Doppler-Free Spetroscopy of Rubidium

Similar documents
Laser Locking with Doppler-free Saturated Absorption Spectroscopy

DIODE LASER SPECTROSCOPY (160309)

R. J. Jones Optical Sciences OPTI 511L Fall 2017

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Electrical Engineering and Computer Science

Lecture 6 Fiber Optical Communication Lecture 6, Slide 1

PHY 451: Advanced Laboratory Manual for Diode Laser Spectroscopy. Derek Neben, Lennart Dabelow

Spectrometer using a tunable diode laser

Basic concepts. Optical Sources (b) Optical Sources (a) Requirements for light sources (b) Requirements for light sources (a)

LEDs, Photodetectors and Solar Cells

Figure 1. Schematic diagram of a Fabry-Perot laser.

Ph 77 ADVANCED PHYSICS LABORATORY ATOMIC AND OPTICAL PHYSICS

Luminous Equivalent of Radiation

Review of Semiconductor Physics

Multi-photon Absorption in Optical Pumping of Rubidium

A Narrow-Band Tunable Diode Laser System with Grating Feedback

Fabry-Perot Cavity FP1-A INSTRUCTOR S MANUAL

Experimental Physics. Experiment C & D: Pulsed Laser & Dye Laser. Course: FY12. Project: The Pulsed Laser. Done by: Wael Al-Assadi & Irvin Mangwiza

The Saturated Absorption Spectroscopy Lab

Chapter 3 OPTICAL SOURCES AND DETECTORS

6 Experiment II: Law of Reflection

Optodevice Data Book ODE I. Rev.9 Mar Opnext Japan, Inc.

High-power semiconductor lasers for applications requiring GHz linewidth source

Introduction Fundamentals of laser Types of lasers Semiconductor lasers

Wavelength Control and Locking with Sub-MHz Precision

NORTHWESTERN UNIVERSITY MULTI-SPECTRAL RAMAN GAIN IN DUAL-ISOTOPE RUBIDIUM VAPOR A THESIS SUBMITTED TO THE GRADUATE SCHOOL

CHAPTER 5 FINE-TUNING OF AN ECDL WITH AN INTRACAVITY LIQUID CRYSTAL ELEMENT

Applications of Steady-state Multichannel Spectroscopy in the Visible and NIR Spectral Region

B. Cavity-Enhanced Absorption Spectroscopy (CEAS)

Spectroscopy in the UV and Visible: Instrumentation. Spectroscopy in the UV and Visible: Instrumentation

SUPPLEMENTARY INFORMATION

Period 3 Solutions: Electromagnetic Waves Radiant Energy II

Measurement of the Speed of Light in Air

LASER Transmitters 1 OBJECTIVE 2 PRE-LAB

UNIT-III SOURCES AND DETECTORS. According to the shape of the band gap as a function of the momentum, semiconductors are classified as

R. J. Jones College of Optical Sciences OPTI 511L Fall 2017

Spectrophotometer. An instrument used to make absorbance, transmittance or emission measurements is known as a spectrophotometer :

Pound-Drever-Hall Locking of a Chip External Cavity Laser to a High-Finesse Cavity Using Vescent Photonics Lasers & Locking Electronics

Lecture 4 Fiber Optical Communication Lecture 4, Slide 1

Frequency evaluation of collimated blue light generated by wave mixing in Rb vapour

Powerful Single-Frequency Laser System based on a Cu-laser pumped Dye Laser

ECE 340 Lecture 29 : LEDs and Lasers Class Outline:

USING LASER DIODE INSTABILITIES FOR CHIP- SCALE STABLE FREQUENCY REFERENCES

Semiconductor Lasers Semiconductors were originally pumped by lasers or e-beams First diode types developed in 1962: Create a pn junction in

Key Questions. What is an LED and how does it work? How does a laser work? How does a semiconductor laser work? ECE 340 Lecture 29 : LEDs and Lasers

SECOND HARMONIC GENERATION AND Q-SWITCHING

Semiconductor Optical Communication Components and Devices Lecture 18: Introduction to Diode Lasers - I

Photoassociative Spectroscopy of Strontium Along the 1 S 0-3 P 1. Transition using a Littman/Metcalf Laser. Andrew Traverso. T.C.

CONTENTS. Chapter 1 Wave Nature of Light 19

LAB V. LIGHT EMITTING DIODES

Spectroscopy Lab 2. Reading Your text books. Look under spectra, spectrometer, diffraction.

A continuous-wave Raman silicon laser

Examination Optoelectronic Communication Technology. April 11, Name: Student ID number: OCT1 1: OCT 2: OCT 3: OCT 4: Total: Grade:

Spectroscopy of Ruby Fluorescence Physics Advanced Physics Lab - Summer 2018 Don Heiman, Northeastern University, 1/12/2018

Swept Wavelength Testing:

CHAPTER 7. Components of Optical Instruments

Physics of Waveguide Photodetectors with Integrated Amplification

Light for Ultra Cold Molecules Final Report for PHYS349

White Paper Laser Sources For Optical Transceivers. Giacomo Losio ProLabs Head of Technology

Functional Materials. Optoelectronic devices

Direct frequency comb saturation spectroscopy with an ultradense tooth spacing of 100 Hz D. A. Long, 1,* A. J. Fleisher, 1 and J. T.

FIBER OPTICS. Prof. R.K. Shevgaonkar. Department of Electrical Engineering. Indian Institute of Technology, Bombay. Lecture: 18.

Ultra stable laser sources based on molecular acetylene

Chemistry Instrumental Analysis Lecture 10. Chem 4631

High-frequency tuning of high-powered DFB MOPA system with diffraction limited power up to 1.5W

Characterization and Development of an Extended Cavity Tunable Laser Diode

Chemistry 524--"Hour Exam"--Keiderling Mar. 19, pm SES

Lecture 18: Photodetectors

External Cavity Diode Laser for Ultra-cold Atom Experiments

A transportable optical frequency comb based on a mode-locked fibre laser

Copyright 2006 Crosslight Software Inc. Analysis of Resonant-Cavity Light-Emitting Diodes

Components of Optical Instruments. Chapter 7_III UV, Visible and IR Instruments

Semiconductor Lasers Semiconductors were originally pumped by lasers or e-beams First diode types developed in 1962: Create a pn junction in

Temporal coherence characteristics of a superluminescent diode system with an optical feedback mechanism

Nd:YSO resonator array Transmission spectrum (a. u.) Supplementary Figure 1. An array of nano-beam resonators fabricated in Nd:YSO.

RECENTLY, studies have begun that are designed to meet

PERFORMANCE OF PHOTODIGM S DBR SEMICONDUCTOR LASERS FOR PICOSECOND AND NANOSECOND PULSING APPLICATIONS

1 Semiconductor-Photon Interaction

Supplementary Information:

Zeeman Shifted Modulation Transfer Spectroscopy in Atomic Cesium

Observation of Rb Two-Photon Absorption Directly Excited by an. Erbium-Fiber-Laser-Based Optical Frequency. Comb via Spectral Control

Electron Spin Resonance v2.0

Exercise 8: Interference and diffraction

LASER DIODE MODULATION AND NOISE

Wave optics and interferometry

Lasers PH 645/ OSE 645/ EE 613 Summer 2010 Section 1: T/Th 2:45-4:45 PM Engineering Building 240

Transfer Cavity Stabilization Using the Pound-Drever-Hall Technique with Noise Cancellation

visibility values: 1) V1=0.5 2) V2=0.9 3) V3=0.99 b) In the three cases considered, what are the values of FSR (Free Spectral Range) and

A novel tunable diode laser using volume holographic gratings

VERTICAL CAVITY SURFACE EMITTING LASER

Lecture 4 INTEGRATED PHOTONICS

Lecture 9 External Modulators and Detectors

LAB V. LIGHT EMITTING DIODES

Fundamentals of Laser

3550 Aberdeen Ave SE, Kirtland AFB, NM 87117, USA ABSTRACT 1. INTRODUCTION

Photon Count. for Brainies.

Coherent Receivers Principles Downconversion

Dr. Rüdiger Paschotta RP Photonics Consulting GmbH. Competence Area: Fiber Devices

atom physics seminar ultra short laser pulses

Chapter 14. Tunable Dye Lasers. Presented by. Mokter Mahmud Chowdhury ID no.:

EE119 Introduction to Optical Engineering Spring 2003 Final Exam. Name:

Transcription:

Doppler-Free Spetroscopy of Rubidium Pranjal Vachaspati, Sabrina Pasterski MIT Department of Physics (Dated: April 17, 2013) We present a technique for spectroscopy of rubidium that eliminates doppler broadening. Using this, we probe the hyperfine structure of rubidium and determine the hyperfine coupling constants. We find for 85 Rb a coupling constant of 973 ± 21MHz in the 5S 1/2 state and constants of A = 24 ± 2MHz, B = 22 ± 2 for the 5P 3/2 state. We find for 87 Rb a coupling constant of 3260 ± 60MHz in the 5S 1/2 state and constants of A = 83 ± 8MHz, B = 14 ± 1 for the 5P 3/2 state. I. INTRODUCTION Spectroscopy is a powerful technique that is widely used for an enormous range of purposes. Here we present an application of atomic spectroscopy to verify results from theoretical quantum mechanics. Theoretical knowledge of the structure of the atom predicts a series of refinements to the energy spectrum of a hydrogenlike atom such as rubidium. Naïve spectroscopy fails to probe this spectrum to a satisfactory degree due to doppler broadening effects, so a hole-burning technique is applied to eliminate those effects. By doing so, we can confirm quantum mechanical results and measure quantitative properties of the rubidium atom. II. THEORY ν transition = ( ) 1 + v c ν, removing them from the ground state. At the same time, a weak probe beam with the same frequency is sent through the cell in the opposite direction, exciting atoms with velocity such that ν transition = ( ) 1 v c ν. When the laser is close to, but not very close to the transition frequency, the probe and pump excite different groups of atoms, and the probe beam shows a normal doppler-broadened spectrum. However, when the laser is tuned very close to the transition frequency, the pump beam depopulates the state that the probe beam tries to excite. This creates a Lamb dip in the absorption spectrum, as there are no ground-state atoms left for the probe beam to excite. Measuring the size and location of the Lamb dip allows more precise determination of the absorption spectrum. II.1. Doppler-Free Absorption Spectroscopy II.2. Rubidium Spectrum Absorption spectroscopy provides a method for determining transitions between energy levels of a material. A quantum system can absorb only excitations that allow it to transition from one energy eigenstate to another. By exciting the system with a range of energies, we can detect which are absorbed, determining the spacings in the energy levels of the material. In this case, we excite a gas of rubidium atoms with a near-infrared laser. Because the gas has a wide velocity distribution, particles in the gas experience a range of doppler shifts, so a sharp laser line appears as a broad line to the gas, obscuring certain spectral features. To remove this doppler broadening, a technique known as hole-burning is used. A strong pump beam with frequency ν is sent through the cell in one direction, exciting atoms with velocity such that pranjal@mit.edu, fermilab@mit.edu Rubidium has one valence electron, so its electron energy levels resemble hydrogen[5]. As with hydrogen, the roughest approximation of the spectrum ignores spin and relativistic effects, specifying the n = 1, 2,... states. Here, we do not probe transitions between these states. Instead, we examine first and second order perturbations to the ground state. The first perturbation (the fine structure) comes from spin-orbit interactions. This splits the ground state into states that depend on the electron angular wavefunction. The ground state is specified in Russell-Saunders notation[3] as 5 2 S 1/2, and the first two excited states are 5 2 P 1/2 and 5 2 P 3/2. The transition we probe here is between the ground state and the 5 2 P 3/2, which occurs at 780 nm. This transition is further perturbed by the nuclear spin, causing hyperfine splitting. 87 Rb has a nuclear spin of I = 3/2 and 85 Rb has a nuclear spin of I = 5/2. This splits the ground state into two levels with total angular momentum F = 1 and F = 2 in the case of 85 Rb, and

between two transitions, the pump beam will depopulate the ground state into one excited state, while the probe beam will attempt to excite into a different excited state. This still creates a Lamb dip, since the ground state is depopulated even though the pump and probe are exciting slightly different transitions. 2 II.3. Resonating Cavities FIG. 1. Nuclear spin effects cause hyperfine splitting, which splits each electron angular wavefunction into separate energy levels[4]. with F = 2 and F = 3 in the case of 87 Rb. It also splits the excited state into four levels, with momenta F = 0, 1, 2, 3 in the case of 85 Rb, and F = 1, 2, 3, 4 in the case of 87 Rb. A selection rule limits transitions to a change in angular momentum of at most 1, so there are three transitions between each ground state splitting and the excited state. The hyperfine splittings for the ground state are 10 to 100 times larger than for the excited state, so this gives four clusters (two per isotope) of three transitions each, for a total of twelve peaks in the spectrum. The transition frequencies are given by ν F = ν J + A C 2 + B 3C(C+1) 4 I(I + 1)J(J + 11) 2I(2I 1)J(2J 1) (1) where C = F (F + 1) J(J + 1) I(I + 1), ν J is the fine frequency, and A, B are experimentally determined constants. The energy level diagram for the hyperfine splittings of this transition is given in 1. II.2.1. Crossover Peaks Lamb dips can also occur when the laser is not at a transition frequency. In each cluster of transitions from a single ground state, when the laser is halfway Optical resonating cavities are critical to measuring and determining the laser wavelength. An optical cavity consists of two very parallel semi-transparent mirrors between which a standing wave can develop. Such a cavity with length l and refractive index n will amplify only light with wavelength mλ = 2nl, for integer m. Thus, the free spectral range of a cavity, given as the spacing between adjacent peaks, is given by II.4. ν = c 2nL Diode Laser (2) A diode laser is used to excite the rubidium atoms used in this experiment. Diode lasers use semiconductors as the lasing medium. A standard semiconductor diode consists of the interface between a P-doped semiconductor and an N-doped semiconductor. When unbiased, the interface becomes a depletion region as the holes from the P-doped region combine with electrons from the N-doped region to form a layer with no mobile charge carriers but a net electric field. When an external voltage is applied that cancels out this electric field, electrons and holes can diffuse across the junction. At the interface, electrons and holes combine, releasing energy. In an indirect bandgap semiconductor like silicon, this energy change corresponds to a momentum change, and most of the energy is released as heat. However, in a direct bandgap semiconductor like aluminum gallium arsenide, the energy change does not change the momentum, and the combination of the charge carriers releases a photon instead. This creates a light-emitting diode (LED). The LED is turned into a laser with the addition of parallel semi-reflecting surfaces on the sides of the diode. The photons which are reflected back into the semiconductor cause simulated emission when they are absorbed by charge carriers, which recombine and emit two coherent photons. The wavelength of the laser can be changed by changing the temperature and power go-

3 III.3. Balanced Photodetector FIG. 2. The beam is split twice, first to measure frequency changes with the Fabry-Perot interferometer, then into probe, pump, and control beams to measure the spectrum. As seen in Figure 2, the pump beam travels through the cell in the opposite direction as the probe and control beams. The pump beam and the probe beam travel through the same region, allowing the probe beam to measure a spectrum with Lamb dips, while the control beam does not intersect the pump beam, ensuring that the control beam measures only a doppler-broadened spectrum. A balanced photodetector[1] contains two photodiodes that measure the intensity of both the probe and control beam, then subtracts them to determine a spectrum of only Lamb dips, allowing the observation of the hyperfine splitting. Since various factors could shift the relative intensity of the beams, it was sometimes necessary to take successive measurements with the pump beam blocked and unblocked, and subtract them afterwards to get the best possible spectrum. ing to the diode, which change the size and refractive index of the diode. III.4. Fabry-Perot Interferometer III. III.1. EQUIPMENT & PROCEDURE Overview and Beam Path The optical table and beam path are shown in Figure 2. The diode laser passes through an optical isolator, which protects the laser and reduces the likelihood of mode-hopping, and then is split twice. First, a small fraction of the beam is reflected off a glass plate into the Fabry-Perot interferometer (III.4) to measure frequency changes. Then, the beam passes through a prism which refects a small portion of the light twice, to form the pump, probe and control beams. A balanced photodetector (III.3) measures the difference between the probe and control beams to get the spectrum. III.2. Diode Laser Package We use a commercial TUI external cavity diode system with a AlGaAs laser head[2]. An externally mounted diffraction grating is pivoted by a servomotor to act as a wavelength-selective mirror. This allows a frequency sweep of around 10 GHz - approximately 1 nm with a linewidth of around 30 MHz. The relative wavelengths of the laser are determined with a Fabry-Perot interferometer. This is a optical cavity as described in II.3, with two partially reflecting mirrors separated by 47.1 ± 0.7 cm. This distance is measured by placing a ruler in the same plane as the cavity and counting pixels in a digital photograph of the scene. Such a cavity has a free spectral range of 318 ± 5 MHz. Uncertainties in this measurement come from the pixel precision of the camera and from uncertainty in the precise mirror position in the mirror mount, as well as the fact that light might bounce off both sides of the mirror IV. RESULTS A full spectrum showing all four clusters of transitions is shown in Figure 3. The ground state hyperfine splittings can be determined by measuring the distance between the clusters, and the excited state splittings can be determined by measuring the distances between peaks in the individual clusters. A fit to the zero-crossings in the Fabry-Perot output, as shown in Figure 4, allows for interpolation to determine peak frequencies to a high degree of precision. Ideally, the laser would have a linear frequency ramp, but it is significantly better described by a slightly quadratic

4 Rb Spectrum Showing Four Transition Clusters 87 Rb F = 2 Transitions Differential Voltage 0.2 0.15 0.1 0.05 0-0.05 87 Rb 85 Rb 85 Rb 87 Rb Differential Voltage 0.2 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04-0.1-1 0 1 2 3 4 5 6 7 8 9 10 0.02 1.7 1.8 1.9 2.0 2.1 2.2 Frequency (GHz, Relative) Frequency (GHz, Relative) FIG. 3. A full spectrum taken over approximately 8 GHz shows four clusters of transition peaks corresponding to the split ground states of 85 Rb and 87 Rb. From this graph the splitting constants of the 5S 1/2 ground state can be derived. Photodiode Voltage -0.31-0.315-0.32-0.325-0.33-0.335 Fabry-Perot Zero Crossings -0.34 0.2 0.4 0.6 0.8 1 1.2 1.4 Frequency Sweep FIG. 4. The Fabry-Perot outputs a sinusoidal signal that can be converted to a relative frequency measurement by locating and fitting to the zero-crossings. ramp, indicating that the laser scans slightly faster towards one end of the ramp. The peaks are then located by a quadratic fit to the area around the peak. From the full spectrum in Figure 3, we find splittings of 2920 ± 64 MHz for 85 Rb and 6520 ± 120 MHz for 87 Rb. Both these values slightly underestimate the known value of the splittings, which are 3040 MHz and FIG. 5. A slow sweep over a smaller frequency range allows precise measurement of transitions from a single ground state. The polarity of the signal is arbitrary and depends on the configuration of the balanced photodetector. Only the lowest energy transition and crossover peaks are visible in this region. 6830 MHz respectively. Next, the laser is swept more slowly over small regions to determine the splittings of the first excited state. For most of these transitions, the crossover peaks are much more visible than the normal transition peaks, so the frequency differences between the peaks must be doubled to find the actual transition frequencies. This method gives spacings of 262 ± 10 MHz and 151 ± 16 MHz for the two higher 87 Rb frequencies, and spacings of 113 ± 6 MHz, 62 ± 4 MHz for 85 Rb. These agree with the accepted values of 266.7 MHz, 159.6 MHz, 121.0 MHz, and 63.4 MHz. From these spacings and equation 1, we can derive the coupling constants for each of the rubidium states. ν For the 5S 1/2 state, we find A = C F =3 C F =2 = ν 3 = 973 ± 21MHz for 85 Rb and ν 2 = 3260 ± 60MHz for 87 Rb. These both underestimate the accepted values of 1011 MHz and 3417 MHz, which is discussed in V. Analagously, we derive for the 5P 3/2 state A = 83 ± 8MHz, B = 14 ± 1MHz for 87 Rb, and A = 24 ± 2MHz, B = 22 ± 2MHz for 85 Rb. These numbers agree with the accepted values of A = 84.7, B = 12.5 for 87 Rb, and A = 25.0, B = 25.7 for 85 Rb at approximately one standard deviation.

5 V. ERRORS The errors in this measurement are derived from a combination of resolution limitations of the voltage measuring devices and noise in the experiment, as well as statistical error between runs. When measuring the full spectrum, the time resolution of the oscilloscope and scan speed of the laser controller allowed us only 20 MHz resolution, limiting the fidelity with which the peak could be detected. Furthermore, errors in the measurement of the Fabry-Perot length as described in III.4 accumulated over the large wavelength differences to arrive at the total wavelength estimate. The underestimate of the ground state splitting can be explained by carefully examining the energy level diagram (1). The resolution of the full spectrum is only good enough to fit to each of the four clusters, not to the individual transitions themselves. This means that the two peaks measured for 87 Rb represent the F = 1 F = 1 and F = 2 F = 2 transitions. However, the splitting that must be measured is between the F = 1 F = 2 peak and the F = 2 F = 2 peak. To correct for this, the splitting between the F = 1 and F = 2 levels can be added to the measured frequency, producing results that better match the known values, but are still slightly low. Fitting the peaks was also challenging for the excited state splittings. Stray reflections in the rubidium cell could cause Lamb dips to appear in the control beam, causing dips within dips in the subtracted spectrum. As a result, it was sometimes difficult to determine which section of a peak was the real peak of the Lamb dip. The errors are significantly larger in the 87 Rb spectrum compared to 85 Rb; this is because 85 Rb is roughly 2.5 times more prevalent and therefore gives a stronger signal. VI. CONCLUSION We find values for the coupling constants that agree for the most part with accepted values; where they do not, we theorize that the measured values actually measure a slightly different attribute of the spectrum, which can be corrected for to give values that agree more closely with the accepted figures. [1] MIT Balanced Detector. [2] Tui-Optics Laser System Brochure: DL100. [3] MIT Department of Physics. Doppler-free spectroscopy. November 2012. [4] G.T. Purves. Rubidium d2 line data, 2005. [5] Daniel A Steck. Rubidium 87 d line data, 2010. [6] Daniel A Steck. Rubidium 85 d line data, 2012.