtained.
This is within 5 percent of the currently accepted value
of = 7.29 103 [10], where is in CGS units. An
analysis of the fringe widths by comparing the standard
deviations of the Gaussians used to fit the intensity plots
mD
= 1.85 0.38 for the mass ratio
of the fringes gives m
H
of deuterium to hydrogen, which is within 20 percent of
the accepted value of 1.998 [10]
Further improvements on the experiment might involve having pressure and temperature readings within
the Woods tube so that these variables can be accounted
for in the analysis of the mass ratio of deuterium to hydrogen. In addition, the setup still appears so be somewhat out of alignment, as evidenced by the slant in the
mercury fringes in FIG. 6. More accurate results may
be obtained if this alignment could be improved. It may
also be beneficial for future students to ensure that the
etalons alignment is checked before the start of the experiment to avoid unnecessary frustration.

(8)

This gives the standard deviation for the hydrogen


fringes to be 16.2 1.8 and the standard deviation for
the deuterium fringes to be 11.9 1.8. Thus we obtain a
value of 1.36 0.26 for the ratio of the standard deviamD
= 1.850.38 for the mass
tions. This gives a value of m
H
ratio of deuterium to hydrogen. The large percentage uncertainty is attributed to the statistical uncertainty in the
standard deviation of the Gaussians that were fitted to
the intensity plot.

[1] W. E. Lamb and R. C. Retherford, Physical Review 72,


241 (1947).
[2] R. C. Williams, Phys. Rev. 54, 558 (1938).
[3] D. Griffiths, Introduction to Quantum Mechanics (Pearson Prentice Hall, Upper Saddle River, NJ, USA, 2005)
pp. 278288.
[4] P. A. M. Dirac, Royal Society of London Proceedings
Series A 117, 610 (1928).
[5] P. A. M. Dirac, The Principles of Quantum Mechanics,
4th ed., International Series of Monographs on Physics
(The Clarendon Press, Oxford, 1958) pp. 253275.
[6] A. Whitehead, A Relativistic Electron in a Coloumb Po-

ACKNOWLEDGMENTS

I would like to thank Prof. Don Hartill for his assistance and guidance with the experimental setup. I would
also like to thank Dr. Eric Smith, who gave me useful
tips on the data collection.

[7]

[8]
[9]
[10]

tential, http://www.physics.drexel.edu/~bob/Term_
Reports/Whitehead_3.pdf (2009), accessed: 2014-09-23.
C. U. Department of Physics, S-2 Fine Structure of Hydrogen,
http://pages.physics.cornell.edu/p510/
S-2_Fine_Structure_in_Hydrogen (2013), accessed:
2014-09-24.
H. E. White, Introduction to Atomic Spectra (McGrawHill Kogakusha, ltd., 1934).
F. A. Jenkins and H. E. White, Fundamentals of Optics
(McGraw-Hill, 1981).
N. I. of Standards and Technology, CODATA Recommended Values of the Fundamental Physical Constants:
2010 (2010).

Você também pode gostar