The Chip-Scale Atomic Clock Recent Development Progress

Similar documents
THE CHIP-SCALE ATOMIC CLOCK COHERENT POPULATION TRAPPING VS. CONVENTIONAL INTERROGATION

MICROFABRICATED ATOMIC CLOCKS AT NIST

THE CHIP-SCALE ATOMIC CLOCK COHERENT POPULATION TRAPPING VS. CONVENTIONAL INTERROGATION

USING LASER DIODE INSTABILITIES FOR CHIP- SCALE STABLE FREQUENCY REFERENCES

A CPT-BASED 87 Rb ATOMIC CLOCK EMPLOYING A SMALL SPHERICAL GLASS VAPOR CELL

Quantum frequency standard Priority: Filing: Grant: Publication: Description

THE CHIP-SCALE ATOMIC CLOCK LOW-POWER PHYSICS PACKAGE

Continuous-Wave Characteristics of MEMS Atomic Clock VCSELs

Introduction Fundamentals of laser Types of lasers Semiconductor lasers

Doppler-Free Spetroscopy of Rubidium

Laser Locking with Doppler-free Saturated Absorption Spectroscopy

A New Compact High-Stability Oscillator

BMC s heritage deformable mirror technology that uses hysteresis free electrostatic

Optical phase-locked loop for coherent transmission over 500 km using heterodyne detection with fiber lasers

DIODE LASER SPECTROSCOPY (160309)

A PORTABLE RUBIDIUM FOUNTAIN 1

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

Chapter 1 Introduction

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

Lecture 6 Fiber Optical Communication Lecture 6, Slide 1

Oscillator for Chip-Scale Atomic

64 Channel Flip-Chip Mounted Selectively Oxidized GaAs VCSEL Array

Frequency Noise Reduction of Integrated Laser Source with On-Chip Optical Feedback

레이저의주파수안정화방법및그응용 박상언 ( 한국표준과학연구원, 길이시간센터 )

EXPERIMENTAL STUDY OF THE LASER DIODE PUMPED RUBIDIUM MASER

MICRO ION FREQUENCY STANDARD

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

VERTICAL CAVITY SURFACE EMITTING LASER

Vertical External Cavity Surface Emitting Laser

Zeeman Shifted Modulation Transfer Spectroscopy in Atomic Cesium

EE C245 ME C218 Introduction to MEMS Design Fall 2010

IST IP NOBEL "Next generation Optical network for Broadband European Leadership"

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

Spatial Investigation of Transverse Mode Turn-On Dynamics in VCSELs

A COMMERCIAL CPT RUBIDIUM CLOCK

Optoelectronics ELEC-E3210

VCSELS FOR RUBIDIUM D1 (795 NM)

THE SA.45S CHIP-SCALE ATOMIC CLOCK EARLY PRODUCTION STATISTICS

Integrated High Speed VCSELs for Bi-Directional Optical Interconnects

VCSELs With Enhanced Single-Mode Power and Stabilized Polarization for Oxygen Sensing

Optical cesium beam clock for eprtc telecom applications

Press Release A New Architecture for Miniaturization of Atomic Clocks Exploiting a Piezoelectric-Thin-Film Vibration

Low Thermal Resistance Flip-Chip Bonding of 850nm 2-D VCSEL Arrays Capable of 10 Gbit/s/ch Operation

Characteristics of absorption and dispersion for rubidium D 2 lines with the modulation transfer spectrum

THE CHIP-SCALE ATOMIC CLOCK PROTOTYPE EVALUATION

LONG-TERM STABILITY OF NIST CHIP-SCALE ATOMIC CLOCK PHYSICS PACKAGES

High-efficiency, high-speed VCSELs with deep oxidation layers

Supplementary Figures

Bistability in Bipolar Cascade VCSELs

Cavity QED with quantum dots in semiconductor microcavities

Narrow line diode laser stacks for DPAL pumping

Detection Beyond 100µm Photon detectors no longer work ("shallow", i.e. low excitation energy, impurities only go out to equivalent of

VCSELs for Atomic Clock Demonstrators

Wavelength Control and Locking with Sub-MHz Precision

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

High-power semiconductor lasers for applications requiring GHz linewidth source

Micro-nanosystems for electrical metrology and precision instrumentation

Instruction manual and data sheet ipca h

Optical phase-coherent link between an optical atomic clock. and 1550 nm mode-locked lasers

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

Leading Edge Technology Enables a Chip Scale Atomic Clock

21.0 Quantum Optics and Photonics

Chapter 1. Overview. 1.1 Introduction

White Paper. Miniature Atomic Clock (MAC) Pre-Production Results

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Electrical Engineering and Computer Science

Integrated Optoelectronic Chips for Bidirectional Optical Interconnection at Gbit/s Data Rates

A Narrow-Band Tunable Diode Laser System with Grating Feedback

InP-based Waveguide Photodetector with Integrated Photon Multiplication

Quantum-Well Semiconductor Saturable Absorber Mirror

Photonic Magnetometry at a (Short) Distance

Vixar High Power Array Technology

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

A SIMPLIFIED LASER AND OPTICS SYSTEM FOR LASER-COOLED RB FOUNTAIN FREQUENCY STANDARDS *

SUPPLEMENTARY INFORMATION

MMA RECEIVERS: HFET AMPLIFIERS

Novel Integrable Semiconductor Laser Diodes

Vertical Cavity Surface Emitting Laser (VCSEL) Technology

Improved Output Performance of High-Power VCSELs

DBR based passively mode-locked 1.5m semiconductor laser with 9 nm tuning range Moskalenko, V.; Williams, K.A.; Bente, E.A.J.M.

Holography Transmitter Design Bill Shillue 2000-Oct-03

A SPACE RUBIDIUM PULSED OPTICAL PUMPED CLOCK CURRENT STATUS, RESULTS, AND FUTURE ACTIVITIES

First results of a high performance optically-pumped cesium beam clock

3 General Principles of Operation of the S7500 Laser

EE119 Introduction to Optical Engineering Fall 2009 Final Exam. Name:

A HIGH PRECISION QUARTZ OSCILLATOR WITH PERFORMANCE COMPARABLE TO RUBIDIUM OSCILLATORS IN MANY RESPECTS

Mode analysis of Oxide-Confined VCSELs using near-far field approaches

Understanding the Magnetic Resonance Spectrum of Nitrogen Vacancy Centers in an Ensemble of Randomly-Oriented Nanodiamonds, Supporting Information

Small, Low Power, High Performance Magnetometers

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

Spectrometer using a tunable diode laser

RF-INTERROGATED END-STATE CHIP-SCALE ATOMIC CLOCK

SUPPLEMENTARY INFORMATION DOI: /NPHOTON

STUDIES ON AN IMPROVED COMPACT PHYSICS PACKAGE FOR RUBIDIUM STANDARDS

Ph 77 ADVANCED PHYSICS LABORATORY ATOMIC AND OPTICAL PHYSICS

CHAPTER 2 MICROSTRIP REFLECTARRAY ANTENNA AND PERFORMANCE EVALUATION

MEMS BASED QUARTZ OSCILLATORS and FILTERS for on-chip INTEGRATION

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

A new picosecond Laser pulse generation method.

A novel tunable diode laser using volume holographic gratings

ABSTRACT. This paper describes the performance characteristics of a new, rugged 5 MHz quartz crystal oscillator

Transcription:

The Chip-Scale Atomic Clock Recent Development Progress R. Lutwak*, D. Emmons, T. English, and W. Riley Symmetricom - Technology Realization Center A. Duwel and M. Varghese Charles Stark Draper Laboratory D. K. Serkland and G. M. Peake Sandia National Laboratories Abstract We have undertaken the development of a chip-scale atomic clock (CSAC) whose design goals include short-term stability, σ y (τ = 1 hour), of 1x10-11 with a total power consumption of 30 mw and an overall device volume of 1 cm 3. The stringent power requirement dominates the physics package architecture, dictating a small (< 5 mm 3 ) gaseous atomic ensemble interrogated by a low-power semiconductor laser. At PTTI 2002, we reported on initial experimental investigations leading to the decision to employ the coherent population trapping (CPT) interrogation technique. This paper describes our further progress on the CSAC effort, including the development of custom vertical cavity surface emitting laser (VCSEL) sources and techniques for microfabricating miniature cesium vapor cells comprised of anodically bonded silicon and glass. Measurements of the signal contrast and linewidth of both the cesium D1 and D2 resonance transitions are compared, and frequency stability measurements of the CSAC testbed are presented. INTRODUCTION We are developing a chip-scale atomic clock (CSAC) with the goal of producing a 1 cm 3 device that can provide a stability of 1x10-11 over 1 hour while consuming only 30 mw. While the stability goals are 10 modest, equivalent to ( τ = 1) = 6 10, the power and size goals are quite ambitious, each more than σ y two orders of magnitude smaller than existing technology. Our initial work, reported at this conference last year [1], focused on the general architecture of the physics package, a small cesium gas cell interrogated using the method of coherent population trapping (CPT). In this paper, we report on our * E-mail: RLutwak@Symmetricom.com 34 Tozer Rd., Beverly, MA 01915 Sandia is a multiprogram laboratory operated by Sandia Corporation, a Lockheed Martin Company, for the United States Department of Energy s National Nuclear Security Administration under contract DE-AC04-94AL85000. -1-

continued progress, including comparative spectroscopy of the cesium D1 and D2 CPT resonances and experimental investigations of small silicon-glass resonance cells. SHORT-TERM STABILITY FIGURE-OF-MERIT Generally speaking, the short-term stability of any atomic frequency standard is proportional to the resonance linewidth and inversely proportional to the signal-to-noise ratio of the detected signal, i.e. γ σ y ( τ = 1sec), S / N ( ) 1Hz where γ is the resonance linewidth (FWHM) and (S/N) 1 Hz is the Signal-to-Noise ratio, measured in a 1 Hz bandwidth at the modulation frequency, F MOD, typically on the order of γ/2. In the following sections, we discuss the phenomena which determine these quantities in the CSAC and the trade-offs which must be made in order to meet the CSAC stability requirements within its power budget. Our goal is to establish a figure-of-merit, ξ, which predicts frequency stability based on spectroscopic measurements of linewidth and signal contrast. There are many contributions to the overall noise figure of the CSAC including statistical shot noise of the detected signal, phase noise of the local oscillator, pre-amplifier current and voltage noise, and laser frequency noise. The performance of traditional lamp-pumped vapor cell atomic frequency standards is predominantly limited by statistical shot noise of the detected photocurrent, which is proportional to the square-root of the DC level of detected signal, i.e. N (DC Level) 1/2. In the laser-pumped CSAC, we have the additional noise contribution due to laser frequency (or phase) noise, which is converted to amplitude noise by non-linear resonant scattering from the atomic sample. The mechanism of FM-AM conversion due to resonant scattering is illustrated schematically in Figure 1. In normal operation, the laser frequency is locked to the peak of the atomic resonance line, where the signal is maximized and the sensitivity to frequency excursions is minimized. However, for small excursions of the laser frequency, which arise due to noise on its bias current, temperature fluctuations, and statistical properties of the VCSEL itself, the amplitude fluctuations on the detected signal increase quadratically with detuning from the peak. Moreover, because the atomic response is non-linear, the vapor cell effectively mixes noise components at all frequencies, limited only by the intrinsic 5 MHz bandwidth of the atomic response, into the detection band at F MOD. The relative magnitude of this noise process is determined by the inhomogeneous linewidth of the atomic resonance and the frequency noise properties (linewidth) of the VCSEL, and is linearly proportional to the light intensity, i.e. N (DC Level). Amplitude f A Frequency Figure 1: Laser FM AM noise conversion -2-

The two principal noise mechanisms, shot noise and laser frequency noise can thus be differentiated by their power-law dependence on the laser intensity. In order to determine their relative impact in the CSAC device, we ve made measurements of Noise as a function of DC Level in our testbed apparatus. Figure 2 shows the RMS noise, measured in a 1 Hz bandwidth at 500 Hz, as a function of laser power, over two orders of magnitude spanning the designed operating region of the CSAC. Also shown is a fit to a power law model of (S/N) which indicates N (DC Level) 0.998. We conclude that, for typical CSAC components and operating conditions, the principal noise mechanism in the CSAC is laser frequency noise and, for practical design purposes, N (DC Level). For the purposes of definition, Figure 3 shows schematically the elements of the CPT resonance signal. The CPT spectrum is dominated by the Gaussian laser absorption feature, whose width may be several GHz due to a combination of Doppler and pressure gas broadening. The Lorentzian CPT signal sits atop the Doppler absorption feature and is characterized by its amplitude, designated Signal, and linewidth, γ. We define the contrast, η, of the CPT resonance as Signal divided by DC Level. Noise [na RMS /Hz 1/2 ] 10 1 0.1 0.1 1 10 DC Level [µa] Light Transmission Figure 2: Noise vs. DC Level Signal Linewidth, γ DC Level Laser Frequency Figure 3: CPT resonance features With these definitions, and the aforementioned linear relationship between DC Level and Noise, we arrive at the appropriate figure-of-merit, ξ, for predicting short-term stability of the CSAC: ξ = η / γ, where η and γ are measured in ppm and Hz respectively. This leads to the relationship 1 1/ 2 σ y ( τ ) τ, ξ where the constant of proportionality is determined by the spectrum of frequency noise on the VCSEL. This presents a valuable relationship, from the perspective of CSAC engineering, in that it permits the evaluation of different atomic transitions and species and operating conditions without explicit dependence on the noise properties of the particular laser source employed in the measurement. -3-

LABORATORY TESTBED We have improved our laboratory CSAC testbed over the past year, in order to better support precision CPT resonance spectroscopy and evaluation of prototype components. Figure 4, below shows a layout of the experimental setup as configured for spectroscopic measurements. The experimental testbed includes independent temperature-stabilized assemblies for the VCSEL and resonance cell as well as a separate reference cell, containing cesium and buffer gas, which allows consistent and reproducible laser stabilization independent of the laser intensity, vapor density, and geometric characteristics of the resonance cell. A variable attenuator at the input to the reference cell allows for two orders of magnitude of intensity variation. The RF chain provides microhertz resolution tunability at 4.6 GHz, with negligible offset and phase noise, continuous power adjustment to +20 dbm, and modulation and sweep capabilities. The VCSEL is locked to the reference cell by applying audio frequency modulation to its drive current, synchronously detecting the response from the reference cell and applying a correction to the bias current from the proportional/integral loop filter. The RF frequency is repetitively swept across the CPT resonance while the cell transmission is collected and averaged on a digital storage oscilloscope. The data is transferred to a personal computer for subsequent analysis and archive. Figure 5 shows a typical CPT signal, measured on the D1 resonance of cesium at λ=894 nm in a small glass cell with 1 mm path length, similar to that shown in Figure 5 of reference [1], containing cesium and 50 Torr of buffer gas at a temperature of 75 C. Also shown in the figure is a fit to a Lorentzian lineshape. The fit demonstrates a linewidth of γ=879 Hz and peak contrast η=7828 ppm, with reducedχ 2 =0.996. Note that this measurement leads to a figure-of-merit of ξ=8.9. VCSEL Laser Current Atten Laser Diode Oven Temp Ctrl Laser Lock Loop & Modulation Bias T RF Chain OCXO 4.6 GHz Variable Beam Splitter Loop Filter DDS PD MEMS Gas Cell Atten λ/4 Reference Cell Lock-In Amplifier Sweep Gen Cell Oven Temp Ctrl PD Preamp Y Digital Oscilloscope X CPT Signal Data To PC For Analysis Figure 4: Testbed configured for CPT spectroscopy Contrast [ppm] 10000 7500 5000 2500 0-5 -4-3 -2-1 0 1 2 3 4 5 Detuning from 4.596 GHz [khz] Data Lorentzian γ = 880 Hz Figure 5: CPT resonance signal -4-

D1 VS. D2 EXCITATION There are two principal optical transitions, accessible with diode laser technology, which could be employed for CPT interrogation of the cesium ground state hyperfine resonance. These two, termed D2 and D1, are shown in Figure 6. Research on a similar alkali system, rubidium, has indicated that the D1 CPT transition exhibits a narrower linewidth and higher contrast than D2, both of which contribute to higher ξ [2]. Because the cesium structure is substantially similar to that of rubidium, a similar improvement could be expected in cesium CPT spectroscopy, though to some lesser extent, due to the higher degeneracy and details of the cesium hyperfine structure. Prior to this work, all published data of CPT spectroscopy in cesium was performed on the D2 line, due to the availability of 852 nm laser devices. In the past year, we have developed VCSEL devices at 894 nm, permitting a direct comparison of the performance advantages to a CSAC based on either transition. Figure 6 shows the transitions of the cesium atomic structure relevant for CPT interrogation, including both the D1 and D2 principal optical transitions at λ=894 nm and λ=852 nm, respectively, each of which contains a cluster of hyperfine states. The D2 optically excited 6 P 3/2 state is composed of four angular momentum substates, F=2-5, while the D1 optically excited 6 P 1/2 state is composed of only two sub-states, F=3,4. The optical dipole selection rules, applicable to CPT interrogation, limit the participating excited state transitions to F=3,4. The additional F=2 and F=5 excited state components in the D2 manifold scatter light, reducing the available atomic population without contributing to the CPT signal, thus reducing the contrast. In addition, the F=2 and F=5 excited states in D2 provide a pathway for CPT decoherence, resulting in a broader resonance line. Note that the spacing of the F=3 and F=4 excitedstate transitions is 6X wider in the 6 P 1/2 (D1) Figure 6: Cesium D1 and D2 CPT resonance levels manifold than in 6 P 3/2 (D2). In a Doppler-broadened gas cell, the inhomogeneous atomic linewidth is roughly 400 MHz and the D1 lines, which are separated by 1.2 GHz, are fully resolved while the D2 lines, separated by only 200 MHz, are not. As a result, in a Doppler-broadened cell, the D2 CPT resonance operates simultaneously on both excited-state transitions, while the D1 does not, thus reducing the effective transition strength of D1 excitation by a factor of 2. This detrimental effect is somewhat reduced in the CSAC, wherein the additional inhomogeneous line broadening due to the necessarily high buffer gas pressure limits the resolution of the D1 transitions. The relative advantage of D1 CPT interrogation is less pronounced in cesium than in rubidium for several reasons. First, note that the Zeeman degeneracy of both the ground and excited states limits the fraction of optical signal which contributes constructively to the CPT resonance signal. Because the cesium system has higher degeneracy than rubidium not only is the signal contrast lower for both D1 and D2 excitation, but the relative advantage of D1 is also proportionately reduced. Secondly, in rubidium the splitting of -5-

the P 1/2 sub-levels is only 800 MHz and the Doppler width is 500 MHz, such that both components contribute to the D1 CPT resonance signal, even at relatively low buffer gas pressure. The D1 VCSELs, developed at Sandia National Laboratories in support of this program, have been optimized in several respects to support the CSAC architecture. The 1-λ cavity length ensures that the VCSELs operate in a single longitudinal mode and thus are free of the mode-hopping behavior which characterizes conventional cleaved facet edge-emitting laser devices. The CSAC requirements impose additional demands on the optical performance of the laser device, including that it must operate on a single transverse cavity mode, that the polarization remain stable, and that the wavelength tune to the cesium atomic resonance over life and at the operating temperature of the physics package. In addition to these optical requirements, the devices have been optimized for high efficiency to meet the stringent power requirements of the CSAC application, delivering 100 µw of optical power while dissipating only 2 mw, and for high modulation efficiency, producing optimal sideband amplitude with less than 1 mw of RF drive power at 4.6 GHz. Figure 7, below shows a representative photograph and light-current-voltage data for the devices developed for the CSAC program. 10 8 j278n.axg 5 4 Power (mw) 6 4 2 3 2 1 Voltage (V) 0 0 0 2 4 6 8 10 12 14 16 18 Current (ma) (a) (b) Figure 7: Prototype VCSEL device (a) photograph and (b) light-current-voltage data The D1 VCSELs were fabricated using the same semiconductor processing steps as for 850-nm oxideaperture VCSELs. The oxide aperture simultaneously confines both the current and the optical mode, thereby resulting in lower threshold currents than VCSELs that use proton implantation to create a current aperture. In order to reach a wavelength of 894 nm, the epitaxial semiconductor structure was modified slightly from the standard 850-nm design. The VCSEL was grown on a semi-insulating GaAs substrate to minimize the parasitic capacitance of the wire-bond pads. The bottom distributed Bragg reflector (DBR) is designed as a high reflector, containing 36 pairs of n-doped quarter-wave Al 0.16 Ga 0.84 As and Al 0.92 Ga 0.08 As layers. The active region contains 5 In 0.07 Ga 0.93 As quantum wells, each 6-nm thick, producing a room-temperature optical gain peak near 885 nm. A quarter-wave layer of Al 0.98 Ga 0.02 As immediately above the active region is selectively oxidized to form a circular oxide aperture. The oxide aperture diameter is kept below 4 microns to force the VCSEL to operate only in the fundamental transverse mode. The top DBR is designed as an output coupling mirror, containing 21 pairs of p-doped quarter-wave Al 0.16 Ga 0.84 As and Al 0.92 Ga 0.08 As layers (in the pair immediately above the active region, the Al 0.92 Ga 0.08 As layer is replaced with Al 0.98 Ga 0.02 As). The doping levels in the DBRs are lower than in typical commercial devices in order to obtain higher conversion efficiency. -6-

For the measurements reported below, the Sandia D1 VCSEL is compared with a D2 VCSEL manufactured by MODE corporation [3]. The two devices are similar, though the Sandia device is roughly 10X more efficient, both at DC and RF, and exhibits a 5X higher noise floor, due perhaps to increased sensitivity to noise on the DC bias current associated with the higher device bandwidth. The noise performance, which is an area of ongoing improvement, has no impact on the measurements contained herein due to our definition of the figure-of-merit, ξ, which is independent of the actual noise figure of the laser source. 1.0 1.0 0.8 0.8 Transmission [Arb] 0.6 0.4 Transmission [Arb] 0.6 0.4 0.2 D1 D2 0.2 D1 D2 0.0-20 -10 0 10 20 Laser Tuning [GHz] 0.0-20 -10 0 10 20 Laser Tuning [GHz] (a) (b) Figure 8: D1 vs. D2 optical absorption spectra with (a) RF off and (b) RF on Figure 8, (a), and (b), show the laser absorption spectra as the laser frequency is swept across the D1 and D2 transitions with the 4.6 GHz RF drive off and on, respectively. In all cases the resonance cell was the small glass cell from reference [1], filled with cesium and 70 Torr of a temperature-compensated nitrogen/argon buffer gas mixture at our design operating temperature of 65 C. Note that the D2 absorption is roughly 4X that of D1, due partly to the aforementioned resolution of the D1 excited states and partly due to additional scattering from the F=2 and F=5 substates of D2. 10000 D1 D2 8000 7000 D1 D2 Linewidth, λ [Hz] 1000 Contrast, η [ppm] 6000 5000 4000 3000 2000 1000 1 10 100 Laser Power [µw] 0 1 10 100 Laser Power [µw] (a) (b) Figure 9: D1 vs. D2 CPT resonance (a) linewidth and (b) contrast vs. laser power -7-

Figures 9(a) and (b) show the measured linewidth and contrast for the D1 and D2 transitions. Note that in both cases the linewidth appears to exhibit a simple power law relationship to the laser power, due to optical power broadening of the optical resonance. At low laser power, the linewidth is limited to γ 500 Hz due to cesium-cesium (spin exchange) collisions in the vapor cell. If necessary, the linewidth can be further reduced by lowering the cell temperature, at the expense of both signal and contrast, and thus ξ. In both the D1 and D2 cases, the contrast improves with increasing laser power until the point at which the atomic sample is fully utilized, at which point the DC level continues to increase without increase in signal. The results of Figure 9 are combined to produce our figure-of-merit, shown below in Figure 10. Our results indicate a 3-4X advantage for CSAC of employing the D1 transition instead of the D2 transition. In both cases the optimum laser power is less than 10 µw, though the slightly higher optimum power level of D1 is an additional advantage in that it provides a higher signal level into the signal recovery electronics. Figure-of-Merit, ξ 9 8 7 6 5 4 3 2 1 0 D1 D2 1 10 100 Laser Power [µw] MEMS CELL FABRICATION Figure 10: D1 vs. D2 figure-of-merit vs. laser power In addition to the power and stability requirements, it is also important that the CSAC design be costeffective for high-yield large-volume production. Towards this goal we have begun development at Draper Laboratory of a silicon-glass cesium resonance cell, utilizing the manufacturing techniques of micro-electromechanical systems (MEMS). The proposed manufacturing scheme is illustrated below. (a) 4" Figure 11: MEMS cell fabrication concept (b) -8-

Figure 11 illustrates the planned methodology for large-scale production of cesium resonance cells for CSAC. On the left, (a), an array of approximately 500 1.5 mm holes is etched into a 1.5 mm thick 4 diameter silicon wafer by the technique of deep reactive ion etching (DRIE). A 4 Pyrex wafer is bonded to the bottom of the wafer using the vacuum tight technique of anodic bonding. The cesium and buffer gas are introduced into the cells and then a second Pyrex wafer is bonded to provide the upper window and complete the vacuum seal. The cells are then diced from the wafer for subsequent incorporation into the CSAC. Figure 11 (b) shows isometric and cross-sectional views of a single diced cell. While these techniques for batch fabrication are under development, we have begun experimentation in the CSAC testbed system with prototype cells of isomorphic geometry and material composition. Prototype cells were manufactured via a hybrid process of MEMS cell fabrication combined with conventional glassblowing and cell processing techniques. (a) (b) Figure 12: MEMS prototype cell components (a) micromachined silicon wafer and (b) tubulated cell Photolithography followed by DRIE was employed to create the dumbbell pattern in the silicon substrate as shown in Figure 12 (a). On the right hand side, a 1.5 mm diameter hole is etched all the way through the 1.5 mm thick silicon wafer. This is connected, via a 1 mm wide by 100 µm deep pathway, to a second 1.5 mm aperture, also only 100 µm deep. The top and bottom Pyrex wafers are attached via anodic bonding as described above, only the top plate has been ultrasonically drilled with an array of 0.5 mm holes which align with the shallow fill area of the dumbbells. After the anodic bonding process is complete, the cells are diced from the wafer into 1 cm squares, each containing a single 1.5 mm diameter x 1.5 mm thick resonance cell, with a small additional volume accommodating the tubulation and attachment. Tubulations are individually attached to each cell using conventional glassblowing techniques, as shown in Figure 12(b), and the cells are filled with cesium and buffer gas using a conventional glass manifold processing system. The resultant cell is shown below in Figure 13. -9-

1.5 mm Figure 13: Prototype silicon-glass MEMS cell Figure 14 reflects efforts in our laboratory to measure narrow linewidths in the MEMS cell. The cell employed for this measurement has an elliptical cross-section, 2 mm x 3 mm, and 800 µm path length. The cell contains cesium and 90 Torr of temperaturecompensated N 2 /Ar buffer gas mixture. The cell is stabilized at our designed operating temperature of 65 C and illuminated with only 1.2 µw of laser power. The fit to a Lorentzian lineshape has a reduced χ 2 =0.98 and indicates a linewidth of γ=438 Hz and peak contrast η=110, leading to ξ=0.25. Contrast [ppm] 120 100 80 60 γ=438 Hz 40 20 0-20 -4-3 -2-1 0 1 2 3 4 Detuning from 4596.3553 MHz [khz] Figure 14: Relatively narrow CPT resonance in MEMS cell As discussed earlier, it is possible to observe even narrower linewidths at the expense of ξ, and thus clock stability. By reducing the cell temperature to 50 C and the laser power to 500 nw, we have observed linewidths as narrow as 265 Hz in similar micromachined silicon-glass cells, though at values of ξ<0.03 with little practical value for CSAC application. -10-

Our laboratory CSAC testbed can be configured to operate as a closed-loop frequency standard as shown in Figure 15. Our experimental setup for stability measurements is similar to that used for spectroscopy (compare Figure 4) with the following alterations. Rather than sweeping the 4.6 GHz RF frequency and observing the cell transmission on an oscilloscope, the RF is sinusoidally modulated at a rate 500 Hz and depth of 1 khz. The detected signal is coherently demodulated by a lock-in amplifier and the resultant error signal is integrated by the loop filter and fed back to the 5 MHz quartz oscillator which serves both as the reference signal to the RF system and provides a buffered output for direct comparison with our house standard hydrogen maser via standard frequency measurement techniques. VCSEL Laser Current To Freq Meas System Atten Laser Diode Oven Temp Ctrl Bias T RF Chain OCXO 4.6 GHz 5 MHz CV Atten Loop Filter Loop Filter MEMS Gas Cell λ/4 Cell Oven Temp Ctrl Lock-In Amplifier Lock-In Amplifier PD Laser Lock Loop & Modulation Freq Lock Loop & Modulation Preamp CPT Signal Figure 15: Testbed configuration for closed-loop operation Figure 16: Frequency stability of CSAC testbed with MEMS cell Figure 16 shows the Allan deviation for the CSAC testbed, utilizing the MEMS prototype cell, D1 laser excitation and typical operating parameters for temperature and laser intensity. For this 12-hour measurement, the cell employed had a 1.5 mm diameter aperture, 1.5 mm path length, and was stabilized at T=70 C. The laser power was 20 µw and the figure-of-merit of the resonance signal was ξ=3.2. At short times, the frequency stability displays the τ -1/2 behavior which typifies atomic frequency standards and displays a 1-second intercept of σ y ( τ = 1) = 6 10 10 σ y ( τ = 1) = 1.5 10, roughly 4X margin over our design goal of. We expect that further optimization of cell geometry, buffer gas pressure, and laser intensity will lead to further improvement in ξ and thus σ. At longer times (>1000 seconds), the device exhibits flicker behavior, reflecting environmental sensitivity, most likely due to thermal variation in the required bias voltage of the VCSEL. We expect this medium-term environmental sensitivity to be 10-11-

greatly reduced when the thermal isolation of the physics package is implemented in the next phase of the CSAC development effort. CONCLUSIONS We have performed an experimental comparison of the D1 and D2 cesium CPT resonances in a CSAC experimental system and have found the D1 resonance to provide a 3-4X better figure-of-merit for CSAC stability. We have developed prototype silicon-glass MEMS cells and verified short-term stability utilizing these cells as well as the low-power VCSEL devices designed for implementation in the CSAC. At this point, we have tested all of the key physics package components necessary to implement a small low-power CSAC and have found them capable of supporting the performance and power goals of the project. ACKNOWLEDGEMENTS The authors wish to thank Mike Garvey, of the Symmetricom TRC, and Jim Sitomer of Draper Laboratory, for supporting and directing this effort. We are also grateful for the exceptional glassblowing skills of Patrick DeFlorio of the Yankee Glassblowing Company and of Luke Tomlin and Joe Reid of the Symmetricom TRC. E. L. Blansett of Sandia National Laboratories contributed to investigations of polarization stability in our VCSEL development. This work is supported by the Defense Advanced Research Products Agency, Contract # NBCHC020050. REFERENCES [1] R. Lutwak, et. al., The Chip-Scale Atomic Clock Coherent Population Trapping vs. Conventional Interrogation, Proceedings of the 34 th Annual Precise Time and Time Interval (PTTI) Systems and Applications Meeting, December 3-5, 2002, Reston, VA, pp. 539-550. [2] M. Stahler, et. al., Coherent population trapping resonances in thermal 85 Rb vapor: D 1 vs D 2 line excitation, Optics Letters, vol. 27, August 15, 2002, pp. 1472-1474. [3] MODE Corporation, Product Number 8085-1000. -12-