Docsity
Docsity

Prepare for your exams
Prepare for your exams

Study with the several resources on Docsity


Earn points to download
Earn points to download

Earn points by helping other students or get them with a premium plan


Guidelines and tips
Guidelines and tips

Doppler free Saturated Absorption Spectroscopy: Laser Spectroscopy | PHYS 3340, Lab Reports of Physics

Material Type: Lab; Class: Introductory Research in Optical Physics; Subject: Physics; University: University of Colorado - Boulder; Term: Unknown 1989;

Typology: Lab Reports

Pre 2010

Uploaded on 02/13/2009

koofers-user-ktf
koofers-user-ktf 🇺🇸

4

(1)

10 documents

1 / 28

Toggle sidebar

Related documents


Partial preview of the text

Download Doppler free Saturated Absorption Spectroscopy: Laser Spectroscopy | PHYS 3340 and more Lab Reports Physics in PDF only on Docsity! Advanced Laboratory Spring 2001 Laser Spectroscopy LS.1 Spring 2001 DOPPLER-FREE SATURATED ABSORPTION SPECTROSCOPY: LASER SPECTROSCOPY Overview In this experiment you will use a diode laser to carry out laser spectroscopy of rubidium atoms. You will study the Doppler broadened optical absorption lines, and will then use the technique of saturated absorption spectroscopy to study the lines with resolution beyond the Doppler limit. This will enable you to measure the hyperfine splittings of one of the excited states of rubidium. You will use a Michelson interferometer to calibrate the frequency scale for this measurement. This experiment was developed by Carl Wieman and uses techniques currently in use in his and other research laboratories at CU. It is being duplicated at a number of other schools. To facilitate the widespread use of this experiment, which is new to undergraduate labs, Daryl Preston (author of your optional text), in collaboration with Wieman wrote a very lengthy writeup. The purpose of this writeup was to allow schools to duplicate the experiment, even if the lab instructors were not familiar with laser spectroscopy. As a result, it is somewhat different in style and far more detailed than your other writeups have been. Although it is long, it does explain in detail everything you will need to know to carry out and analyze this experiment so you will have little need of other references. References to other chapters or experiments refer to items in Preston's book "The Art of Experimental Physics". Historical Note One half of the 1981 Nobel prize in Physics was awarded jointly to Arthur L. Schawlow, Stanford University. During the 70's, Professor Schawlow's research group developed and applied the technique of Doppler-Free Saturation Absorption Spectroscopy. Apparatus Tunable 780-nm diode laser system (see reference 1) Rubidium vapor cell Photodiode detector circuit (see reference 1) Triangle waveform generator 2' x 2' x 1/2" optical breadboard 3/8"-thick transparent plastic or glass (beam splitter) 4 flat mirrors 4 mirror mounts 9 posts (for mounting vapor cell, photodiode detectors, mirrors, and beam splitter) Oscilloscope (Optional: storage scope with a plotter) Oscilloscope camera Kodak IR detection card Hand held IR viewer or a CCD surveillance camera Purpose 1) To appreciate the distinction between linear and nonlinear spectroscopy. 2) To understand term states, fine structure, and hyperfine structure of rubidium. 3) To record and analyze the Doppler-broadened 780-nm rubidium spectral line (linear optics). 4) To record and analyze the Doppler-free saturated absorption lines of rubidium (nonlinear optics), and thereby determine the hyperfine splitting of the 5P3/2 state. Laser Spectroscopy LS.2 Spring 2001 Key Concepts Linear optics Electric dipole selection rules Nonlinear optics Doppler broadening Absorption spectroscopy Fine structure Saturated absorption spectroscopy Hyperfine structure References 1. K. B. MacAdam, A. Steinbach, and C. Wieman, Am. J. Phys. 60, 1098-1111, 1992. The construction of the diode laser system and the Doppler-free saturated absorption experiment are discussed in detail in this paper. Parts, suppliers, and cost are listed. 2. M. D. Levenson, Introduction to Nonlinear Laser Spectroscopy, Academic Press,1982 3. Carl E. Wieman and Leo Hollberg, Rev. Sci. Instrum. 62(1), January 1991, pp. 1- 20. Diode lasers in atomic physics are reviewed. The list of references is extensive. 4. J. C. Camparo, Contemp. Phys., Vol. 26, No. 5, 1985, 443-477. This is the first paper to review the diode laser in atomic physics. 5. A. Corney, Atomic and Laser Spectroscopy. Oxford Press 1977. Hyperfine interactions are discussed in chapter 18. 6. V. S. Letokhov, “Saturation Spectroscopy”, Chapter 4 of High Resolution Laser Spectroscopy (Topics in Applied Physics, Vol. 13, ed. K. Shimoda), Springer-Verlag, 1976. 7. T. W. Hansch, Nonlinear high resolution spectroscopy of atoms and molecules, in “Nonlinear Spectroscopy” (Proc. Int. School Phys., Enrico Fermi, Course 64) (N. Bloemberger, ed.). North-Holland Publi. Amsterdam, 1977. 8. W. Demtroder, Laser Spectroscopy (Springer series in Chemical Physics, Vol. 5), Springer, New York, 1982. 9. E. Arimondo, M. Inguscio, and P.Violino, Rev. Mod. Phys., Vol. 49, No. 1, 1977. The experimental determinations of the hyperfine structure in the alkali atoms are reviewed. 10. R. Gupta, Am. J. Phys. 59(10), 874 (1991). This paper is a Resource Letter that provides a guide to literature on laser spectroscopy. 11. T. W. Hansch, A. L. Schawlow, and G. W. Series, Sci. Am.240(18), 94 (March 1979). An excellent semi-quantitative discussion of laser spectroscopy with many good figures. Introduction Figure 1 compares an older method with a modern method of doing optical spectroscopy. Figure 1a shows the energy levels of atomic deuterium and the allowed transitions for n = 2 and 3, 1b shows the Balmer α emission line of deuterium recorded at 50 K with a spectrograph, where the vertical lines in 1b are the theoretical intensities, and 1c shows an early high resolution laser measurement, where one spectral line is arbitrarily assigned 0 cm-1. The “crossover” resonance line shown in 1c will be discussed later. (See figure 12.3 for more details on the energy levels of hydrogen.) Even at 50 K the emission lines are Doppler-broadened by the random thermal motion of the emitting atoms, while the laser method using a technique known as Doppler-free, saturated absorption spectroscopy eliminates Doppler-broadening. What feature of the laser gives rise to high resolution spectroscopy? Well, it is the narrow spectral linewidth, which is about 20 MHz for the diode laser, and the tunability of lasers that have revolutionized optical spectroscopy. Note that when the 780-nm diode laser is operating at its center frequency then most of its power output is in the frequency range of 3.85 × Laser Spectroscopy LS.5 Spring 2001 output versus forward current and monitor current. The maximum operating forward current and power output are specified by the manufacturer for each diode laser. Figure 6 shows the wavelength versus case temperature. Figure 7 shows how the power output depends on the wavelength, where the spectral width of the laser line at 30 mW is about 20 MHz. The mirrors for a diode laser chip are the cleaved facets of the semiconductor, which are smoother and flatter than any mechanically polished mirror. If there are no coatings on the end surfaces of the laser chip, then the reflectivity R of a surface is given by 2( )c a c a n n R n n − = + (3) where nC and na, are the indices of refraction of the chip and air, respectively. Exercise 1. The index of refraction of GaAs and air are about 3.6 and l.O. What is the reflectivity of the chip facet? Compare your answer with the reflectivity of the mirrors used in the HeNe laser, which is about 0.99. Exercise 2 Assuming the length of the diode laser cavity is 250 µm and the index of refraction of the cavity is 3.6, show that the frequency spacing •ν of the axial modes is 167 GHz. (You may want to refer to the discussion of axial modes under Laser Cavity Modes of the Introduction to Laser Physics.) The LT025MD0 diode laser has a reduced reflective coating on the output facet of the chip; therefore, the reflectivity of this facet is less than that calculated in Exercise 1. The grating reflectivity is about 40%, hence it dominates the front facet of the chip in forming the laser cavity. Therefore, the length of the cavity is the distance from the grating to the far chip facet. The purpose of this configuration is to change the laser wavelength by displacing the grating. Figure 5. Optical power output versus forward current (for three laser temperatures) and monitor current. Figure 6. Emitted wavelength versus case temperature. Note the discontinuities. Laser Spectroscopy LS.6 Spring 2001 Atomic Structure of Rubidium The ground electron configuration of Rb is: ls2; 2s2, 2p6; 3s2, 3p6, 3d10; 4s2, 4p6; 5s1, and with its single 5s1 electron outside of closed shells it has an energy-level structure that resembles hydrogen. For Rb in its first excited state the single electron becomes a 5p1 electron. Also natural rubidium has two isotopes, the 28% abundant 87Rb, where the nuclear spin quantum number I = 3/2, and the 72% abundant 85Rb, where I = 5/2. Term States A term state is a state specified by the angular momenta quantum numbers s, l, and j (or S, L, and J), and the notation for such a state is 2s + 1lj (or 2s + 1Lj). The spectroscopic notation for l values is l = 0(S), 1(P), 2(D), 3(F), 4(G), 5(H),.. . The total angular momentum J is defined by J = L + S (Js) (4) where their magnitudes are ( 1); ( 1); ( 1)J j j L S s s= + = + = += = A A = (5) and the possible values of the total angular momentum quantum number j are |l - s|, |l - s| + 1,..., l + s - 1, l + s; where for a single electron s = 1/2. It is recommended that you read the discussion on LS coupling and term states in Experiment 16. The 5s1 electron gives rise to a 52S1/2 ground term state. The first excited term state corresponds to the single electron becoming a 5p1 electron, and there are two term states, the 52P1/2 and the 5 2P3/2 Exercise 3 Show for a 5s1 electron that the term state is a 52S1/2. For a 5p 1 electron show that the term states are 52P1/2 and 5 2P3/2. Figure 7. Power outuput dependence of wavelength. Laser Spectroscopy LS.7 Spring 2001 Hamiltonian Assuming an infinitely massive nucleus, the nonrelativistic Hamiltonian for an atom having a single electron is given by 22 0 2 ( ) L S J I 2 4 3 3(I J) (I J) ( 1) ( 1) ( ) 2 (2 1) (2 1) 2 effZ ep H r m r I I J J J I I J J ζ α ε β = − + ⋅ + ⋅ + π  ⋅ + ⋅ − + + − −   (6) We label the 5 terms is this equation, in order, as K, V, Hso, H1,hyp, and H2,hyp respectively. K is the kinetic energy of the single electron; where p i= ∇= . Classically p is the mechanical momentum of the electron of mass m. V is the Coulomb interaction of the single electron with the nucleus and the core electrons (this assumes the nucleus and core electrons form a spherical symmetric potential with charge Zeffe, where Zeff is an effective atomic number). Hso is the spin orbit interaction, where L and S are the orbital and spin angular momenta of the single electron. H1,hyp is the magnetic hyperfine interaction, where J and I are the total electron and nuclear angular momenta, respectively. This interaction is -µn•Be where µn, the nuclear magnetic dipole moment, is proportional to I, and Be, the magnetic field produced at the nucleus by the single electron, is proportional to J. Hence the interaction is expressed as αI•J. α is called the magnetic hyperfine structure constant, and it has units of energy, that is, the angular momenta I and J are dimensionless. H2,hyp is the electric quadrupole hyperfine interaction, where ß is the electric quadrupole interaction constant, and nonbold I and J are angular momenta quantum numbers. The major electric pole of the Rb nucleus is the spherical symmetric electric monopole, which gives rise to the Coulomb interaction; however, it also has an electric quadrupole moment (but not an electric dipole moment). The electrostatic interaction of the single electron with the nuclear electric quadrupole moment is -eVq, that is, it is the product of the electron's charge and the electrostatic quadrupole potential. Although it is not at all obvious, this interaction can be expressed in terms of I and J. In both hyperfine interactions I and J are dimensionless, that is, the constants α and ß have units of joules. We will not use the above Hamiltonian in the time independent Schrodinger equation and solve for the eigenvalues or quantum states of Rb, but rather we present a qualitative discussion of how each interaction effects such states. K + V The K + V interactions separate the 5s ground configuration and the 5p excited configuration. This is shown in figure 8a. Qualitatively, if the potential energy is not a strictly Coulomb potential energy then for a given value of n, electrons with higher l have a higher orbital angular momentum (a more positive kinetic energy) and on the average are farther from the nucleus (a less negative Coulomb potential energy), hence higher l value means a higher (more positive) energy. This scenario does not occur in hydrogen because the potential energy is coulombic. Laser Spectroscopy LS.10 Spring 2001 Exercise 6 Assuming all of the spectral lines are resolved for transitions from the 52P3/2 excited state to the 52S1/2 ground state, how many spectral lines do you expect to observe for 87Rb? In addition to the normal resonance lines, there are “crossover” resonances peculiar to saturated absorption spectroscopy, which occur at frequencies (ν1 + ν2)/2 for each pair of true or normal transitions at frequency ν1 and ν2. A crossover resonance is indicated in figure Ic. The crossover transitions are often more intense than the normal transitions. In figure 9 six crossover transitions, b, d, e, h, j, and k, and six normal transitions, a, c, f, g, i, and 1, are shown, where for the normal transitions •F = 0, ±1. The frequency of the emitted radiation increases from a to l. What are the expected frequencies of the normal transitions a, c, f, g, i, and l? To answer this question we first determine the energies of the hyperfine levels. Using equation 9 and forming the dot product of F•F, we solve for J•I and obtain 2 2 2( ) / 2 [ ( 1) ( 1) ( 1)] / 2 / 2 J I F J I F F J J I I C • = − − = + − + − + = (11) where dimensionless magnitudes were used in the second equality and the last equality defines C. Replacing J•I in the hyperfine interactions of equation 6 with equation 11, the magnitude of the interactions or the energy EJ,F is given by ,J F J hypE E E= + 23 3 ( 1) ( 1) 4 4 2 2 (2 1) (2 1)J C C I I J J C E I I J J α β + − + + = + + − − (J) (12) where EJ is the energy of the n 2S+1LJ state, that is, the 5 2P3/2 or the 5 2S1/2 state shown in figure 9. From figure 9 note that in equation 12 for the 52P3/2 state I = 3/2, J = 3/2, and F' = 0,1,2,3; and for the 52S1/2 state I = 3/2, J = 1/2, and F =1,2. (drawing) (fig 9) The frequencies νJ,F (energy/h) of the various hyperfine levels are obtained by dividing equation 12 by Planck's constant h: , 3 ( 1) ( 1) ( 1) 4 2 2 (2 1) (2 1)J F J C C I I J J C A B I I J J ν ν  + − + +  = + + − − (Hz) (13) Laser Spectroscopy LS.11 Spring 2001 where A ≡ α/h and B ≡ ß/h have units of hertz. For the 52S1/2 state of 87Rb, the term that multiplies B in equation 13 reduces to zero and the accepted value of A is 3417.34 Mhz. For the 52P3/2 state of 87Rb, the accepted values of A and B are 84.85 MHz and 12.52 MHz, respectively. For the 52S1/2 of 85Rb the accepted value of A is 1011.91 MHz, and for the 52P3/2 the accepted values of A and B are 25.01 MHz and 25.9 MHz, respectively. Exercise 7 For the 52S1/2 state of 87Rb substitute numerical values for I, J, and F into the third term of equation 13 and show for both F = 1 and 2 that this formula only makes sense if B = 0 for these two cases Exercise 8 For 87Rb use equation 13 to show that: (1) for the 52S1/2 state the splitting of the F = 1 and F = 2 levels, ν1/2,2 - ν1/2,1 is 6834.7 MHz, as shown in figure 8c, (2) for the 5 2P3/2 state the splitting Figure 9. Saturated absorption transitions for 87Rb. The spectral line separation will be derived in exercises 6 through 8 or can be figured out from figure 8. Laser Spectroscopy LS.12 Spring 2001 of the F' = 3 and F' = 2 levels, ν3/2,3 - ν3/2,2 is 267.1 MHz, as shown in figure 8c, (3) the frequency spacing •ν = ν3/2,2 - νJ=3/2 is 194.0 MHz as shown in figure 9. (4) Now that you know how to do the calculations using equation 13, just use the energy level spacings given in figures 8 and 9 to shown the frequency of transitions a and b are 3.846 × 1014 - 2.7933 × 109 Hz and 3.846 × 1014 - 2.7147 × 109 Hz, respectively, hence the separation of these two spectral lines is 79 MHz. (5) Show that the frequency separation of spectral lines a and I is 6.993 GHz and then show that their wavelength separation is 0.0142 nm. One goal of this experiment is to experimentally determine A and B for the 52P3/2 state of 87Rb. Doppler Broadening and Absorption Spectroscopy Random thermal motion of atoms or molecules creates a Doppler shift in the emitted or absorbed radiation. The spectral lines of such atoms or molecules are said to be Doppler broadened since the frequency of the radiation emitted or absorbed depends on the atomic velocities. (The emission spectral lines in both Experiments 12 and 13 will be Dopplerbroadened.) Individual spectral lines may not be resolved due to Doppler broadening, and, hence, subtle details in the atomic or molecular structure are not revealed. What determines the linewidth of a Doppler broadened line? To answer this question we do some theoretical physics. We first consider the Doppler effect qualitatively. If an atom is moving toward or away from a laser light source, then it “sees” radiation that is blue or red shifted, respectively. If an atom at rest, relative to the laser, absorbs radiation of frequency ν0, then when the atom is approaching the laser it will see blue-shifted radiation, hence for absorption to occur the frequency of the laser must be less than ν0 in order for it to be blue-shifted to the resonance value of ν0. Similarly, if the atom is receding from the laser, the laser frequency must be greater than ν0 for absorption to occur. We now offer a more quantitative argument of the Doppler effect and atomic resonance, where, as before, ν0 is the atomic resonance frequency when the atom is at rest in the frame of the laser. If the atom is moving along the z-axis, say, relative to the laser with vZ«c, then the frequency of the absorbed radiation in the rest frame of the laser will be νL, where 0 1 Z L v C ν ν  = +   . (Hz) (14) If vZ is negative (motion toward the laser) then νL < ν0, that is, the atom moving toward the laser observes radiation that is blue-shifted from νL up to νL. If vZ is positive (motion away from the laser) then νL > ν0, that is, the atom observes radiation that is red-shifted from νL down to ν0. Therefore, an ensemble of atoms having a distribution of speeds will absorb light over a range of frequencies. The probability that an atom has a velocity between vZ and vZ + dvZ is given by the Maxwell distribution Laser Spectroscopy LS.15 Spring 2001 Doppler-Free Saturated Absorption Spectroscopy The apparatus for the Doppler-free saturated absorption spectroscopy of Rb is shown in figure 11. The output beam from the laser is split into three beams, two less intense probe beams and a more intense pump beam, at the beamsplitter BS. The two probe beams pass through the Rb cell from left to right, and they are separately detected by two photodiodes. After being reflected twice by mirrors M1 and M2, the more intense pump beam passes through the Rb cell from right to left. Inside the Rb cell there is a region of space where the pump and a probe beam overlap and, hence, interact with the same atoms. This overlapping probe beam will be referred to as the first probe beam and the other one the second probe beam. The signal from the second probe beam will be a linear, absorption spectroscopy signal, where the spectral lines will be Doppler-broadened. The signal is shown in figure 12a, and it was photographed from the screen of an oscilloscope. This signal was obtained by blocking the pump and first probe beams. There are two Doppler-broadened lines shown in the 12a, and a portion of the triangular waveform that drives the PZT, and hence, sweeps the laser frequency, is also shown. The larger amplitude signal is that of the 72% abundant 85Rb and the smaller amplitude signal is that of the 28% abundant 87Rb. The 87Rb transition is the F = 2 to F’ = 1, 2 and 3 transition, and the 85Rb transition is F = 3 to F’ = 2, 3, and 4. Figure 11. Apparatus for Doppler-free saturated absorption spectroscopy of 87Rb. Laser Spectroscopy LS.16 Spring 2001 If the second probe beam only is blocked then the signal from the first probe beam will be a nonlinear, saturated absorption spectroscopy signal “riding on” the Dopplerbroadened line. This signal is shown in figure 12b, where the two Doppler-broadened lines are the same transitions as in 12a, but note the hyperfine structure riding on these lines. If the two signals in 12a and 12b are subtracted from each other, then the Dopplerbroadened line cancels and the hyperfine structure remains. The two photodiodes shown in figure 11 are wired such that their signals subtract, and the signal obtained when none of the beams are blocked is shown in figure 12c for 87Rb. In 12c note that the amplitude of the triangular waveform is one-fourth as large as in 12a and b, that is, in 12c the laser is being swept over a range that is one-fourth as large as in 12a and b. The signal shown in 12c is the Doppler- free saturated absorption signal. We now consider in detail the physics behind figure 12. Figure 12. In all three photographs the frequency is increasing to the left. (a) Doppler- broadened spectral lines. (b) Doppler-broadened spectral lines with hyperfine structure. (c) Doppler-free saturated absorption spectral lines, where the spectral lines are labeled consistent with figure 9. Laser Spectroscopy LS.17 Spring 2001 We start by focussing on the first probe and pump beams. The pump beam changes the populations of the atomic states and the probe detects these changes. Let us first consider how the pump beam changes the populations, and then we will discuss how these changes effect the first probe signal. As discussed above, because of the Doppler shift only atoms with a particular velocity vZ will be in resonance with the pump beam, and thereby be excited. This velocity dependent excitation process changes the populations in two ways, one way is known as “hyperfine pumping” and the other as “saturation”. Hyperfine pumping is the larger of the two effects, and it will be discussed first. Hyperfine pumping is optical pumping of the atoms between the hyperfine levels of the 52S1/2 state. This happens in the following manner. Suppose the laser frequency is such that an atom in the F = 1 ground state is excited to the F' = 1 excited state. The •F selection rule indicates that this state can then decay back to either the F = 1 or F = 2 ground states, with roughly equal probabilities. When it decays back to the F = 1 state it will be reexcited by the laser light and the process repeated. Thus after a very short time interval most of the atoms will be in the F = 2 state, and only a small fraction will remain in the F = 1 state. If the atoms never left the pump laser beam, even a very weak laser would quickly pump all the atoms into the F = 2 state. However, the laser beam diameter is small and the atoms are moving rapidly so that in a few microseconds the optically pumped atoms leave the beam and are replaced by unpumped atoms whose populations are equally distributed between the F = 1 and F = 2 levels. ( You are ask to show below, in Exercise 12, that the populations of these two levels are essentially equal at room temperature with the laser turned off.) The average populations are determined by the balance between the rate at which the atoms are being excited and hence optically pumped, and the rate they are leaving the beam to be replaced by fresh ones. Without solving the problem in detail, one can see that if the laser intensity is sufficient to excite an atom in something like 1 microsecond it will cause a significant change in the populations of the F = 1 and F = 2 levels, that is, more atoms will be in the F = 2 level than the F = 1 level. This mechanism is called hyperfine pumping since the net effect is pumping electrons from the F = 1 ground hyperfine level to the F = 2 excited hyperfine level of the 52S1/2 state. Although the example used was for the F = 1 to F' = 1 transition, similar hyperfine pumping will occur for any excitation where the excited state can decay back into a ground state which is different from the initial ground state. Exercise 12 With the laser off, the ratio of the number of atoms in the energy level E1/2,2 (J = 1/2, F = 2), N(E1/2,2), to the number of atoms in the energy level E1/2,1 (J = 1/2, F = 1), N(E1/2,1), is determined by Maxwell-Boltzmann statistics: 1 / 2,2 1 / 2,2 1 / 2,1 1 / 2,1 ( ) exp ( ) N E E E N E kT −  = −     (20) Assuming a temperature of 295 K, show that the above population ratio is 0.999, hence the two levels are essentially equally populated with the laser off. You may want to refer to Exercise 4 and the discussion that precedes it in the Introduction to Magnetic Resonance. Laser Spectroscopy LS.20 Spring 2001 Figure 13. Absorption of pump and probe beams by ground state atoms, assuming a Maxwell velocity distribution, for the cases when the laser frequency νL is (a) νL < ν0, (b) νL = ν0, and (c) νL > ν0. Laser Spectroscopy LS.21 Spring 2001 A crossover peak appears midway between any two transitions that have the same lower level and two different excited levels. This occurs because, when the laser is tuned to the frequency midway between two such transitions, atoms with a particular nonzero velocity can simultaneously be in resonance with both the saturating beam and the probe beam and thus have nonlinear absorption. This happens because the two beams excite resonances to different transitions. For example, those atoms which are moving toward the probe beam with a velocity which gives a Doppler shift exactly equal to half the frequency difference between the two transitions will be shifted into resonance with the higher frequency transition. These atoms will see the pump beam with exactly the opposite Doppler shift. This shift will make the pump beam frequency just right to excite these same atoms to the lower frequency transition. As a result, the absorption is saturated not in stationary atoms but rather in two classes of moving ones. Experiments Remark: The diode laser used in this experiment is a research-quality instrument, and it should be handled with considerable care. It can be destroyed by a spurious voltage spike, for example, by a voltage induced in the current controller from other equipment. It is recommended that other equipment being used in this experiment be turned on and allowed to stabilize before the laser is turned on. Before discussing how to use the laser we consider the diode laser assembly shown in figure 14, which is reproduced from reference 1. The labeled parts in figure 14 are: T-thermistor, LD-laser diode mount, C-collimating lens mount, G-diffraction grating, S-precision screw, P-PZT disk, MM-mirror mount, B- baseplate, H-controls horizontal motion of the grating, V-controls vertical motion of the grating, W-window. Lift the top off of the laser assembly and observe the parts. Also reference marks on the H and V knobs are desirable, for example, a white dot made with typewriter correction fluid on the top of each knob. The discussion below assumes that the temperatures of the laser and the baseplate, the operating current, the position of the screw S, and the positions of the adjustments H and V have been made as described in reference I such that the laser wavelength is approximately 780 nm. The operating current and the knob settings for the temperature controllers should be recorded on or near the laser housing. Also the maximum laser current should be recorded. The following adjustments SHOULD NOT BE ALTERED: 1 The precision screw S. 2. The vertical motion control V. 3 Baseplate or laser diode temperature. The following should only be adjusted as specified later in the experiment: 4. Diode current. 5. The horizontal motion control H. Turn off the laser in the following way, where the knobs and switches are on the current controller: 1. Set the laser current to zero using the current control knob. 2 Set the “run-short” switch in the short position. 3. Leave the power switch of the current controller in the on position. Laser Spectroscopy LS.22 Spring 2001 To turn on the laser go through the three steps in reverse order, where the laser current should be set at the previously established value. NEVER EXCEED THE MAXIMUM CURRENT. Place an IR detection card in the beam path as you increase the laser current in order to observe the beam. Diode Laser Power Output vs. Forward Current The purpose of this measurement is to better understand the diode laser by measuring uncalibrated power output vs forward current. Wire the current to voltage amplifier circuit shown in figure 15 or use a box which contains this circuit already wired. The OP913 SL is a PIN silicon photodiode with a large active area chip. Other photodiodes may be used in its place if available. Place the OP913 between M1 and M2 shown in figure 11 such that it intercepts the pump beam. Measure the output voltage from the amplifier as a function of the diode laser forward current, but do not exceed the specified operating current. Plot your data, where the resulting curve should resemble those shown in figure 5. After finishing these measurements set the current at the operating current. Remark: If the laser light saturates the photodiode, then you can reduce the laser intensity by inserting pieces of darkened plexiglass in the beam. By using several pieces, you can measure the transmission of a single piece. Exercise 12 What is the threshold current for your laser? Remove the OP 913 photodiode from the path of the pump beam. Figure 14. Assembly top view of laser. Figure 15. Circuit to measure uncalibrated diode laser power output as a function of diode laser current. The switch should be a shorting switch. Laser Spectroscopy LS.25 Spring 2001 whereν and c are the wave frequency and velocity. The frequency is not fixed, rather it is being swept by the triangle waveform applied to the PZT, hence as the frequency changes the phase difference will change. The changes in frequency and phase are expressed by writing equation 21 as 1 2 1 2( ) 2 ( ) 2L LC νφ φ ∆∆ − = − π (22) The intensity of the two superimposed waves, aside from a constant of proportionality, is given by, 1 2 1 2 2 2 ( 2 2 ( 2 2 ) 1 2 2 2 ( 2 2 ) ( 2 2 ) 1 2 2 2 1 2 1 2 1 2 ( ) ( ) 2 cos ( ) L L i t i t L L i t i t I E E E e E e E e E e E E E E π πν π πν λ λ π πν π πν λ λ φ φ − −∗ − − = • = + • + = + + • ∆ − (23) where • was inserted to indicate the change in the phase from sweeping the frequency. The interference is a maximum whenever 1 2( ) 2φ φ∆ − = π (24) Substituting equation 24 into 22 yields the frequency spacing of the interference maxima 1 22 ( ) C L L ν∆ = − (Hz) (25) Figure 17 is a photo of the scope screen showing the triangle waveform and the interference pattern, where the frequency spacing between maxima is given by equation 25. There are a few things it is helpful to know in order to get good fringes from the interferometer. First, you should not be misled by the weak fringes you will get from the interference of the beams reflecting off the two sides of the beamsplitter. Figure 17. Interference fringes and the triangle wave sweep. In this case: L1 - L2 = 19.5 cm, and •ν = 0.769 GHz. Laser Spectroscopy LS.26 Spring 2001 They will not have the correct spacing and they will not require both beams returning off M3 and M4. Second, you need to realize that in order to give good fringes the two beams must not only overlap at the photodiode, but they must also be parallel. It is possible to have them come back from M3 and M4 and go though different parts of the beamsplitter but still overlap at the photodiode. By drawing some simple triangles you should be able to convince yourself that in this case the path length difference L1 - L2 will be different at different places where the beams overlap. This will result in a series of bright and dark fringes across the photodiode which you can see if you look closely with the viewer. If the phase difference across the face of the photodiode is a full 2• there will always be a bright and dark fringe present and changing the laser frequency will give no change in the total power on the photodiode, and hence no signal. The larger the angle between the beams the closer will be the spacing of the fringes. As you adjust the beams to be parallel (but still overlapping) the spacing between the dark fringes will become larger until it is as large as the photodiode, and you will see a large modulation in the photodiode output as you change the frequency. Often the easiest way to get good fringes is to first make the beams as parallel and overlapping as possible and then do the final adjustments by looking at the photodiode output and align the beams to get the largest fringes as you ramp the laser frequency. The final thing you need to realize is that you want the beams to go nearly but not exactly back along the incident laser beam. If they are going exactly back they will go back into the laser and cause the frequency to jump around. This usually shows up as a lot of noise on the fringe pattern signal you see on the oscilloscope. If you keep these things in mind you should not have trouble aligning and using the Michelson interferometer with the following procedure. Choose L1 and L2 such that •ν is approximately 0.5 GHz, and start with the amplifier gain set to 100 k•. 1. Place an IR detection card over PD and adjust the tilts of M3 and M4 such that the two beams are superimposed. 2. Place the IR detection card on M1 near the beam spot incident from the laser, adjust M3 and M4 such that the two spots coming from BS2 are superimposed on each other but slightly offset from the beam spot incident from the laser. Check the superposition of the two spots at PD, re-align if necessary. Once alignment is achieved the scope trace should resemble figure 17. Carefully calibrate the horizontal axis of the oscilloscope for the gain setting of the triangle wave generator you used in measuring the spectra. If you determine how large a frequency change is produced for each volt that is applied to the PZT drive, you can then convert the horizontal scale of the spectra from ramp voltage to frequency, and then determine splittings and linewidths. You might check that different ranges on the ramp generator give the same calibration factor. As the laser frequency is changed you will observe a cosine modulation on the oscilloscope. However, if you make a large sweep you will see sudden jumps in the signal as if the phase has abruptly changed. What has actually happened is that the laser has jumped to a different frequency (a so called “modehop”), probably one which corresponds to one more or one less half wavelength fitting into the laser cavity as defined by the diffraction grating “mirror” and the high reflecting facet of the diode chip, and hence is different by about 3 Ghz. The frequency where the laser jumps will move if you change the laser current. It could also be jumping to a frequency which corresponds to one more or one less half wavelength fitting into the actual diode chip itself. This causes a much larger change in the frequency. Under ideal conditions with this laser setup you can get a scan of about 8 GHz without a modehop. A single scan of this Laser Spectroscopy LS.27 Spring 2001 length, which shows all 4 rubidium peaks at once, is shown in fig. 18. More typically you will get continuous scans (no mode hops) of 3 or 4 Ghz. If the frequency range over which you get a continuous scan is shorter than this it probably means that the vertical alignment is off. The length of continuous scan can also be affected by the laser temperature. Data Analysis 1. Measure the FWHM for each of the two Doppler-broadened lines of 87Rb. Compare your values with your calculated value from Exercise 9. 2. Measure the separation of the hyperfine lines for the F = 2 to F' transitions, use equation 13 to obtain a theoretical expression for the frequency separation of the hyperfine lines, and then solve for the constants A and B for the 52P3/2 state of 87Rb. Compare your results with the accepted values given earlier. If you have time, carry out the same analysis for the other transitions.
Docsity logo



Copyright © 2024 Ladybird Srl - Via Leonardo da Vinci 16, 10126, Torino, Italy - VAT 10816460017 - All rights reserved