physics0609133/Text.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: 
3: %% Trim Size: 9.75in x 6.5in
4: 
5: %% Text Area: 8in (include Runningheads) x 5in
6: 
7: %% ws-mplb.tex   :   10 September 2002
8: 
9: %% TeX file to use with ws-mplb.cls written in Latex2E.
10: 
11: %% The content, structure, format and layout of this style file is the
12: 
13: %% property of World Scientific Publishing Co. Pte. Ltd.
14: 
15: %% Copyright 1995, 2002 by World Scientific Publishing Co.
16: 
17: %% All rights are reserved.
18: 
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: 
21: 
22: 
23: \documentclass{ws-mplb}
24: 
25: 
26: 
27: \begin{document}
28: 
29: 
30: 
31: \markboth{F. Sorrentino, G. Ferrari, N. Poli, R. Drullinger and G.
32: M. Tino}{Laser Cooling of Atomic Strontium}
33: 
34: 
35: 
36: %%%%%%%%%%%%%%%%%%%%% Publisher's Area please ignore %%%%%%%%%%%%%%%
37: 
38: %
39: 
40: \catchline{}{}{}{}{}
41: 
42: %
43: 
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: 
46: 
47: 
48: \title{LASER COOLING AND TRAPPING OF ATOMIC STRONTIUM FOR ULTRACOLD ATOMS
49: PHYSICS, HIGH-PRECISION SPECTROSCOPY AND QUANTUM SENSORS}
50: 
51: 
52: 
53: \author{\footnotesize F. Sorrentino, G. Ferrari, N. Poli, R. Drullinger\footnote{on leave from NIST, 325 Broadway, Boulder, Colorado 80305} and G. M. Tino}
54: 
55: 
56: 
57: \address{Dipartimento di Fisica and LENS - Universit\`a di Firenze, INFN,
58: INFM,\\ I-50019 Sesto Fiorentino (FI), Italy\\
59: guglielmo.tino@fi.infn.it}
60: 
61: 
62: 
63: \maketitle
64: 
65: 
66: 
67: \begin{history}
68: 
69: \received{(received date)}
70: 
71: \revised{(revised date)}
72: 
73: %\accepted{(Day Month Year)}
74: 
75: %\comby{(xxxxxxxxxx)}
76: 
77: \end{history}
78: 
79: 
80: 
81: \begin{abstract}
82: 
83: This review describes the production of atomic strontium samples
84: at ultra-low temperature and at high phase-space density, and
85: their possible use for physical studies and applications. We
86: describe the process of loading a magneto-optical trap from an
87: atomic beam and preparing the sample for high precision
88: measurements. Particular emphasis is given to the applications of
89: ultracold Sr samples, spanning from optical frequency metrology to
90: force sensing at micrometer scale.
91: \end{abstract}
92: 
93: 
94: 
95: 
96: 
97: 
98: 
99: \section{Introduction}
100: 
101: \label{Introduction}
102: 
103: 
104: 
105: Recently, laser-cooled atomic strontium has been the subject of
106: active research in several fields spanning from all-optical
107: cooling towards quantum degeneracy for
108: bosonic,\cite{KatoriMOT1999,Ferrari2006} and
109: fermionic\cite{Mukaiyama2003} isotopes, cooling
110: physics,\cite{Xu2003,Loftus2004} continuous atom
111: laser,\cite{Katori2001} detection of ultra-narrow
112: transitions,\cite{Takamoto2003,Courtillotclock,Ferrari2003,Ido2005}
113: multiple scattering,\cite{Bidel2002} and collisional
114: theory\cite{Derevianko2003}.
115: 
116: 
117: 
118: Much of that interest relies on the features of the electronic
119: level structure of alkali-earth atoms, that make them ideal
120: systems for laser manipulation and for the realization of quantum
121: devices. Among the alkali-earth metals, strontium summarizes most
122: of the useful properties both for the preparation of ultracold
123: samples and for applications.
124: 
125: 
126: 
127: The ground-state Sr atom presents a strong and quasi-closed
128: optical transition well suited for efficient trapping in
129: magneto-optical traps (MOT's) from thermal
130: samples,\cite{Raab1987,Kurosu1990} and a multiplet of weak
131: intercombination transitions with large interest in laser cooling
132: and optical metrology. All of these transitions are easily
133: accessible with compact solid-state laser sources, as described in
134: \ref{ExperimentalSetup}. Second-stage cooling on the narrow
135: $^1$S$_0$-$^3$P$_1$ intercombination line was proven to be an
136: efficient method to reduce the sample temperature down to the
137: photon recoil limit.\cite{KatoriMOT1999}
138: 
139: 
140: 
141: The absence of a nuclear spin in all of the bosonic isotopes
142: greatly simplifies the electronic energy spectrum with respect to
143: the already simple structure of alkali atoms, and allows a direct
144: verification of theories in the field of light
145: scattering\cite{Chanelière2004} and laser
146: cooling.\cite{Loftus2004}  Moreover, the lack of nuclear spin
147: results in minimum sensitivity to stray magnetic fields: this has
148: important consequences on high-precision measurements, as
149: discussed in sections  \ref{Microsensor} and \ref{1S3PFreqMeasure}.
150: 
151: 
152: 
153: Interatomic collisions may represent an important source of
154: perturbations in atomic quantum devices. In this respect, the
155: $^{88}$Sr atom exhibits excellent features due to its remarkably
156: small collisional cross-section, resulting in the longest
157: coherence time for Bloch oscillations
158: observed so far.\cite{FerrariBloch}
159: 
160: 
161: 
162: As a consequence of their zero magnetic moment, ground-state
163: bosonic Sr isotopes cannot be magnetically trapped. However,
164: optical dipole trapping with far-off resonant laser fields
165: provides an effective method for trapping in conservative
166: potentials. By properly choosing the wavelength of the trapping
167: laser field, it is possible to design optical potentials that
168: shift equally levels belonging to the singlet and triplet
169: manifolds, hence avoiding perturbations to a given
170: intercombination line.\cite{KatoriFORT1999} This feature is
171: the basis of laser cooling in dipole traps and allows one to reach
172: a temperature close to the recoil limit with a phase-space density
173: close to quantum degeneracy both with bosonic and fermionic
174: isotopes.\cite{Ido2000,Mukaiyama2003,Ferrari2006}
175: 
176: 
177: 
178: Several groups are presently working on laser cooled strontium.
179: The Tokyo group first demonstrated the double-stage optical
180: cooling down to sub-$\mu$K temperatures;\cite{KatoriMOT1999} they
181: also proposed and realized a Sr ``optical lattice
182: clock'',\cite{Takamoto2005} and made pioneering studies on
183: electric microtraps for strontium.\cite{Katori2006}  The JILA
184: group has made interesting studies on cooling
185: physics,\cite{Xu2003,Loftus2004} and on metrological
186: applications.\cite{Ido2005} Our group is concerned both with
187: ultracold physics studies,\cite{Ferrari2006} and with applications
188: to optical metrology\cite{Ferrari2003} and quantum
189: devices.\cite{FerrariBloch} The BNM-SYRTE group is mainly
190: concerned with metrological
191: applications,\cite{Courtillotclock,Lemonde2005,Targat2006} the Houston
192: group with ultracold Sr physics,\cite{Nagel2005,Mickelson2005} while the
193: Nice group has studied multiple scattering of light from cold Sr
194: samples,\cite{Chanelière2004} and analyzed several details of the
195: Sr cooling mechanisms.\cite{Chanelière2005}
196: 
197: 
198: 
199: 
200: 
201: From the experimental point of view, some of the techniques and
202: solutions involved in atomic strontium cooling are specific to
203: this atom and to some extent uncommon in the field of atomic
204: physics and laser cooling. In this paper we give a detailed
205: description of how to prepare an ultra-cold strontium sample which
206: is well suited for high-precision spectroscopy, the study of
207: quantum degenerate gases and quantum sensors. The presentation has
208: the following structure:
209: 
210: 
211: 
212: \begin{itemize}
213: \item section \ref{Strontiumatom} summarizes the properties of the
214: strontium atom, \item section \ref{MOT} describes the process of
215: laser cooling of strontium starting from the slowing of a thermal
216: beam, to the cooling down to the photon recoil limit and trapping
217: in a conservative potential, \item section
218: \ref{CollisionalMeasure} deals with the study of ultracold
219: collisions on the ground-state bosonic Sr isotopes, \item in
220: section \ref{SrBEC} we show some recent advances towards the
221: realization of a Bose-Einstein condensate (BEC) of strontium,
222: \item in section \ref{Microsensor} we present the use of ultracold
223: Sr atoms as force sensors at micrometric scale, \item in section
224: \ref{1S3PFreqMeasure} we review the recent frequency measurements
225: on the Sr intercombination transitions with application to optical
226: metrology, \item in \ref{ExperimentalSetup} we give a detailed
227: description of the experimental setup employed in laser cooling of
228: strontium, namely the vacuum apparatus and the laser sources.
229: \end{itemize}
230: 
231: 
232: 
233: 
234: 
235: 
236: 
237: \section{Properties of the strontium atom}
238: \label{Strontiumatom}
239: 
240: 
241: The main properties of atomic strontium are common to almost all
242: alkaline-earth metals. At ambient temperature it appears as a
243: metal. Its vapor pressure is quite low, and it reaches 1 mTorr at
244: 1000\,Kelvin. The Sr atom is rather reactive: it forms compounds
245: with oxygen, nitrogen, water and silicates, while it is inert
246: against sapphire. Thus in working with Sr vapors it is common to
247: employ sapphire windows.\cite{Neuman1995}
248: 
249: 
250: 
251: Strontium has four natural isotopes, whose properties are listed
252: in tab. 1. The bosonic (even) isotopes have zero nuclear spin,
253: thus they are perfect scalar particles in the $J=0$ states. This
254: has important consequences for applications to optical metrology
255: and quantum sensors (see sections \ref{Microsensor} and
256: \ref{1S3PFreqMeasure}).
257: 
258: 
259: 
260: 
261: 
262: \begin{table}[pt]
263: \tbl{Natural Sr isotopes (NIST data)} {\begin{tabular}{@{}cccc@{}}
264: \toprule Isotope & Relative atomic mass & Relative abundance &
265: Nuclear spin  \\ \colrule
266: $^{88}$Sr &   87.905 6143(24) & \hphantom{0}82.58(1)\% & 0 \\
267: $^{86}$Sr &  85.909 2624(24) & \hphantom{00}9.86(1)\% & 0 \\
268: $^{87}$Sr &  86.908 8793(24) & \hphantom{00}7.00(1)\% & 9/2 \\
269: $^{84}$Sr &  83.913 425(4) & \hphantom{000}0.56(1)\% & 0\\
270: \botrule
271: \end{tabular}}
272: \end{table}
273: 
274: 
275: 
276: 
277: 
278: Concerning the electronic level structure, due to the presence of
279: two electrons in the outer shell the atomic states can be grouped
280: into two separate classes: singlets and triplets. Since the
281: spin-orbit interaction breaks the spin symmetry, intercombination
282: transitions between singlet and triplet states are weakly
283: allowed.\cite{LibroFisicaAtomica} A simplified scheme of relevant
284: Sr levels and transitions is shown in figure \ref{LevelScheme}.
285: 
286: 
287: 
288: \begin{figure}[t]
289: \centerline{\psfig{file=fig1,width=8cm}} \vspace*{8pt}
290: \caption{Electronic level structure of atomic strontium. The
291: transitions relevant for laser cooling and optical manipulation
292: are indicated as well as their linewidths.}\label{LevelScheme}
293: \end{figure}
294: 
295: 
296: 
297: 
298: 
299: The strong $^1$S$_0$-$^1$P$_1$ transition at 461 nm has a natural
300: width of 32 MHz, and it is used for laser cooling and
301: trapping.\cite{Raab1987,Kurosu1990} Such transition is not
302: perfectly closed, due to a small leakage towards the 4d\,$^1$D$_2$
303: state (branching ratio $\sim 10^{-5}$). The direct
304: $^1$D$_2$-$^1$S$_0$ decay channel is forbidden in dipole
305: approximation, and atoms from the $^1$D$_2$ basically decay
306: towards the 5p\,$^3$P$_2$ (branching ratio 33\,\%) and
307: 5p\,$^3$P$_1$ (branching ratio 67\,\%) states.
308: 
309: 
310: 
311: The line strength of the three 5s$^2$\,$^1$S-5s5p\,$^3$P
312: intercombination transitions range from the relatively high value
313: (7.6\,kHz) of the 689\,nm $^1$S$_0$-$^3$P$_1$ line to the
314: virtually zero value of the $^1$S$_0$-$^1$P$_0$ line for the
315: even isotopes in the absence of external fields. In $^{87}$Sr the
316:  0-0 line has a natural
317: linewidth of
318: $\sim 1$\,mHz due to hyperfine mixing.
319: 
320: 
321: 
322: 
323: 
324: 
325: \section{Laser cooling and trapping}
326: \label{MOT}
327: 
328: 
329: A magneto-optical trap operated on the
330: $^1$S$_0$-$^1$P$_1$ line requires the use of a blue laser source,
331: that can be realized through second-harmonic generation from a
332: semiconductor laser, as described in section \ref{BlueLaser}.
333: 
334: 
335: 
336: The final temperature in such a ``blue MOT'' is limited to few mK
337: by the linewidth of the $^1$S$_0$-$^1$P$_1$ transition: the ground
338: state of alkaline-earth even isotopes has no hyperfine or Zeeman structure, and this
339: prevents the application of sub-Doppler cooling techniques.
340: However, the presence of narrow intercombination transitions
341: allows efficient second-stage Doppler cooling down to sub-recoil
342: temperatures, as described in \ref{RedMOT}. In $^{87}$Sr, the
343: ground-state hyperfine structure offers the chance for sub-Doppler
344: cooling on the
345: $^1$S$_0$-$^1$P$_1$ transition, as demonstrated by the JILA
346: group.\cite{Xu2003} Alternatively, it is possible to apply sub-Doppler
347: cooling techniques to atoms trapped in the metastable $^3$P$_2$ level, as
348: already demonstrated on Ca.\cite{Grunert}
349: 
350: 
351: 
352: In this section we illustrate the Sr cooling and trapping in
353: detail, as performed in our laboratory. We start by preparing a mK
354: sample in the blue MOT from a Zeeman-slowed atomic beam, using optical pumping to recycle atoms
355: from the metastable $^3$P$_2$ level. Then we transfer the atoms to
356: a ``red MOT'' operated on the $^1$S$_0$-$^3$P$_1$ intercombination
357: line, where we cool them down to $\mu$K temperatures. Our
358: apparatus allows us to trap different Sr isotopes simultaneously,
359: as described in \ref{mixture}. The final step consists in loading
360: an optical dipole trap.
361: 
362: 
363: 
364: 
365: 
366: 
367: \subsection{Zeeman slowing and atomic beam collimation} \label{BeamCollimation}
368: 
369: 
370: The sequence for cooling and trapping on the 461\,nm resonance
371: line follows standard techniques, as already reported by other
372: groups.\cite{KatoriMOT1999,Xu2003,Kurosu1990,Oates1999} We slow
373: the thermal atomic beam to few tens of m/s in a 30-cm long
374: Zeeman-slower\cite{Prodan1982} based on a two-stages tapered coil
375: with a zero crossing magnetic field, and a counter-propagating
376: laser beam frequency shifted by 480\,MHz to the red of the
377: $^1$S$_0$-$^1$P$_1$ transition. The beam, with typical power of
378: 40\,mW, has a 7\,mm $1/e^2$ radius at the MOT and it is focused on
379: the capillaries (see appendix \ref{VacuumSystem}). The distance
380: between the capillaries and the MOT region is 100\,cm.
381: 
382: 
383: The atomic beam brightness can be increased with a 2-D transverse
384: cooling stage before the Zeeman slower.\cite{Shimizu,Rasel1999}
385: This improves the MOT loading both by increasing the atomic flux
386: coupled into the differential pumping tube (see
387: \ref{VacuumSystem}), and by effectively reducing the final
388: diameter of the atomic beam after the Zeeman slower.
389: 
390: 
391: 
392: We implement a 2D optical molasses with two 461\,nm beams sent
393: perpendicularly to the atomic beam, in multipass geometry on two
394: orthogonal planes. The beams are red detuned by 20\,MHz with
395: respect to the $^1$S$_0$-$^1$P$_1$ resonance. At given optical power, the
396: multi-pass geometry improves the transverse cooling efficiency  by
397: increasing the length of the interaction region. In the present setup we
398: use beams with
399: $1/e^2$ diameter of 2.5\,mm$^2$ bouncing about 12 times to cover an
400: interaction length of 4\,cm. The interaction starts 15\,cm after the
401: capillaries (see
402: \ref{VacuumSystem}). In that region the beam diameter has already
403: reached about 11\,mm.
404: 
405: 
406: 
407: Without beam collimation we can couple into the differential
408: pumping tube (placed 24\,cm far from collimation region) about
409: 10\,\% of the total atomic flux. With the 2D molasses we increase
410: the coupling by a factor of 4 in optimal conditions (laser detuning
411: $\sim 20$\,MHz, optical power $\sim 20$\,mW). This simple scheme can also
412: be used to deflect the atomic beam.
413: 
414: 
415: 
416: 
417: 
418: 
419: 
420: 
421: \subsection{Blue MOT} \label{BlueMOT}
422: 
423: 
424: 
425: The slowed atoms are then cooled and trapped in a MOT operated
426:  on the $^1$S$_0$-$^1$P$_1$ transition 40\,MHz to the red
427: with respect to resonance. Three retro-reflected beams with a
428: 5\,mm $1/e^2$ radius cross in the MOT region with almost mutually
429: orthogonal directions. The vertical beam is collinear with the
430: magnetic quadrupole axis of an anti-Helmholz pair of coils. The
431: field gradient at the quadrupole center is 50\,Gauss/cm. Taking
432: into account the coupling efficiency of the blue laser source into
433: the fiber (70\,\%) and the efficiency of the AOMs ($75 \div
434: 80$\,\%), the remaining 120\,mW are split into the three channels:
435: typical power values are respectively 60\,mW for MOT beams, 40\,mW
436: for Zeeman slower and 20\,mW for transverse cooling beams. The
437: total 461\,nm light incident on the MOT region amounts to $\sim
438: 90$\,mW/cm$^2$.
439: 
440: 
441: 
442: 
443: 
444: We have a rough estimate of the number of trapped atoms by
445: collecting the 461\,nm fluorescence on a calibrated photodiode. By
446: changing the blue laser detuning we are able to separately trap
447: the four natural Sr isotopes, as shown in figure
448: \ref{IsotopeFluorescence}. For a more accurate measure of the atom
449: number we perform absorption imaging on a CCD with a resonant
450: 461\,nm horizontally propagating probe beam. We measure the sample
451: temperature with standard time-of-flight technique.
452: 
453: 
454: 
455: 
456: 
457: \begin{figure}[t]
458: \centerline{\psfig{file=MOTfluor,width=7cm}} %\vspace*{8pt}
459: \caption{Fluorescence of blue MOT as a function of laser detuning.
460: Resonances corresponding to the different isotopes are visible.
461: }\label{IsotopeFluorescence}
462: \end{figure}
463: 
464: 
465: 
466: 
467: 
468: The lowest temperatures measured in the blue MOT are typically
469: higher than the Doppler limit by at least a factor two. The Nice
470: group has shown that the extra-heating mechanisms causing such
471: discrepancy can be explained in terms of intensity fluctuations in the MOT
472: laser beams.\cite{Chanelière2005}
473: 
474: 
475: 
476: 
477: 
478: 
479: 
480: \subsection{Optical repumping from metastable states} \label{OpticalPumping}
481: 
482: 
483: 
484: 
485: 
486: The $^1$S$_0$-$^1$P$_1$ transition used in the first stages of
487: cooling and trapping is not perfectly closed due to the decay
488: channel of the 5s5p\,$^1$P$_1$ towards the 5s4d\,$^1$D$_2$ state,
489: that has a branching ratio of $2 \times 10^{-5}$. Atoms in the
490: latter state may decay back to the ground state trough the
491: 5s5p\,$^3$P$_1$ within less than 1 \,ms, or may decay to the
492: metastable 5s5p\,$^3$P$_2$ and be lost (see figure
493: \ref{LevelScheme}). Under typical MOT conditions this process
494: limits the MOT lifetime to few tens of milliseconds.
495: 
496: 
497: 
498: Some groups already circumvented this limitation by optical
499: pumping atoms from the metastable 5s5p\,$^3$P$_2$ to the ground
500: state via the 5s6s\,$^3$S$_1$ state with 707\,nm
501: light.\cite{Vogel1999} Since the 5s6s\,$^3$S$_1$ state is also
502: coupled to the metastable 5s5p\,$^3$P$_0$ an additional laser at
503: 679 nm is necessary in this scheme. To reduce the number of
504: repumping lasers we use a different approach which involves the
505: 5s5d\,$^3$D$_2$ state and requires one single laser at 497 nm.
506: Concerning the optical pumping, the $^3$D$_2$ state is coupled
507: basically to the $^3$P$_2$ and $^3$P$_1$ states, then efficient
508: pumping is insured within few absorption cycles. To this purpose
509: during the loading we illuminate the blue MOT with light at
510: 497\,nm, produced by the source described in section
511: \ref{BlueLaser}, which is kept on resonance with the repumping
512: transition. Figure \ref{BlueMOTFluorescence} shows the effect of
513: the repumping field on the $^{88}$Sr MOT loading in different
514: conditions. Without the repumper the MOT loading time ($1/e$) is
515: 10.2 ms regardless to the presence of the atomic beam transverse
516: cooling, and correspondingly to the flux of atoms captured. In
517: presence of the repumper we observe a 0.24 s lifetime when few
518: atoms are present in the MOT, and a decreasing lifetime down to
519: 0.11 s at full MOT charging. This reduction in lifetime is
520: explained in terms of light assisted collision.\cite{Vogel1999}
521: 
522: 
523: 
524: 
525: 
526: \begin{figure}[t]
527: \centerline{\psfig{file=MOTChargingBlue,width=9cm}} %\vspace*{8pt}
528: \caption{Fluorescence of the charging MOT. Solid line: with
529: transverse cooling and repumper. Dash: with repumper. Dot: with
530: transverse cooling. Dash-dot: without repumper or transverse
531: cooling.}\label{BlueMOTFluorescence}
532: \end{figure}
533: 
534: 
535: 
536: 
537: 
538: Due to the difference in the natural isotope abundance (see
539: section \ref{Strontiumatom}), the loading flux into the MOT varies
540: correspondingly. In typical conditions, i.e. when operating with
541: the transverse cooling and optical repumping from the metastable
542: state, and with the Sr oven kept at $\sim 700$\,K, the
543: steady-state blue MOT population amounts to $\sim 10^8$ atoms for
544: $^{88}$Sr and $\sim 10^7$ atoms for $^{86}$Sr.
545: 
546: 
547: 
548: 
549: 
550: 
551: 
552: \subsection{Red MOT} \label{RedMOT}
553: 
554: 
555: 
556: The Tokyo group first realized a strontium MOT operated on the
557: $^1$S$_0$-$^3$P$_1$ line.\cite{KatoriMOT1999} Such system has been
558: exhaustively studied by the JILA group, both theoretically and
559: experimentally.\cite{Loftus2004} The dynamics of laser cooling on
560: narrow transitions presents several interesting features. Unlike
561: in the case of ordinary magneto-optical traps, for the
562: $^1$S$_0$-$^3$P$_1$ transition in alkaline-earth atoms the natural
563: linewidth $\Gamma$ is of the order of the recoil frequency shift
564: $\Gamma_R$. In such conditions both mechanical and thermodynamical
565: MOT properties cannot be explained by the ordinary semiclassical
566: Doppler theory of laser cooling. In particular, the role of
567: gravity becomes non negligible, and the atomic temperature can be
568: lower than the recoil limit $T_R = \frac{\hbar \Gamma_R}{k_B}$.
569: 
570: 
571: 
572: When the laser detuning $\delta$ is negative and larger than the
573: power-broadened linewidth $\Gamma\sqrt{1+\frac{I}{I_{sat}}}$, the
574: atoms interact with the MOT laser beams only on a thin ellipsoidal
575: shell, corresponding to the surface where the laser frequency
576: offset compensates for the Zeeman shift. The maximum radiation
577: force is only one order of magnitude larger than gravity. As a
578: result, the atoms sag on the bottom of the ellipsoid, as
579: shown in the inset of figure \ref{RedMOTtemperature}. In such
580: conditions the atomic temperature is mainly determined by the
581: interaction with the upward propagating vertical MOT beam. With
582: typical laser intensities of $I \simeq 10^1 \div 10^4 I_{sat}$,
583: measured temperatures are in good agreement with a semiclassical
584: model, and can be expressed as
585: 
586: 
587: 
588: \begin{equation}
589: T=\frac{\hbar \Gamma \sqrt{1+\frac{I}{I_{sat}}}}{2 k_B},
590: \end{equation}
591: which is lower than the Doppler temperature at the same detuning.
592: At low laser intensity ($I \approx I_{sat}$), the cooling
593: mechanism becomes fully quantum-mechanical, and the minimal
594: attainable temperature is $T_R \over 2$.\cite{Loftus2004} Figure
595: \ref{RedMOTtemperature} shows the dependence of the MOT
596: temperature on the 689\,nm laser intensity, as measured in our
597: laboratory.
598: 
599: 
600: 
601: The JILA group also studied the case of positive laser detuning,
602: revealing the appearance of novel striking phenomena such as
603: momentum-space crystals. They showed that driving the atomic
604: system with 689\,nm light blue detuned from the
605: $^1$S$_0$-$^3$P$_1$ resonance may result in a periodic pattern in
606: the atomic velocity distribution.\cite{Loftus2004}
607: 
608: 
609: 
610: Second-stage cooling of the odd Sr isotope is more complex,
611: due to the hyperfine structure of both ground and excited states.
612: The Tokyo group has shown that using two lasers at 689\,nm
613: it is possible to cool and trap $^{87}$Sr atoms at phase-space
614: densities close to the Fermi degeneracy.\cite{KatoriFermion}
615: 
616: 
617: 
618: We here illustrate the red MOT for bosonic Sr isotopes as realized
619: in our laboratory. The 689\,nm light for the MOT beam (14\,mW) is
620: provided by a slave laser injection-locked with light coming from
621: the stable master laser described in \ref{RedLaser}. The beam spot
622: is then enlarged to the same radius as the 461\,nm MOT beams and
623: overlapped to the blue beams with a dichroic mirror. After that
624: mirror, the two wavelengths share the same broad-band optics:
625: polarizing beam splitters, mirrors and waveplates.
626: 
627: 
628: 
629: The linewidth of the intercombination transition is not sufficient
630: to cover the Doppler broadening corresponding to the velocity
631: distribution of the sample trapped in the blue MOT. An efficient
632: solution consists in broadening the spectrum of the 689\,nm laser
633: field.\cite{KatoriMOT1999} On the contrary, groups working on
634: other alkaline-earth atoms (i.e. Ca and Mg) employ a resonant
635: coupling to some higher level to quench the $^3$P$_1$
636: lifetime,\cite{Binnewies2001} thus increasing the effective
637: strength of the intercombination transition, as the maximum
638: radiation force would be otherwise lower than gravity.
639: 
640: We add a frequency modulation at 50\,kHz in the first 200\,ms. The
641: total frequency span is 2\,MHz, corresponding to 40 sidebands,
642: with an intensity of 120\,$\mu$W/cm$^2$ per sideband (the
643: saturation intensity is $I_{sat} = 3$\,$\mu$W/cm$^2$). The central
644: frequency is red detuned by 1 MHz with respect to resonance. At
645: the end of the FM recapture, we normally obtain a cloud at a
646: temperature of about 20\,$\mu$K. After that we remove the
647: frequency modulation, set a fixed red detuning from
648: $^1$S$_0$-$^3$P$_1$ transition and reduce the beam intensity in
649: the last 60\,ms of cooling. Working at 350\,kHz below resonance
650: and reducing the total light intensity on the MOT to
651: 70\,$\mu$W/cm$^2$, we can then transfer about 10\,\% of the atoms
652: from the blue MOT to the red MOT at temperatures below 1\,$\mu$K.
653: 
654: 
655: 
656: \begin{figure}[t]
657: \centerline{\psfig{file=redMOTtemp,width=9cm}} %\vspace*{8pt}
658: \caption{Measured atomic temperature and population of the red MOT
659: as a function of the laser intensity with our apparatus. The inset
660: shows in-situ and free-fall images of the trapped
661: atoms.}\label{RedMOTtemperature}
662: \end{figure}
663: 
664: 
665: 
666: 
667: 
668: 
669: 
670: \subsection{Cooling and trapping isotopic mixtures} \label{mixture}
671: 
672: 
673: Among the experiments on ultracold atoms, much work is being
674: concentrated on the study of mixtures of different atomic
675: species\cite{Santos1995,Modugno2001,Murdrich2002} or different
676: isotopes of the same
677: species.\cite{Mewesl999,Loftus2001,Honda2002,Stas2004} Mixtures
678: offer a way to exploit collisional physics not applicable in
679: single species samples.\cite{Schreck2001} They also offer
680: additional degrees of freedom, such as sympathetic cooling, in
681: order to achieve the degenerate quantum regime with atoms for
682: which evaporative cooling is not efficient.\cite{Khaykovich2001}
683: 
684: 
685: 
686: For simultaneous trapping of different isotopes previous
687: experiments employed laser sources delivering the necessary
688: frequency components for each isotope
689: involved.\cite{Suptitz1994,Mewesl999} In the case of the strontium
690: blue MOT, this approach may be difficult to apply because of the
691: complexity of the laser sources and the limited laser power. An
692: alterative solution is presented by the use of the magnetically
693: trapped $^3$P$_2$  state as a dark atom
694: reservoir.\cite{Stuhler2001,Nagel2003,Poli2005} During the blue
695: MOT phase without the repumper, the small loss channel of the
696: excited $^1$P$_1$ state towards the metastable $^3$P$_2$ state
697: provides a continuous loading of atoms into the potential given by
698: the MOT's magnetic quadrupole. Figure \ref{BlueIsotopeLifetime}
699: reports our lifetime measurements on the magnetically trapped
700: metastable isotopes. By using the same blue laser source, one can
701: sequentially load different isotopes into the magnetic potential
702: by just stepping the laser frequency to the different resonances.
703: 
704: 
705: 
706: We typically start by collecting one isotope (say $^{86}$Sr) for a
707: few seconds, then we tune the blue laser on resonance to the other
708: isotope (say $^{88}$Sr). Once the isotopic mixture is prepared in
709: the $^3$P$_2$ state, the blue light is switched off, and the FM
710: red MOT is switched on as well as the repumping beam. The isotopic
711: shift on the repumping transition is smaller than the resonance
712: width of the $^3$P$_2$-$^3$D$_2$ transition observed on the blue
713: MOT fluorescence, which results in efficient, simultaneous optical
714: pumping of $^{88}$Sr and $^{86}$Sr on a time scale short with
715: respect to the red MOT capture time. The loading of a single
716: isotope into the magnetic potential was already described by Nagel
717: et al.,\cite{Nagel2003}  and we did not observe significant
718: differences in the behavior when loading two isotopes. Figure
719: \ref{BlueIsotopeLifetime} shows the measurement of the lifetime
720: for each isotope, both when individually trapped and in the
721: presence of the blue MOT working on the other isotope. The
722: measured lifetime values are all of the order of 5\,s, close to
723: the background pressure limited lifetime of 7\,s.
724: 
725: 
726: 
727: 
728: 
729: \begin{figure}[t]
730: \centerline{\psfig{file=lifetime3P2,width=7cm}} %\vspace*{8pt}
731: \caption{Decay of the $^3$P$_2$ state when trapped in a 56\,G/cm magnetic quadrupole. Circles
732: ($^{86}$Sr) and squares $^{88}$Sr refer  to data taken for the individual isotopes with the blue
733: MOT switched off; diamonds ($^{86}$Sr) and triangles ($^{88}$Sr) with the blue MOT operating on
734: the undetected isotope. The measurements are taken after red MOT recapture. Reprinted figure with
735: permission from N. Poli et al., Phys. Rev. A 71, 061403(R) (2005). Copyright (2005) by the
736: American Physical Society.}\label{BlueIsotopeLifetime}
737: \end{figure}
738: 
739: 
740: 
741: The laser source for the operation of the two-isotope red MOT is
742: composed of two slave lasers injected from the same frequency
743: stabilized master with a frequency offset corresponding to the
744: isotopic shift of 163 817.3\,kHz.\cite{Ferrari2003} Subsequently,
745: the frequency and intensity of the two beams are controlled by
746: double pass AOMs driven by the same RF, the beams are superimposed
747: on a polarizing beam splitter, and then they are overlapped to the
748: blue MOT beams as described previously. Comparing the two-isotope
749: red MOT with the single isotope one, with the same atom numbers we
750: do not observe any effect in the transfer efficiency and final
751: temperature due to the presence of the second isotope. In this
752: way, we obtain samples with up to 10$^7$ (10$^6$) atoms of
753: $^{88}$Sr ($^{86}$Sr) at a temperature 2\,$\mu$K (1\,$\mu$K). We
754: attribute the difference in the loading to the natural abundances
755: and to minor differences in the red MOT parameters. By varying the
756: order of loading and the loading times of the two isotopes we can
757: vary almost arbitrarily the final ratio of populations.
758: 
759: 
760: 
761: The atom number and temperature are measured independently on the
762: two isotopes by absorption imaging with the resonant 461\,nm probe
763: beam, and the contribution of the non-resonant isotope is taken
764: into account.
765: 
766: 
767: 
768: 
769: 
770: 
771: 
772: \subsection{Optical dipole trap} \label{FORT}
773: 
774: 
775: 
776: In magneto-optical traps the atomic temperature and lifetime are
777: fundamentally limited by resonant photon scattering and
778: light-assisted collisions. In all experiments requiring long
779: storage times or ultra-low temperatures, it is convenient to
780: transfer the atoms into a conservative trap. The ground-state even
781: isotopes of alkaline-earth atoms cannot be magnetically trapped,
782: due to the absence of Zeeman structure. Though the Tokyo group has
783: recently demostrated a clever scheme for Sr trapping with AC elecric
784: fields,\cite{Katori2006}  in most cases the best choice consists in
785: optical dipole traps. Moreover on two-electrons atoms it is possible to
786: apply light-shift cancellation techniques to the intercombination
787: transitions,\cite{KatoriFORT1999} opening the way to optical spectroscopy
788: with ultimate accuracy. However, the optical dipole trap is widely
789: employed with magnetic atoms as well,\cite{opticaltrapping} since it
790: generally produces a stronger confinement than magnetic traps, up to the
791: Lamb-Dicke regime in optical lattices,\cite{Dicke} and permits trapping
792: in all the magnetic sub-levels.
793: 
794: 
795: 
796: In an optical dipole trap the confining force originates from the
797: energy level gradient produced by the intensity-dependent light
798: shift. The energy level shift $\Delta E$ of an atom in an optical
799: field is proportional to the light intensity $I$ and, in the rotating-wave approximation, inversely
800: proportional to the frequency detuning $\delta$ from resonance:
801: 
802: 
803: 
804: \begin{equation}
805: \Delta E = \frac{\hbar \Gamma^2 I}{8 \delta I_{sat}},
806: \end{equation}
807: where $\Gamma$ and $I_{sat}$ are the linewidth and the saturation
808: intensity of the resonance transition, respectively, while the
809: photon scattering rate $R_S$ scales as the light intensity and the
810: inverse square of the laser detuning:
811: 
812: 
813: 
814: \begin{equation}
815: R_S = \frac{\Gamma^3 I}{4 \delta^2 I_{sat}}.
816: \end{equation}
817: Thus, at a given trap potential depth, it is possible to reduce
818: the heating due to photon scattering by increasing $\delta$ and
819: $I$ proportionally. In far-off resonant optical dipole traps
820: (FORT) the effect of photon scattering is negligible for most
821: practical purposes.
822: 
823: 
824: 
825: The trapping radiation couples singlet and
826: triplet states independently. This results in a different
827: dependence on the trapping wavelength for the light shift of the
828: $^1$S$_0$ and $^3$P$_1$ levels, so that the differential shift
829: vanishes at a ``magic wavelength''. This enables optical cooling
830: into the dipole trap, since the detuning of the cooling radiation
831: is not position-dependent.
832: 
833: 
834: 
835: For our optical dipole trap we employ the apparatus described in appendix
836: \ref{IRlaser}. In most cases we use the horizontally propagating
837: beam alone. This single-beam FORT has a trap depth of $90\,\mu$K.
838: The computed radial and axial trap frequencies are $\omega_r =
839: 2\pi \times 2$ kHz and $\omega_a = 2\pi \times 26$ Hz
840: respectively. The vertical beam produces a maximum potential depth
841: of about $12\,\mu$K. We find good agreement between computed and
842: measured trap frequencies. The resonant 461\,nm probe beam for
843: time-of-flight absorption imaging propagates horizontally at an
844: angle of about $30^{\circ}$ with the horizontal FORT beam.
845: 
846: 
847: 
848: The transfer efficiency from the red MOT to the FORT results from
849: a balance between loading flux and density dependent losses due to
850: light assisted collisions.\cite{KatoriFORT1999} Figure
851: \ref{DipoleTrap} shows typical loading curves, together with an
852: in-situ image of the trapped atoms.
853: 
854: 
855: 
856: 
857: 
858: \begin{figure}[t]
859: \centerline{\psfig{file=dipolo,width=12cm}} %\vspace*{8pt}
860: \caption{Crossed-beams FORT loading. The inset shows an in-situ
861: image of trapped atoms. The graph shows the measured FORT
862: population as a function of the time overlap between FORT and red
863: MOT. Reprinted figure with permission from N. Poli et al., Phys.
864: Rev. A 71, 061403(R) (2005). Copyright (2005) by the American
865: Physical Society.}\label{DipoleTrap}
866: \end{figure}
867: 
868: 
869: 
870: 
871: 
872: The AC Stark shift of the intercombination line depends on the
873: direction of the FORT field with respect to the bias magnetic field
874: $\vec B$. We use the MOT quadrupole field to resolve the Zeeman structure
875: of the $^1$S$_0$-$^3$P$_1$ line, and we keep the polarization of the FORT
876: beam linear and orthogonal to
877: $\vec B$. In such conditions the resulting light shift is not critical
878: for laser cooling in the FORT. The wavelength used for the dipole trap is
879: only 7\,nm away from the ``magic wavelength'' for the intercombination
880: $^1$S$_0$-$^3$P$_1$ transition.\cite{Ido2003} When the polarization of
881: the dipole trapping light is orthogonal to the magnetic field the Stark
882: shift for the $^1$S$_0$-$^3$P$_1$ transition is lower than 10\,kHz.
883: Thus, it is possible to cool the atoms while loading the
884: dipole trap.\cite{Ido2000}
885: 
886: 
887: 
888: 
889: 
890: 
891: 
892: \section{Collisional studies on ground-state even Sr isotopes}
893: \label{CollisionalMeasure}
894: 
895: 
896: 
897: The study of atomic collisions at low temperature has undergone a
898: rapid development in recent years following the advent of a number
899: of cooling techniques for diluite atomic gases. The measurement of
900: collisional parameters gives insight into an interesting and
901: promising physics, allowing tests of theoretical models for
902: interatomic potentials and molecular
903: wavefunctions.\cite{Vogel1999,Fedichev1996} Moreover, a precise
904: knowledge of collisional properties is essential for experiments
905: aiming to explore regimes of quantum degeneracy in atomic
906: gases.\cite{Delannoy2001,Yamashita2003}
907: 
908: 
909: 
910: Here we describe a systematic analysis on the ground-state collisional
911: properties of the two most abundant bosonic Sr isotopes ($^{88}$Sr
912: and $^{86}$Sr) performed with our apparatus. More specifically, we
913: evaluated the elastic cross-sections $\sigma_{i-j}$ ($i,j =86,88$)
914: for both intra and inter-species collisions, and the three-body
915: recombination coefficients $K_{i}$ ($i =86,88$) in the FORT. The
916: elastic cross-sections were deduced by driving the system out of
917: thermal equilibrium and measuring the thermalization rate together
918: with the sample density. For the inelastic collisions, we measured
919: the density dependence of the trap loss rate. The key point of our
920: experimental procedure is a precise knowledge of the atomic
921: density. This in turn requires a proper control of the trap
922: frequencies. We assumed the equilibrium phase-space distribution
923: in an ideal harmonic trap to infer the atomic density from the
924: measured number of atoms $N$ and temperature $T$. Our assumption
925: is supported by the fact that the ratio $\eta=\frac{U}{k_B T}$ of
926: trap depth and sample temperature was larger than 5 in all of our
927: measurements. The equilibrium peak atomic density in a harmonic trap is
928: 
929: 
930: 
931: \begin{equation}
932: n_0 = N {\bar \nu}^3 \left({2 \pi m \over k_{B}T}\right)^{3 \over
933: 2}
934: \end{equation}
935: where $\bar \nu$ is the average trap frequency, $m$ is the atomic
936: mass and $k_B$ is the Boltzmann constant. Our results show
937: significant differences in the collisional properties of the two
938: isotopes. Both the elastic cross-section and the three-body
939: collision coefficient were found to be several orders of magnitude
940: larger in $^{86}$Sr than in $^{88}$Sr, and the inter-species
941: cross-section $\sigma_{86-88}$ is much larger than the
942: intra-species cross-section $\sigma_{88-88}$.
943: 
944: 
945: We adopted the standard technique of observing the
946: cross-thermalization between orthogonal degrees of
947: freedom.\cite{Monroe1993,Arndt1997,Hopkins2000,Schmidt2003} The
948: single-beam FORT geometry is not well suited for this experiment,
949: as the temperature in the horizontal modes cannot be univocally
950: determined with our apparatus, and because with $^{86}$Sr in our
951: experimental conditions the system turned out to be in
952: hydrodinamic regime along the axial
953: direction.\cite{Odelin1999,Gensemer2001} Given these constraints,
954: we chose the crossed-beams geometry for the thermalization
955: measurements.
956: 
957: 
958: 
959: In order to drive the system out of thermal equilibrium we
960: performed a selective optical cooling along the vertical direction
961: with resonant 689 nm light. For this purpose we took advantage of
962: the fact that the FORT is located at the center of the red MOT,
963: where atoms are not resonant with the horizontal MOT
964: beams.\cite{Loftus2004} We chose the optical intensity and cooling
965: time in such way to create a detectable thermal anisotropy without
966: introducing dramatic losses.
967: 
968: 
969: 
970: We measured the intra-species thermalization rate $\tau_{th}$ by
971: separately loading one isotope (either  $^{88}$Sr or $^{86}$Sr) and
972: observing the temporal evolution towards thermal equilibrium (see
973: figure \ref{86Thermalization}). In both cases the total trap loss
974: rate was much lower than the thermalization rate, so the atom
975: number was constant within shot-to-shot fluctuations.
976: 
977: 
978: 
979: 
980: 
981: \begin{figure}[t]
982: \centerline{\psfig{file=86therm,width=7cm}} %\vspace*{8pt}
983: \caption{Measurement of the thermalization rate for $^{86}$Sr. The
984: graph shows the temporal evolution of the horizontal and vertical
985: temperatures after cooling on the vertical direction. The solid
986: curve is an exponential fit to the vertical temperature
987: data.}\label{86Thermalization}
988: \end{figure}
989: 
990: 
991: 
992: 
993: 
994: We deduced the elastic cross-section from the measured
995: thermalization rate, by assuming a value of 2.7 for the average
996: number of collisions to reach thermalization, that is,
997: $\tau_{th}=2.7
998: \tau_{coll}$.\cite{Monroe1993,Arndt1997,Kavoulakis2000} The
999: collision rate is given by
1000: 
1001: 
1002: 
1003: \begin{equation}
1004: {1\over\tau_{coll}} = \bar n \sigma \bar v
1005: \end{equation}
1006: where $\bar n$ is the average atom density, $\bar v$ is the
1007: average relative velocity of colliding atoms, and $\sigma$ is the
1008: cross-section. We computed $\bar n$ and $\bar v$ from the measured
1009: values for atom number $N$, average sample temperature $T$ and
1010: average trap frequency $\bar \nu$. With our assumptions the
1011: average density is given by $\bar n = {n_0 / 2 \sqrt 2}$, while
1012: the average relative velocity is given by $\bar v = 4 \sqrt {k_B
1013: T/\pi m}$.
1014: 
1015: 
1016: 
1017: We repeated our measurements for different values of the atom
1018: density, in order to check whether the observed
1019: cross-thermalization was due to elastic collisions or ergodic
1020: mixing between different degrees of freedom. The resulting values
1021: are $\sigma_{88-88} = 3(1) \times 10^{-13}$ cm$^2$  and
1022: $\sigma_{86-86} = 1.3(0.5) \times 10^{-10}$ cm$^2$. The
1023: uncertainty is mainly due to shot-to-shot fluctuations in the FORT
1024: population, that reflect in both density and temperature
1025: instabilities. Such results might be compared with the
1026: zero-temperature cross sections deduced from scattering length
1027: values through the relation $\sigma=8\pi\lambda^2$, where
1028: $\lambda$ is the $s$-wave scattering length. Scattering length
1029: values were obtained from photoassociation spectra by the Tokyo
1030: group for $^{88}$Sr,\cite{yasuda2004} and by the Houston group for
1031: both isotopes.\cite{Mickelson2005} As concerning $\sigma_{88-88}$,
1032: our value is consistent with the Tokyo work, while the Houston
1033: group predicts an even smaller cross-section. On the contrary,
1034: there is a fair agreement between our $\sigma_{86-86}$ value and
1035: the Houston work. All of these results are summarized in tab. 2
1036: 
1037: 
1038: 
1039: \begin{table}[pt]
1040: \tbl{Intra-species elastic cross-section for $^{88}$Sr and
1041: $^{86}$Sr, in cm$^2$} {\begin{tabular}{@{}cccc@{}} \toprule
1042: Reference &
1043: $\sigma_{88-88}$  & $\sigma_{86-86}$  & Method  \\
1044: \colrule
1045: Tokyo group\cite{yasuda2004} &  \, $3(1) \times 10^{-13}$ &  & $\lambda$ from PA spectra \\
1046: Houston group\cite{Mickelson2005} &  $<1.2\times 10^{-13}$ & $2.6 \times 10^{-10} < \sigma < 3.7 \times 10^{-9}$ & $\lambda$ from PA spectra \\
1047: Our group\cite{Ferrari2006} & \, $3(1) \times 10^{-13}$ & $1.3(0.5) \times 10^{-10}$ & cross-thermalization \\
1048: \botrule
1049: \end{tabular}}
1050: \end{table}
1051: 
1052: 
1053: 
1054: 
1055: 
1056: We adopted a similar approach for the measurement of the
1057: inter-species cross section.\cite{Goldwin,K-Rb,Mosk,Delannoy} The
1058: single-beam FORT was loaded with a mixture of $^{88}$Sr ($10^5$
1059: atoms) and $^{86}$Sr ($4 \times 10^4$ atoms). We then applied a
1060: selective optical cooling stage (duration 5\,ms) to $^{88}$Sr and
1061: observed the temperature evolution of the two samples (see figure
1062: \ref{MixThermalization}). The $^{88}$Sr is heated by $^{86}$Sr to
1063: the equilibrium temperature in few tens of ms. For comparison, in
1064: absence of $^{86}$Sr the temperature of $^{88}$Sr grows by less
1065: than 5\,\% in 100\,ms. The resulting inter-species cross-section
1066: was found as $\sigma_{88-86} = 4(1) \times 10^{-12}$ cm$^2$. This
1067: value is significantly larger than the intra-species cross-section
1068: $\sigma_{88-88}$, suggesting the way to a novel and efficient
1069: sympathetic cooling mechanism.\cite{Ferrari2006}
1070: 
1071: 
1072: 
1073: \begin{figure}[t]
1074: \centerline{\psfig{file=mixtherm,width=7cm}} %\vspace*{8pt}
1075: \caption{Measurement of the interspecies thermalization rate. The
1076: graph shows the temperature evolution of $^{88}$Sr and $^{86}$Sr
1077: after selective cooling on $^{88}$Sr. The solid curve is an
1078: exponential fit to the $^{88}$Sr data. Trap populations are $10^5$
1079: atoms for $^{88}$Sr and $4.5 \times 10^4$ atoms for
1080: $^{86}$Sr.}\label{MixThermalization}
1081: \end{figure}
1082: 
1083: 
1084: 
1085: We studied the inelastic collisions by loading a single isotope in
1086: the single-beam FORT  (either  $^{88}$Sr or $^{86}$Sr) and looking
1087: at the evolution of the number and temperature of trapped atoms in
1088: the FORT after the MOT operation was finished. We found no
1089: evidence for non-exponential decay with $^{88}$Sr. The measured
1090: lifetime of 7\,s is consistent with the residual background gas
1091: pressure of $10^{-8}$\,torr. With the initial atom density at trap
1092: center of $3 \times 10^{13}$ cm$^{-3}$ this gives an upper limit
1093: of $10^{-27}$ cm$^6$s$^{-1}$ for the $K_{88}$ coefficient.
1094: 
1095: 
1096: 
1097: 
1098: 
1099: \begin{figure}[t]
1100: \centerline{\psfig{file=recomb86,width=9cm}} %\vspace*{8pt}
1101: \caption{Decay of trap population when loading $^{86}$Sr atoms
1102: alone in the single-beam FORT. The sample temperature is
1103: 12\,$\mu$K, trap depth is 90\,$\mu$K.}\label{Recombination86}
1104: \end{figure}
1105: 
1106: 
1107: 
1108: 
1109: 
1110: When loading the single-beam FORT with $^{86}$Sr only, we observed
1111: a clear non-exponential decay (figure \ref{Recombination86}). We
1112: deduced the three-body recombination coefficient from the density
1113: dependence of the loss rate.\cite{Burt} Integrating the loss rate
1114: equation
1115: 
1116: 
1117: 
1118: \begin{equation}
1119: \dot N = -\Gamma_b N - K_{86}{\int {n^3 (\vec r, t) d^3 \vec r}}
1120: \end{equation}
1121: where $\Gamma_b$ is the linear loss rate for background gas
1122: collisions, with our assumptions we can write
1123: 
1124: 
1125: 
1126: \begin{equation}
1127: \ln {N \over N_0} = -\Gamma_b t - {K_{86} (2\pi)^3 m^3 \bar \nu ^6
1128: \over  3^{3\over 2}}{\int {{N^2 (t) \over [k_B T(t)]^3} dt}}
1129: \end{equation}
1130: This formula is valid as long as additional losses due to
1131: evaporation can be neglected. In order to limit such effect, we
1132: select the coldest atoms by means of forced evaporation before the
1133: measurement.
1134: 
1135: 
1136: 
1137: We deduced the linear loss rate from the wing of figure
1138: \ref{Recombination86}. This value is consistent with the loss rate
1139: measured with $^{88}$Sr. Then we performed a linear fit of $\ln {N
1140: \over N_0} + \Gamma_b t$ versus $\int {{N^2 (t) \over [k_B
1141: T(t)]^3} dt}$ to derive the recombination constant. We repeated
1142: the measurement several times to average out density fluctuations
1143: reflecting in large uncertainty on $K_{86}$. The final result was
1144: $K_{86} = 1.0(0.5) \times 10^{-24}$ cm$^6$s$^{-1}$.
1145: 
1146: 
1147: 
1148: 
1149: 
1150: 
1151: 
1152: 
1153: 
1154: \section{Towards a BEC of strontium} \label{SrBEC}
1155: 
1156: 
1157: 
1158: 
1159: Laser cooling is a very effective technique to reach phase-space
1160: densities within few orders of magnitude from quantum degeneracy.
1161: The limits in cooling at high density are set by the optical depth
1162: of the sample, i.e. reabsorption of scattered light, and
1163: light-assisted atom-atom collisions. Forced evaporative cooling
1164: represents the common way to circumvent these
1165: limits.\cite{Hess1986} However, this procedure is not effective
1166: with all atoms. In particular, among the atoms cooled with optical
1167: methods, none of the alkali-earth atoms reached quantum degeneracy
1168: so far, except ytterbium which has an alkali-earth-like electronic
1169: structure.\cite{Takasu2003} A phase-space density of $\simeq
1170: 10^{-1}$ was reported for Sr but BEC could not be
1171: reached.\cite{Ido2000}
1172: 
1173: 
1174: 
1175: On this respect, the results of the collisional measurements
1176: reported in \ref{CollisionalMeasure} suggest that evaporative
1177: cooling on pure samples of either $^{86}$Sr or $^{88}$Sr cannot be
1178: very efficient in our experimental conditions. $^{86}$Sr presents
1179: an extremely large elastic cross-section, a good point to
1180: establish a fast thermalization, but the 3-body recombination rate
1181: introduces a loss channel that is fatal with the typical
1182: geometries accessible through optical dipole trapping. An optical
1183: trap with a much larger trapping volume would partially suppress
1184: this loss channel.\cite{Weber2002} $^{88}$Sr instead turns out to
1185: be stable against 3-body decay, but the small elastic
1186: cross-section results in a long thermalization time compared to
1187: typical trap lifetime.
1188: 
1189: 
1190: 
1191: On the other hand, our results suggest a novel all-optical sympathetic
1192: cooling scheme.\cite{Ferrari2006} In the isotope mixture the relatively
1193: large inter-species cross-section results in thermalization times
1194: typically of the order of few milliseconds. This thermalization is
1195: fast even on the time scale of laser cooling on the
1196: intercombination $^1$S$_0$-$^3$P$_1$ transition. Moreover, due to the
1197: 164\, MHz
1198: $^{86}$Sr-$^{88}$Sr isotopic shift and the natural linewidth of 7.6\,kHz
1199: for the $^1$S$_0$-$^3$P$_1$ transition, laser cooling on one
1200: isotope has negligible effect on the other one. It
1201: is then possible to cool sympathetically a dense and optically thick cloud
1202: of one isotope (for instance
1203: $^{88}$Sr) via optical cooling of a small sample of the other
1204: isotope ($^{86}$Sr). Continuous laser cooling of $^{86}$Sr
1205: provides heat dissipation in the sample, while the small optical
1206: depth on $^{86}$Sr does not limit the achievable minimum
1207: temperature. Sympathetic cooling with neutral atoms normally
1208: requires a thermal bath with heat capacity large with respect to
1209: that of the sample to be cooled. This is due to the fact that when
1210: the thermal bath is cooled by evaporative cooling, each lost atom
1211: carries an energy of the order of few times the temperature of the
1212: sample. In the case of optical-sympathetic cooling, each
1213: laser-cooled atom can subtract an energy of the order of the
1214: optical recoil in a time corresponding to a few lifetimes of the
1215: excited state, without being lost.
1216: 
1217: 
1218: 
1219: \begin{figure}[t]
1220: \centerline{\psfig{file=sympatheticdynamics,width=7cm}} %\vspace*{8pt}
1221: \caption{Dynamics of an optically trapped $^{88}$Sr cloud
1222: sympathetically cooled with laser cooled $^{86}$Sr. Filled
1223: circles: $^{88}$Sr atom number. Open circles: $^{86}$Sr atom
1224: number. The number of $^{88}$Sr atoms remains constant during the
1225: process, while $^{86}$Sr decays exponentially with a 80\,ms time
1226: constant. Under optimized conditions, the temperature (triangles)
1227: decreases with a 150\,ms time constant and the mixture is always
1228: at thermal equilibrium. Reprinted figure with permission from G.
1229: Ferrari et al., Phys. Rev. A 73, 023408 (2006). Copyright (2006)
1230: by the American Physical Society.}\label{SimpatheticCooling}
1231: \end{figure}
1232: 
1233: 
1234: 
1235: 
1236: 
1237: We implemented the optical-sympathetic cooling scheme by extending
1238: the temporal overlap between the FORT and the $^{86}$Sr red MOT
1239: after switching off the $^{88}$Sr MOT. Figure
1240: \ref{SimpatheticCooling} reports the dynamics of
1241: optical-sympathetic cooling, starting from an initial temperature
1242: of 15-20\,$\mu$K, limited by density dependent
1243: heating.\cite{Poli2005,Weber2002} It can be observed that the
1244: cooling process does not induce losses on $^{88}$Sr while the
1245: number of $^{86}$Sr atoms exponentially decays with a 80\,ms
1246: lifetime. About 100\,ms after switching off the $^{88}$Sr red MOT,
1247: we observe that the mixture attains thermal equilibrium.
1248: 
1249: 
1250: 
1251: 
1252: 
1253: Under optimized conditions (overall optical intensity $100
1254: I_{sat}$) the temperature decays from the initial value with a
1255: 150\,ms time constant. The minimum attainable temperature depends
1256: both on the intensity of the $^{86}$Sr cooling beam, and the
1257: $^{88}$Sr density. By keeping the cooling parameters on $^{86}$Sr
1258: fixed at the optimum value and by varying the number of trapped
1259: $^{88}$Sr, we determined the dependence of the final temperature
1260: on the $^{88}$Sr density. For $6 \times 10^5$
1261: $^{88}$Sr atoms trapped in the FORT, the final temperature is
1262: 6.7\,$\mu$K at a peak density of $1.3 \times 10^{14}$ cm$^3$; the
1263: corresponding phase-space density is $5 \times 10^{-2}$. This
1264: value is only a factor of two lower than what was
1265: previously obtained without forced evaporation,\cite{Ido2000} but
1266: it exhibits more favorable conditions for starting evaporative
1267: cooling, considering the higher spatial density (more than one
1268: order of magnitude higher) and the number of trapped atoms (gain
1269: $2 \div 10$).
1270: 
1271: 
1272: Indeed we applied an evaporative
1273: cooling stage on $^{88}$Sr by reducing the FORT intensity after the optical sympathetic cooling.
1274: This produced an
1275: increase in phase-space density by roughly a factor of 4, giving a
1276: maximum of $2 \times 10^{-1}$. Such result is consistent with a numerical simulation of
1277: forced evaporation in our experimental conditions. The gain in phase-space density during evaporative
1278: cooling is basically limited by the  $^{88}$Sr elastic cross-section. A promising way towards
1279:  Bose-Einstein condensation seems to be the use of a dipole trap with variable
1280: geometry, to compensate for the reduction in thermalization rate during
1281: evaporation.\cite{Weiss2003}
1282: 
1283: 
1284: 
1285: %\begin{figure}[t]
1286: %\centerline{\psfig{file=asymptemperature,width=9cm}} %\vspace*{8pt}
1287: %\caption{Asymptotic temperature of the $^{88}$Sr cloud for
1288: %different intensities of the $^{86}$Sr laser cooling beam on the
1289: %$^1$S$_0$-$^3$P$_1$ transition ($I_{sat} = 3 \mu$W/cm$^2$). The
1290: %$^{88}$Sr atom number is $6 \times
1291: %10^5$.}\label{AsymptoticTemperature}
1292: %\end{figure}
1293: 
1294: 
1295: 
1296: 
1297: 
1298: \begin{figure}[t]
1299: \centerline{\psfig{file=phasespacedensity,width=7cm}} %\vspace*{8pt}
1300: \caption{Temperature and phase-space density of the $^{88}$Sr
1301: cloud sympathetically cooled with laser cooled $^{86}$Sr, as a
1302: function of the $^{88}$Sr atom number. Reprinted figure with
1303: permission from G. Ferrari et al., Phys. Rev. A 73, 023408 (2006).
1304: Copyright (2006) by the American Physical
1305: Society.}\label{PhaseSpaceDensity}
1306: \end{figure}
1307: 
1308: 
1309: 
1310: 
1311: 
1312: Figure \ref{PhaseSpaceDensity} shows the dependence of $^{88}$Sr
1313: temperature after the optical sympathetic cooling (without
1314: evaporative cooling)
1315:  on the number of atoms
1316: in the trap. We determine the density-dependent heating $dT/dn \simeq 2
1317: \mu$K/($10^{14}$\,cm$^{-3}$), which is 20 times lower than the
1318: equivalent value for pure laser cooled
1319: $^{88}$Sr.\cite{KatoriMOT1999} This large reduction is a direct
1320: consequence of the strong selectivity of the $^1$S$_0$-$^3$P$_1$
1321: transition with respect to the two isotopes. The limit on the
1322: $^{86}$Sr temperature of $4 \mu$K for zero $^{88}$Sr density can
1323: be attributed to the laser cooling dynamics in the tightly
1324: confining potential of the FORT.
1325: 
1326: 
1327: 
1328: 
1329: 
1330: 
1331: 
1332: \section{Ultracold Sr atoms as quantum sensors} \label{Microsensor}
1333: 
1334: 
1335: 
1336: Ultracold atomic strontium is particularly suited for applications
1337: in the field of quantum sensors. Atom interferometry has already
1338: been used with alkali-metals for precision inertial
1339: sensing,\cite{Peters1999,Gustavson2000} for measuring fundamental
1340: constants,\cite{Wicht2002,Clade2006,Stuhler2003} and testing
1341: relativity.\cite{Fray2004} The extremely small size of ultracold
1342: atomic samples enables precision measurements of forces at
1343: micrometer scale. This is a challenge in physics for studies of
1344: surfaces, Casimir effect,\cite{Antezza2005} and searches for
1345: deviations from Newtonian gravity predicted by theories beyond the
1346: standard model.\cite{Long2003,Dimopoulos2003,Samullin2005}
1347: 
1348: 
1349: 
1350: An interesting class of quantum devices is represented by
1351: ultracold atoms confined in an optical lattice, that is a dipole
1352: trap created by a laser standing wave.\cite{Bloch2005} In
1353: particular, Bloch oscillations were predicted for electrons in a
1354: periodic crystal potential in presence of a static electric
1355: field\cite{Bloch1929} but could not be observed in natural
1356: crystals because of the scattering of electrons by the lattice
1357: defects. They were directly observed using atoms in an optical
1358: lattice.\cite{Raizen1997}
1359: 
1360: 
1361: 
1362: The combination of the periodic optical potential and a linear
1363: potential produced by a constant force $F$ along the lattice
1364: wave-vector gives rise to Bloch oscillations at frequency $\nu_B$
1365: given by $\nu_B = \frac{F \lambda_L}{2h}$, where $\lambda_L$ is
1366: the wavelength of the light producing the lattice, and $h$ is
1367: Planck’s constant. Since $\lambda_L$ is well known, the force
1368: along the lattice axis can be accurately determined by measuring
1369: the Bloch frequency $\nu_B$. In order to perform a sensitive force
1370: measurement, a long coherence time with respect to the measurement
1371: duration is required. The most common effects limiting the
1372: coherence time for ultracold atoms are perturbations due to
1373: electromagnetic fields and atom-atom interactions. $^{88}$Sr is in
1374: this respect an ideal choice because in the ground state it has
1375: zero orbital, spin and nuclear angular momentum. This makes it
1376: virtually insensitive to stray magnetic fields. In addition, as
1377: shown in section \ref{CollisionalMeasure} $^{88}$Sr has remarkably
1378: small atom-atom interactions. Such properties make Sr in optical
1379: lattices a unique sensor for small-scale forces with better
1380: performances and reduced complexity compared to proposed schemes
1381: using degenerate Bose or Fermi gases.\cite{Anderson1998,Roati2004}
1382: This enables new experiments on gravity in unexplored regions.
1383: 
1384: 
1385: 
1386: We tested such idea by measuring the gravitational acceleration in
1387: our laboratory with a $^{88}$Sr sample in a vertical optical
1388: lattice. To this end, we cool $\sim 5 \times 10^5$ atoms in the
1389: red MOT down to the recoil temperature (see section \ref{RedMOT}),
1390: so that the vertical momentum distribution is narrower than the
1391: width of the first Brillouin zone.\cite{Bloch1929}. Then we
1392: release the atoms from the MOT and we switch on adiabatically a
1393: one-dimensional optical lattice.
1394: 
1395: 
1396: The lattice potential is originated by a single-mode
1397: frequency-doubled Nd:YVO$_4$ laser ($\lambda_L = 532$\,nm)
1398: delivering up to 350\,mW on the atoms with a beam waist of $200
1399: \mu$m. The beam is vertically aligned and retro-reflected by a
1400: mirror producing a standing wave with a period
1401: $\frac{\lambda_L}{2} = 266$\,nm. The corresponding photon recoil
1402: energy is $E_R = \frac{h^2}{2m\lambda^2} = k_B \times 381$\,nK. We
1403: populate about 100 lattice sites with $2 \times 10^5$ atoms at an
1404: average spatial density of $\sim 10^{11}$\,cm$^{-3}$. We leave the
1405: atoms in the optical lattice for a variable time $t$, then we
1406: switch off the lattice adiabatically and we measure the momentum
1407: distribution of the sample by time-of-flight imaging, after a free
1408: fall of 12\,ms.
1409: 
1410: 
1411: 
1412: We integrate along the
1413: horizontal direction the optical thickness obtained by absorption imaging.
1414: The resulting curve gives
1415: the vertical momentum distribution of the atomic sample: in figure
1416: \ref{TwoPeaks} we show a typical plot after the integration.
1417: 
1418: 
1419: 
1420: 
1421: 
1422: \begin{figure}[t]
1423: \centerline{\psfig{file=TwoPeakDistrib,width=7cm}} %\vspace*{8pt}
1424: \caption{Vertical momentum distribution of the atoms at the Bragg
1425: reflection.}\label{TwoPeaks}
1426: \end{figure}
1427: 
1428: 
1429: 
1430: 
1431: 
1432: We fit the measured momentum distribution  with the sum of two
1433: Gaussian functions. From each fit we extract the vertical momentum
1434: center of the lower peak and the width of the atomic momentum
1435: distribution. We find that the latter is less sensitive against
1436: noise-induced perturbations to the vertical momentum. We can
1437: observe $\sim 4000$ Bloch oscillations in a time $t = 7$\,s (see
1438: figure \ref{BlochOsc}), with a damping time $\tau\sim 12$\,s. To
1439: our knowledge, the present results for number of Bloch
1440: oscillations, duration, and the corresponding number of coherently
1441: transferred photon momenta, are by far the highest ever achieved
1442: experimentally in any physical system.
1443: 
1444: 
1445: 
1446: 
1447: 
1448: \begin{figure}[t]
1449: \centerline{\psfig{file=fig13.EPS,width=9cm}} %\vspace*{8pt}
1450: \caption{Time evolution of the width of the atomic momentum
1451: distribution, showing Bloch oscillation of $^{88}$Sr atoms in the
1452: vertical 1-dimensional optical lattice under the effect of
1453: gravity. From the the data fit, a Bloch frequency $\nu_B =
1454: 574.568(3)$\,Hz is obtained with a damping time $\tau\sim 12$\,s
1455: for the oscillations.}\label{BlochOsc}
1456: \end{figure}
1457: 
1458: 
1459: 
1460: From the measured Bloch frequency $\nu = 574.568(3)$\,Hz we
1461: determine the gravity acceleration along the optical lattice
1462: $g_{meas} = 9.80012(5)$\,ms$^{-2}$.
1463: %The expected value at the
1464: %laboratory site is $g_{ref} =
1465: %9.805046(9)$\,ms$^{-2}$.\cite{Petelski} The disagreement by $5
1466: %\times 10^{-4} g$ is due to a vertical misalignment of $\sim
1467: %20$\,mrad of the optical lattice beam.
1468: The estimated sensitivity is $5 \times 10^{-6} g$. We expect that
1469: such precision may be increased by one order of magnitude by using
1470: a larger number of atoms, and reducing the initial temperature of
1471: the sample.
1472: 
1473: 
1474: 
1475: 
1476: 
1477: 
1478: 
1479: 
1480: \section{Frequency measurements on the Sr intercombination lines}
1481: 
1482: \label{1S3PFreqMeasure}
1483: 
1484: 
1485: 
1486: The advent of laser cooling techniques had important consequences
1487: on time-frequency metrology, by reducing the uncertainty on the
1488: frequency of atomic clocks caused by atom motion.
1489: 
1490: 
1491: 
1492: The recent progresses in two related fields, namely
1493: high-resolution laser spectroscopy and direct optical-frequency
1494: comb generation, opened the way to a new generation of frequency
1495: standards based on transitions in the optical domain. The use of
1496: frequency combs based on self-mode-locked femtosecond lasers has
1497: made possible, for the first time, relatively simple
1498: optical-frequency measurements.\cite{Udem1999,Diddams2000} On the
1499: other hand, the realization of lasers with ultra-narrow emission
1500: band now enables spectroscopy on forbidden optical transitions
1501: with quality factor $Q = \frac{\nu}{\Delta\nu}$ in excess of
1502: $10^{15}$.
1503: 
1504: 
1505: 
1506: Because of their higher frequency, optical transitions have the
1507: potential for greatly improved accuracy and stability compared to
1508: conventional atomic clocks based on microwave frequency
1509: transitions.\cite{Udem2002} Different transitions are now
1510: considered as optical-frequency standards, involving single ions
1511: and neutral atoms.\cite{Udem2001,Gill2005} While single ions offer
1512: an excellent control on systematic effects, clouds of laser cooled
1513: atoms have the potential for extremely high precision, as the
1514: large number of atoms reduces the quantum projection noise.
1515: Perhaps the use of optical lattices at the magic wavelength to
1516: confine neutral atoms in the Lamb-Dicke regime, as proposed for
1517: the first time on strontium,\cite{KatoriFORT1999,Katori2003}
1518: summarizes the best of both worlds; that is, a large number of
1519: quantum absorbers with negligible shift of the optical clock
1520: transition due to external fields, Doppler effect and collisions.
1521: 
1522: 
1523: 
1524: Among the neutral atoms, Sr has long been considered as one of the
1525: most interesting candidates.\cite{Hall1989} Several features, some
1526: of which are specific to this atom, allow different possibilities
1527: for the realization of a high precision optical clock. The visible
1528: intercombination 5$^1$S-5$^3$P lines from the ground state are
1529: easily accessible with semiconductor lasers. Depending on the
1530: specific fine-structure component and on the isotope,  a wide
1531: choice of transitions with different natural linewidths is
1532: possible (see section \ref{Strontiumatom}).
1533: 
1534: 
1535: 
1536: The first phase-coherent absolute frequency measurement of a Sr
1537: intercombination line was performed by our group on the
1538: 5$^1$S$_0$-5$^3$P$_{1}$ transition, using a thermal atomic
1539: beam.\cite{Ferrari2003} This represented an improvement by several
1540: orders of magnitude with respect to previous data.\cite{Tino1992}
1541: Our result is in agreement with subsequent phase-coherent
1542: measurements performed by the BNM-SYRTE group on a thermal
1543: sample,\cite{Courtillotclock} and by the JILA group on a
1544: free-falling ultracold sample using a red MOT.\cite{Ido2005} The
1545: JILA measurement provided a further improvement in the accuracy by
1546: more than two orders of magnitude.
1547: 
1548: 
1549: 
1550: However, the $^1$S$_0$-$^3$P$_{1}$ transition is not best suited
1551: as a final frequency reference, due to its natural linewidth of
1552: 7.6\,kHz. Some groups recently began working on the 698\,nm
1553: $^1$S$_0$-$^3$P$_{0}$ line, that is strictly forbidden in the even
1554: isotopes since it is a $J=0\to J=0$ transition. In $^{87}$Sr the
1555: hyperfine mixing enables direct $^1$S$_0$-$^3$P$_{0}$ excitation
1556: with a transition probability of $\sim 1 $\,mHz. The BNM-SYRTE
1557: group first measured such transition in a blue MOT with an
1558: uncertainty of 15\,kHz.\cite{Courtillotclock} The Tokyo group
1559: performed the first absolute frequency measurement in an optical
1560: lattice on this transition,\cite{Takamoto2005} followed by the
1561: JILA and BNM-SYRTE groups.\cite{Ludlow2005,Targat2006} All groups
1562: estimated an uncertainty $\leq 20$\,Hz for the absolute transition
1563: frequency. The corresponding 578\,nm $^1$S$_0$-$^3$P$_{0}$
1564: transition in Ytterbium was observed at NIST in a $\sim 70 \mu$K
1565: sample on the two odd isotopes, $^{171}$Yb and $^{173}$Yb, with an
1566: uncertainty of $\sim 4$\,kHz.\cite{Hoyt2005}
1567: 
1568: 
1569: 
1570: In spite of its large quality factor, the $^1$S$_0$-$^3$P$_{0}$
1571: transition in odd Sr and Yb isotopes suffers from residual
1572: sensitivity to stray magnetic fields and optical lattice
1573: polarization, besides a complex line structure due to the presence
1574: of many magnetic sublevels. Several groups are now looking at the
1575: even Sr or Yb isotopes as the best candidates to represent optical
1576: frequency standards based on neutral atoms. Some groups have
1577: proposed different methods to directly excite the clock transition
1578: on the even Sr or Yb isotopes, by properly engineering the atomic
1579: level structure to create a finite and controllable
1580: $^1$S$_0$-$^3$P$_{0}$ transition probability. These methods
1581: basically consist in coupling the metastable $^3$P$_{0}$ level to
1582: other electronic states by using either multiple near-resonant
1583: laser fields,\cite{Santra2005,Hong2005} or simply a small static
1584: magnetic field.\cite{Taichenachev2005} The latter scheme has been
1585: experimentally demonstrated on $^{174}$Yb at
1586: NIST.\cite{Barber2005} The possible instability due to
1587: site-to-site tunneling in optical lattice clocks has been
1588: addressed by the BNM-SYRTE group.\cite{Lemonde2005} For accuracy
1589: goals at the $10^{-18}$ level they propose the use of vertical
1590: optical lattices, in order to lift the degeneracy between adjacent
1591: potential wells.
1592: 
1593: 
1594: 
1595: 
1596: 
1597: 
1598: 
1599: \subsection{Frequency measurement on the $^1$S$_0$-$^3$P$_1$
1600: transition with a thermal beam} \label{OpticalFreqMeasure}
1601: 
1602: 
1603: 
1604: 
1605: 
1606: In this section we discuss our precision frequency
1607: measurements on the intercombination 5$^1$S$_0$-5$^3$P$_{1}$
1608: transition.\cite{Ferrari2003} Using a femtosecond laser comb, we
1609: determined the absolute frequency of the transition for $^{88}$Sr
1610: and $^{86}$Sr and an accurate value for the isotope shift.
1611: 
1612: 
1613: 
1614: 
1615: 
1616: The frequency measurements have been performed through saturation
1617: spectroscopy on a thermal beam with the apparatus described in
1618: appendix \ref{RedLaser}. As frequency-comb generator we employed a
1619: commercial system based on a Kerr-lens mode-locked Ti:Sa laser
1620: with a repetition rate of 1\,GHz.\cite{Cundiff2001} Its repetition
1621: rate and carrier envelope offset frequency were locked to a GPS
1622: stabilized quartz oscillator. The strontium atomic beam is
1623: obtained from the metal heated to 830\,K. Residual atomic beam
1624: divergency is 25\,mrad and the typical atomic density in the
1625: detection region is 10$^8$ cm$^{-3}$. We derive a Doppler-free
1626: signal using a retro-reflected laser beam perpendicular to the
1627: atomic beam. The fluorescence light from the laser excited atoms
1628: is collected on a photomultiplier tube.
1629: 
1630: 
1631: 
1632:  We estimate the indetermination on the reflection angle
1633: of the laser beam to be less than 10\,$\mu$rad. The peak beam
1634: intensity of 60\,$\mu$W/cm$^2$ (to be compared to the saturation
1635: intensity of 3\,$\mu$W/cm$^2$) was chosen to obtain sufficient
1636: signal-to-noise for the RC lock onto the atomic resonance. A
1637: uniform magnetic field of 10\,G (see figure \ref{689Setup})
1638: defines the quantization axis in the interrogation region such
1639: that the light is $\pi$ polarized. The double pass AOM next to the
1640: atomic detection (AOM3) is frequency modulated at 10\,kHz to
1641: derive the locking signal of the cavity onto the atomic line.
1642: 
1643: 
1644: 
1645: \begin{figure}[t]
1646: \centerline{\psfig{file=RedFluorescence,width=9cm}} %\vspace*{8pt}
1647: \caption{Fluorescence spectrum of the strontium
1648: $^1$S$_0$-$^3$P$_1$ line at 689\,nm. The lines of the two bosonic
1649: isotopes $^{86}$Sr and $^{88}$Sr, together with the hyperfine
1650: structure of the fermionic $^{87}$Sr, can be resolved. The
1651: linewidth corresponds to the residual first order Doppler
1652: broadening in the thermal beam. Inset: sub-Doppler resonance of
1653: $^{88}$Sr recorded by saturation spectroscopy using two
1654: counterpropagating laser beams. The amplitude of the dip is 10\,\%
1655: of the Doppler signal. Reprinted figure with permission from G.
1656: Ferrari et al., Phys. Rev. Lett. 91, 243002 (2003). Copyright
1657: (2006) by the American Physical Society.}\label{RedFluor}
1658: \end{figure}
1659: 
1660: 
1661: 
1662: Figure \ref{RedFluor} shows the Doppler broadened resonances of
1663: $^{88}$Sr, $^{86}$Sr, and the hyperfine structure of $^{87}$Sr.
1664: The residual atomic beam divergency produces a residual Doppler
1665: broadening of 60\,MHz FWHM. In the inset, the sub-Doppler signal
1666: for $^{88}$Sr is shown. Two independent measurements of the
1667: sub-Doppler resonance show a FWHM of about 50\,kHz, which is in
1668: agreement with the expected value considering the saturation and
1669: transit time broadening, and the recoil splitting.
1670: 
1671: 
1672: 
1673: 
1674: 
1675: \begin{figure}[t]
1676: \centerline{\psfig{file=FreqPlot,width=8cm}} %\vspace*{8pt}
1677: \caption{Chronological plot of the optical frequency measurements.
1678: The error bars correspond to the standard deviation for each data
1679: set. Reprinted figure with permission from G. Ferrari et al.,
1680: Phys. Rev. Lett. 91, 243002 (2003). Copyright (2006) by the
1681: American Physical Society.}\label{FreqencyMeasure}
1682: \end{figure}
1683: 
1684: 
1685: 
1686: 
1687: 
1688: Figure \ref{FreqencyMeasure} shows the result of the measurement
1689: of the $^{88}$Sr transition frequency taken over a period of
1690: several days. Each data point corresponds to the averaging of the
1691: values resulting from consecutive measurements taken with a 1\,s
1692: integration time over $100 \div 200$\,s. The error bars correspond
1693: to the standard deviation for each data set. We evaluated first-
1694: and second-order Doppler and Zeeman effects, ac Stark shift,
1695: collisional shifts, and mechanical effects of light. The resulting
1696: value for the $^{88}$Sr transition frequency, including the
1697: corrections discussed previously, is 434 829 121 311 (10)\,kHz,
1698: corresponding to a 1$\sigma$ relative uncertainty of $2.3 \times
1699: 10^{-11}$.
1700: 
1701: 
1702: 
1703: 
1704: 
1705: 
1706: \subsection{$^{86}$Sr - $^{88}$Sr isotopic shift measurement} \label{IsotopeShift}
1707: 
1708: 
1709: 
1710: With a minor change in the apparatus, we locked simultaneously the
1711: frequency of two laser beams to the sub-Doppler signals of
1712: $^{86}$Sr and $^{88}$Sr. This system allowed us to measure the
1713: isotopic shift of the $^1$S$_0$-$^3$P$_1$ transition by counting
1714: the beat note between the two interrogating beams. For this
1715: purpose, the reference cavity is locked to the $^{88}$Sr resonance
1716: as described previously and the light for $^{86}$Sr is derived
1717: from the same laser beam and brought to resonance through AOMs.
1718: The two beams are overlapped in a single mode optical fiber and
1719: sent to the interrogation region. By frequency modulating the
1720: beams at different rates and using phase sensitive detection we
1721: get the lock signal for both the isotopes from the same
1722: photomultiplier. In this isotope-shift measurement most of the
1723: noise sources are basically common mode and rejected. The measured
1724: $^{88}$Sr-$^{86}$Sr isotope shift for the $^1$S$_0$-$^3$P$_1$
1725: transition is 163 817.4 (0.2)\,kHz. This value represented an
1726: improvement in accuracy of more than three orders of magnitude
1727: with respect to previously available data.\cite{Buchinger1985} The
1728: $^{86}$Sr optical frequency then amounts to 434 828 957 494
1729: (10)\,kHz.
1730: 
1731: 
1732: 
1733: 
1734: 
1735: \section{Conclusions}
1736: \label{Conclusion}
1737: 
1738: The strontium atom is an attractive candidate both for physical
1739: studies and for applications. We have shown how the properties of
1740: atomic strontium  are suited for laser cooling and trapping, for
1741: the study of ultracold atomic physics, and for applications to
1742: optical frequency metrology or to micrometric force sensors.
1743: 
1744: Future work on ultracold strontium offers intriguing perspectives,
1745: including the possible realization of a nearly non-interacting
1746: Bose-Einstein condensate, the realization of an optical clock with
1747: ultimate stability, or direct experimental tests of theories
1748: beyond the standard model.
1749: 
1750: 
1751: 
1752: 
1753: 
1754: \appendix
1755: 
1756: 
1757: 
1758: 
1759: 
1760: \section{Experimental setup}
1761: \label{ExperimentalSetup}
1762: 
1763: 
1764: 
1765: The typical experimental setup for Sr laser cooling and trapping
1766: basically includes a vacuum system, a blue laser source for the
1767: atom collection and first cooling stage, a red laser for second
1768: stage cooling and precision spectroscopy, and an infrared laser
1769: source for the optical dipole trap. In the following we illustrate
1770: these main parts as they are realized in our laboratory.
1771: 
1772: 
1773: 
1774: 
1775: 
1776: 
1777: 
1778: \subsection{Vacuum system} \label{VacuumSystem}
1779: 
1780: 
1781: 
1782: The apparatus consists in three major parts: the oven, the Zeeman
1783: slower, and the MOT chamber. The oven generates an atomic
1784: strontium vapor by sublimation from metallic strontium kept at
1785: $\sim 800$ Kelvin. The vapor is collimated into an atomic beam
1786: passing through a nozzle filled with about 200 stainless steel
1787: capillaries 8\,mm long which insure a ballistic divergence of the
1788: atomic beam of 20\,mRad.\cite{Courtillot2003} Keeping the
1789: capillaries at a higher temperature prevents internal vapor
1790: condensation and clogging.
1791: 
1792: 
1793: 
1794: 
1795: 
1796: \begin{figure}[t]
1797: \centerline{\psfig{file=vacuum,width=12cm}} \vspace*{8pt}
1798: \caption{Vacuum aparatus. O: oven for St sublimation; IP: ion
1799: pumps; TC: cell for transverse cooling; ZS: differential pumping
1800: tube for the Zeeman slower; C: MOT cell; TS: titanium sublimation
1801: pump; W: sapphire window for the Zeeman slower laser
1802: beam.}\label{VacuumPicture}
1803: \end{figure}
1804: 
1805: 
1806: 
1807: 
1808: 
1809: Following the atomic trajectory (see the sketch in figure
1810: \ref{VacuumPicture}), the atoms pass through a region with radial
1811: optical access for 2D transverse cooling,\cite{Shimizu,Rasel1999}
1812: they are decelerated through the Zeeman slower,\cite{Prodan1982}
1813: and finally stopped in the cell that hosts the MOT. The optical
1814: access for for 2D transverse cooling is provided by a CF35 cube
1815: aligned on the atomic beam. On the two free direction two pairs of
1816: windows, AR coated for 461 and 689\,nm, are sealed with modified
1817: copper gaskets.\cite{WindowSeal} A differential pumping stage
1818: between the transverse cooling region and the Zeeman slower
1819: insures decoupled background pressure between the oven and MOT
1820: region. The oven is pumped by a 20\,l/s ion pump while the MOT is
1821: pumped by a 20\,l/s ion pump and a titanium sublimation pump. With
1822: this setup, in operation condition, we achieve a pressure of
1823: $10^{-7}$ Torr in the transverse cooling region, and a pressure of
1824: $10^{-9}$ Torr on the MOT cell. A BK7 window on the atomic beam
1825: axis provides access for the Zeeman slower decelerating beam. To
1826: prevent chemical reaction of strontium and darkening, the
1827: anti-reflection is only present on the outer side of the window ,
1828: and the window is heated to $\sim 450$\,Kelvin. Paying attention
1829: to block the Sr beam when unnecessary, the window presents a dim
1830: shadow after two years of operation. In the future we plan to
1831: exchange the window with a sapphire one, again anti-reflection
1832: coated on the outer side, from which deposited Sr can be easily
1833: removed.
1834: 
1835: 
1836: 
1837: 
1838: 
1839: \begin{figure}[t]
1840: \centerline{\psfig{file=BlueSetup.ps,width=8cm}} \vspace*{8pt} \caption{Scheme of the laser at 461
1841: nm. A distributed feedback laser (DFB) is amplified in a semiconductor tapered amplifier (TA), and
1842: frequency doubled on a periodically poled KTP crystal. The non-linear conversion is improved by
1843: placing the KTP crystal into an optical resonator. Optical isolators (OI) between the DFB and TA,
1844: and between the TA and the frequency doubling stage. Solid lines represent the optical path,
1845: dashed lines represent electrical connections. BSO: beam shaping optics.}\label{BlueSetup}
1846: \end{figure}
1847: 
1848: 
1849: 
1850: 
1851: 
1852: \subsection{Blue laser sources} \label{BlueLaser}
1853: 
1854: 
1855: 
1856: The light at 461\,nm (see figure \ref{LevelScheme}) used for the
1857: atomic beam deceleration and capture in the MOT is produced by
1858: second-harmonic generation (SHG) of a 922\,nm semiconductor laser
1859: (see figure \ref{BlueSetup}). We generate the 922 nm radiation
1860: with a master oscillator-parametric amplifier system (MOPA). An
1861: anti-reflection coated laser diode mounted in an extended cavity
1862: in Littrow configuration (ECDL)\cite{Littrow1991} delivers 50\,mW
1863: at 922 nm, that is amplified to 1.2 W in a tapered amplifier (TA).
1864: Optical isolators are placed between the ECDL and the TA, and
1865: between the TA and the frequency doubler cavity, in order to
1866: prevent optical feedback into the master laser, and optical
1867: damages on the amplifier. The frequency doubler is composed of a
1868: 20\,mm long periodically-poled KTP crystal, placed in an optical
1869: build up cavity. The crystal facets are anti-reflection coated
1870: both at 922 and 461\,nm (R\,$<$\,0.2\,\%) and the poling period is
1871: chosen to fulfill quasi-phase matching of our wavelength at room
1872: temperature. The resonator has an input coupling mirror with
1873: 11\,\% transmission and it is held in resonance with the input
1874: light feeding the error signal from a H\"{a}nsch-Couillaud
1875: detector\cite{Couillaud81} to a PZT controlled folding mirror.
1876: Under optimal conditions we obtain 300\,mW in the blue and
1877: routinely we work with 200\,mW. We frequency stabilize this blue
1878: source to the $^1$S$_0$-$^1$P$_1$ line of $^{88}$Sr by means of
1879: conventional saturated spectroscopy in a strontium heatpipe. The
1880: servo loop acts on the piezo of the ECDL.
1881: 
1882: 
1883: The light used for atomic manipulation and detection is brought to
1884: the vacuum system through single-mode polarization-maintaining
1885: optical fibers to improve the long-term beam pointing stability.
1886: 
1887: 
1888: 
1889: 
1890: 
1891: 
1892: As we discussed in section \ref{OpticalPumping}, we employ a
1893: turquoise laser at 497\,nm to increase the MOT lifetime and number
1894: of atoms loaded. As in the case of 461\,nm source, there are no
1895: laser diodes available at 497\,nm and the simplest method to
1896: produce this light is frequency doubling an IR laser at 994\,nm.
1897: In this application 1\,mW of light is sufficient to saturate the
1898: process, then we do not need any amplification of the IR light
1899: before frequency doubling. The source at 497\,nm differs from that
1900: at 461\,nm in few parts. The master laser is an anti-reflection
1901: coated diode in ECDL Littrow configuration. After the beam shaping
1902: optics, the IR light is coupled into a bow-tie cavity that
1903: contains a 17\,mm long, b-cut potassium niobate crystal which is
1904: kept at 328\,K in order to satisfy non-critical phase matching for
1905: SHG. The crystal facets are AR coated both for the fundamental and
1906: the blue light and, like the 461\,nm source, the cavity is kept
1907: resonant to the 994\,nm laser with a H\"{a}nsch-Couillaud locking.
1908: Since this laser operates among Sr excited states it is not
1909: possible to lock the laser to the atomic line on a simple
1910: heatpipe. Possible frequency stabilization methods include locking
1911: to a reference cavity or spectroscopy on an atomic sample with a
1912: suitable fraction of excited atoms, such as a hollow cathode lamp,
1913: or an heatpipe with gas discharge, or simply the MOT. We chose the
1914: latter system, in spite of the fact that the blue MOT fluorescence
1915: signal is not continuously available during our measurement
1916: cycles. In fact the short-term laser frequency stability is
1917: sufficient to keep it in resonance with the $^3$P$_2$-$^3$D$_2$
1918: transition for some tens of seconds. Thus we leave the laser free
1919: running, and we manually adjust its frequency at the beginning of
1920: the measurement cycle by acting on the piezo of the ECDL to
1921: maximize the MOT fluorescence.
1922: 
1923: 
1924: 
1925: 
1926: 
1927: 
1928: \subsection{Red laser source} \label{RedLaser}
1929: 
1930: 
1931: 
1932: 
1933: 
1934: The 689\,nm source is composed of a laser diode frequency-locked
1935: to an optical cavity whose modes are stabilized to keep the laser
1936: on resonance with the atomic line. A scheme of the experimental
1937: setup is given in figure \ref{689Setup}. An extended cavity
1938: laser-diode mounted in the Littrow configuration delivers 15\,mW
1939: at 689 nm. Optical feedback into the ECDL is prevented by a 40\,dB
1940: optical isolator and a single pass acusto-optic modulator in
1941: cascade. The laser linewidth is reduced by locking the laser to an
1942: optical reference cavity (RC) with standard Pound-Drever-Hall
1943: technique;\cite{Pound-Drever-Hall} the phase modulation is
1944: produced by an electro-optic modulator (EOM) driven at 11\,MHz and
1945: leaves 85\,\% of the power in the carrier. The reference cavity
1946: has a free spectral range (FSR) of 1.5\,GHz and a finesse of $\sim
1947: 7000$. On one side of the quartz spacer we glued a concave mirror
1948: (R$=50$\,cm) while on the other side a PZT is glued between the
1949: spacer and a flat mirror in order to steer the modes of the cavity
1950: by more than one FSR.
1951: 
1952: 
1953: 
1954: 
1955: 
1956: \begin{figure}[t]
1957: \centerline{\psfig{file=689Setup.eps,width=10cm}} \vspace*{8pt}
1958: \caption{Scheme of the 689\,nm laser. Experimental setup used for
1959: the frequency measurement on the Sr intercombination line. Optical
1960: isolators (OI) and acusto-optic modulators (AOM) eliminate
1961: feedback among the master laser (ECDL), the slave laser, the
1962: electro-optic modulator (EOM) and the reference cavity (RC).  QWP:
1963: quarter wave-plate. CO: collimation optics. PMF: polarization
1964: maintaining fiber. PMT: photo multiplier tube. The same apparatus
1965: is used for Sr second stage cooling and trapping, but the thermal
1966: beam is replaced by an heat-pipe.}\label{689Setup}
1967: \end{figure}
1968: 
1969: 
1970: 
1971: 
1972: 
1973: The locking loop includes a low frequency channel acting on the
1974: PZT of the ECDL (1\,kHz bandwidth), and a high frequency channel
1975: acting on the laser-diode current supply ($\sim 3$\,MHz
1976: bandwidth). Under lock condition more than 55\,\% of the incident
1977: light is transmitted through the cavity. From the frequency noise
1978: spectrum obtained by comparison with an independent cavity we can
1979: infer a laser linewidth lower than 20\,Hz, and more than 90\,\% of
1980: the optical power in the carrier.\cite{Ferrari2004,Poli2006} The
1981: RC is acoustically isolated, though we do not keep it under
1982: vacuum.\cite{UltraStableCavity} The optical table is actively
1983: isolated from seismic noise with pneumatic legs. The frequency
1984: drifts of the cavity are compensated by the servo to the atomic
1985: signal which acts on the PZT of the high finesse cavity with a
1986: 20\,Hz bandwidth. In the frequency measurement experiment
1987: described in section \ref{FreqencyMeasure} the Doppler free
1988: saturated fluorescence on a thermal strontium beam provides the
1989: signal for cavity stabilization on the atomic line. The thermal
1990: beam source has a similar design as described in
1991: \ref{VacuumSystem} and it is pumped by a 20 l/s ion pump. However,
1992: when using the stable 689\,nm laser for second stage cooling and
1993: trapping, as described in section \ref{RedMOT}, we employ an
1994: heat-pipe kept at $\sim 750$\,K instead of the thermal beam, thus
1995: obtaining a larger signal and a more robust lock.
1996: 
1997: 
1998: 
1999: 
2000: 
2001: 
2002: \subsection{Infrared laser sources for dipole trapping} \label{IRlaser}
2003: 
2004: 
2005: 
2006: After the production of an ultracold sample in double stage
2007: magneto-optical trapping, we transfer the atoms in an optical
2008: dipole trap made of two infrared laser beams crossing each other
2009: near the waist. The two beams, respectively aligned along the
2010: horizontal and vertical direction, are produced with two
2011: independent TA injected with light coming from the same infrared
2012: source used for producing the blue light (see figure
2013: \ref{FORTsources}), that is close enough to the ``magic wavelength'' for the $^1$S$_0$-$^3$P$_{1}$
2014: transition.
2015: Typically 300\,mW of
2016: light coming from that source are coupled into a fiber and sent to the two amplifiers.
2017: The injection of the two TAs it is regulated by two AOMs which are
2018: used to shift the frequency of the two beams (avoiding
2019: interference at the center of the dipole trap) and to apply a fast
2020: control to the output TA optical power. For mode cleaning and
2021: delivering the output beams from the TAs to the atoms, we use two
2022: independent single mode fibers. Typically we obtain about 650\,mW
2023: and 440\,mW at the fiber output for the horizontal and vertical
2024: beams, respectively. We finally focus the beams at the MOT center,
2025: with 1/e$^2$ radii of 15 $\mu$m and 35 $\mu$m respectively.
2026: 
2027: 
2028: 
2029: \begin{figure}[t]
2030: \centerline{\psfig{file=laserFORT,width=7cm}} %\vspace*{8pt}
2031: \caption{The 922\,nm laser source for the optical dipole trap.
2032: }\label{FORTsources}
2033: \end{figure}
2034: 
2035: 
2036: 
2037: 
2038: 
2039: 
2040: 
2041: 
2042: 
2043: 
2044: \begin{thebibliography}{0}
2045: 
2046: 
2047: 
2048: 
2049: 
2050: 
2051: \bibitem{KatoriMOT1999}
2052: H. Katori, T. Ido, Y. Isoya and M. Kuwata-Gonokami, {\it Phys. Rev.
2053: Lett.} {\bf 82}, 1116 (1999).
2054: 
2055: 
2056: 
2057: 
2058: \bibitem{Ferrari2006}
2059: G. Ferrari, N. Poli, R. E. Drullinger, F. Sorrentino and G. Tino,
2060: {\it Phys. Rev. A.} {\bf 73}, 023408 (2006).
2061: 
2062: 
2063: 
2064: 
2065: \bibitem{Mukaiyama2003}
2066: T. Mukaiyama, H. Katori, T. Ido, Y. Li and M. Kuwata-Gonokami,
2067: {\it Phys. Rev. Lett.} {\bf 90}, 113002 (2003).
2068: 
2069: 
2070: 
2071: \bibitem{Xu2003}
2072: X. Xu, T. H. Loftus, J. W. Dunn, C. H. Greene, J. L. Hall, A.
2073: Gallagher and J. Ye, {\it Phys. Rev. Lett.} {\bf 90}, 193002
2074: (2003).
2075: 
2076: 
2077: 
2078: \bibitem{Loftus2004}
2079: T. H. Loftus, T. Ido, A. D. Ludlow, M. M. Boyd and J. Ye, {\it
2080: Phys. Rev. Lett.} {\bf 93}, 073003 (2004).
2081: 
2082: 
2083: 
2084: \bibitem{Katori2001}
2085: H. Katori, T. Ido, Y. Isoya and M. Kuwata-Gonokami, in {\it Atomic
2086: Physics 17}, p. 382, edited by E. Arimondo, P. De Natale, and M.
2087: Inguscio (AIP, New York, 2001).
2088: 
2089: 
2090: 
2091: \bibitem{Takamoto2003}
2092: M. Takamoto and H. Katori, {\it Phys. Rev. Lett.} {\bf 91}, 223001
2093: (2003).
2094: 
2095: 
2096: 
2097: \bibitem{Courtillotclock}
2098: I. Courtillot, A. Quessada, R. P. Kovacich, A. Brusch, D. Kolker,
2099: J.-J. Zondy, G. D. Rovera and P. Lemonde, {\it Phys. Rev. A} {\bf
2100: 68}, 030501 (R) (2003).
2101: 
2102: 
2103: 
2104: 
2105: \bibitem{Ferrari2003}
2106: G. Ferrari, P. Cancio, R. Drullinger, G. Giusfredi, N. Poli, M.
2107: Prevedelli, C. Toninelli and G. M. Tino, {\it Phys. Rev. Lett.}
2108: {\bf 91}, 243002 (2003).
2109: 
2110: 
2111: 
2112: \bibitem{Ido2005}
2113: T. Ido, T. H. Loftus, M. M. Boyd, A. D. Ludlow, K. W. Holman, and
2114: J. Ye, {\it Phys. Rev. Lett.} {\bf 94}, 153001 (2005).
2115: 
2116: 
2117: 
2118: 
2119: 
2120: \bibitem{Bidel2002}
2121: Y. Bidel, B. Klappauf, J. C. Bernard, D. Delande, G. Labeyrie, C.
2122: Miniatura, D. Wilkowski and R. Kaiser, {\it Phys. Rev. Lett.} {\bf
2123: 88}, 203902 (2002).
2124: 
2125: 
2126: 
2127: \bibitem{Derevianko2003}
2128: A. Derevianko, S. G. Porsev, S. Kotochigova, E. Tiesinga and P. S.
2129: Julienne, {\it Phys. Rev. Lett.} {\bf 90}, 063002 (2003).
2130: 
2131: 
2132: 
2133: 
2134: \bibitem{Raab1987}
2135: E. L. Raab, M. Prentiss, A. Cable, S. Chu and D. Pritchard, {\it
2136: Phys. Rev. Lett.} {\bf 59}, 2631 (1987).
2137: 
2138: 
2139: 
2140: \bibitem{Kurosu1990}
2141: T. Kurosu and F. Shimizu, {\it Jpn. J. Appl. Phys.} {\bf 29},
2142: L2127 (1990).
2143: 
2144: 
2145: 
2146: 
2147: \bibitem{Chanelière2004}
2148: T. Chaneli\'ere, D. Wilkowski, Y. Bidel, R. Kaiser and C.
2149: Miniatura, {\it Phys. Rev. E.} {\bf 70}, 036602 (2004)
2150: 
2151: 
2152: 
2153: \bibitem{FerrariBloch}
2154: G. Ferrari, N. Poli, F. Sorrentino and G. Tino, {\it Phys. Rev.
2155: Lett.} {\bf 97}, 060402 (2006).
2156: 
2157: 
2158: 
2159: 
2160: \bibitem{KatoriFORT1999}
2161: H. Katori, T. Ido and M. K.-Gonokami, {\it J. Phys. Soc. Jpn.}
2162: {\bf 68}, 2479 (1999).
2163: 
2164: 
2165: 
2166: \bibitem{Ido2000}
2167: T. Ido, Y. Isoya and H. Katori, {\it Phys. Rev. A} {\bf 61},
2168: 061403 (2000).
2169: 
2170: 
2171: 
2172: 
2173: \bibitem{Takamoto2005}
2174: M. Takamoto, F.-L. Hong, R. Higashi and H. Katori, {\it Nature}
2175: {\bf 435}, 321 (2005).
2176: 
2177: 
2178: 
2179: \bibitem{Katori2006}
2180: T. Kishimoto, H. Hachisu, J. Fujiki, K. Nagato, M. Yasuda, and H.
2181: Katori, {\it Phys. Rev. Lett.} {\bf 96}, 123001 (2006).
2182: 
2183: 
2184: 
2185: 
2186: 
2187: \bibitem{Lemonde2005}
2188: P. Lemonde and P. Wolf, {\it Phys. Rev. A} {\bf 72}, 033409
2189: (2005).
2190: 
2191: 
2192: 
2193: \bibitem{Targat2006}
2194: R. Le Targat, X. Baillard, M. Fouch\'e, A. Brusch, O. Tcherbakoff,
2195: G. D. Rovera and P. Lemonde, arXiv:physics/0605200 (2006).
2196: 
2197: 
2198: 
2199: 
2200: \bibitem{Nagel2005}
2201: S. B. Nagel, P. G. Mickelson, A. D. Saenz, Y. N. Martinez, Y. C.
2202: Chen, T. C. Killian, P. Pellegrini and R. C\^ot\'e,  {\it Phys.
2203: Rev. Lett.} {\bf 94}, 083004 (2005).
2204: 
2205: 
2206: 
2207: \bibitem{Mickelson2005}
2208: P. G. Mickelson, Y. N. Martinez, A. D. Saenz, S. B. Nagel, Y. C.
2209: Chen, T. C. Killian, P. Pellegrini and R. C\^ot\'e, {\it Phys.
2210: Rev. Lett.} {\bf 95}, 223002 (2005).
2211: 
2212: 
2213: 
2214: 
2215: \bibitem{Chanelière2005}
2216: T. Chaneli\'ere, J.-L. Meunier, R. Kaiser, C. Miniatura and D.
2217: Wilkowski, arXiv:physics/0412119v1 (2004).
2218: 
2219: 
2220: 
2221: \bibitem{Neuman1995}
2222: J. A. Neuman, P. Wang and A. Gallagher, {\it Rev. Sci. Instrum.}
2223: {\bf 66}, 3021 (1995)
2224: 
2225: 
2226: 
2227: \bibitem{LibroFisicaAtomica}
2228: A. Corney, {\it Atomic and laser spectroscopy}, Oxford, New York:
2229: Clarendon Press (1977).
2230: 
2231: 
2232: 
2233: \bibitem{Grunert}
2234: J. Gr\"unert and A. Hemmerich, {\it Phys. Rev. A} {\bf 65}, 041401
2235: (2002).
2236: 
2237: 
2238: 
2239: \bibitem{Oates1999}
2240: C.W. Oates, F. Bondu and L. Hollberg, {\it Eur. Phys. J. D} {\bf
2241: 7}, 449 (1999).
2242: 
2243: 
2244: 
2245: \bibitem{Prodan1982}
2246: J. V. Prodan, W. D. Phillips and H. Metcalf, {\it Phys. Rev.
2247: Lett.} {\bf 49}, 1149 (1982).
2248: 
2249: 
2250: 
2251: \bibitem{Shimizu}
2252: F. Shimizu, K. Shimizu and H. Takuma, {\it Chem. Phys.} {\bf 145},
2253: 327 (1990).
2254: 
2255: 
2256: 
2257: \bibitem{Rasel1999}
2258: E. Rasel, F. Pereira Dos Santos, F. S. Pavone, F. Perales, C. S.
2259: Unnikrishnan and M. Leduc, {\it Eur. Phys. J. D} {\bf 7}, 311
2260: (1999).
2261: 
2262: 
2263: 
2264: \bibitem{Vogel1999}
2265: T. P. Dinnen, K. R. Vogel, E. Arimondo, J. L. Hall and A.
2266: Gallagher, {\it Phys. Rev. A} {\bf 59}, 1216 (1999).
2267: 
2268: 
2269: 
2270: \bibitem{KatoriFermion}
2271: T. Mukaiyama, H. Katori, T. Ido, Y. Li and M. Kuwata-Gonokami,
2272: {\it Phys. Rev. Lett.} {\bf 90}, 113002 (2003).
2273: 
2274: 
2275: 
2276: \bibitem{Binnewies2001}
2277: T. Binnewies, G. Wilpers, U. Sterr, F. Rihele, J. Helmke, T.E.
2278: Mehlst\"aubler, E.M. Rasel and W. Ertmer, {\it Phys. Rev. Lett.}
2279: {\bf 87}, 123002 (2001).
2280: 
2281: 
2282: 
2283: \bibitem{Santos1995}
2284: M. S. Santos, P. Nussenzveig, L. G. Marcassa, K. Helmerson, J.
2285: Flemming, S. C. Zilio and V. S. Bagnato, {\it Phys. Rev. A} {\bf
2286: 52}, R4340 (1995).
2287: 
2288: 
2289: 
2290: \bibitem{Modugno2001}
2291: G. Modugno, G. Ferrari, G. Roati, R. J. Brecha, A. Simoni, M.
2292: Inguscio, {\it Science} {\bf 294}, 1320 (2001).
2293: 
2294: 
2295: \bibitem{Murdrich2002}
2296: M. Murdrich, S. Kraft, K. Singer, R. Grimm, A. Mosk and M.
2297: Weidemüller, {\it Phys. Rev. Lett.} {\bf 88}, 253001 (2002).
2298: 
2299: 
2300: 
2301: 
2302: \bibitem{Mewesl999}
2303: M.-O. Mewes, G. Ferrari, F. Schreck, A. Sinatra and C. Salomon,
2304: {\it Phys. Rev. A} {\bf 61}, 011403(R) (1999).
2305: 
2306: 
2307: \bibitem{Loftus2001}
2308: T. Loftus, J. R. Bochinski and T. W. Mossberg, {\it Phys. Rev. A}
2309: {\bf 63}, 053401 (2001).
2310: 
2311: 
2312: 
2313: \bibitem{Honda2002}
2314: K. Honda, Y. Takasu, T. Kuwamoto, M. Kumakura, Y. Takahashi and T.
2315: Yabuzaki, {\it Phys. Rev. A} {\bf 66}, 021401(R) (2002).
2316: 
2317: 
2318: 
2319: \bibitem{Stas2004}
2320: R. J. W. Stas, J. M. McNamara, W. Hogervorst and W. Vassen, {\it
2321: Phys. Rev. Lett.} {\bf 93}, 053001 (2004)
2322: 
2323: 
2324: 
2325: \bibitem{Schreck2001}
2326: F. Schreck, G. Ferrari, K. L. Corwin, J. Cubizolles, L.
2327: Khaykovich, M.-O. Mewes and C. Salomon, {\it Phys. Rev. A} {\bf
2328: 64}, 011402 (2001).
2329: 
2330: 
2331: 
2332: \bibitem{Khaykovich2001}
2333: F. Schreck, L. Khaykovich, K. L. Corwin, G. Ferrari, T. Bourdel,
2334: J. Cubizolles and C. Salomon, {\it Phys. Rev. Lett.} {\bf 87},
2335: 080403 (2001).
2336: 
2337: 
2338: 
2339: 
2340: \bibitem{Suptitz1994}
2341: W. Suptitz, G. Wokurka, F. Strauch, P. Kohns and W. Ertmer, {\it
2342: Opt. Lett.} {\bf 19}, 1571 (1994).
2343: 
2344: 
2345: 
2346: \bibitem{Stuhler2001}
2347: J. Stuhler, P. O. Schmidt, S. Hensler, J. Werner, J. Mlynek, and
2348: T. Pfau, {\it Phys. Rev. A} {\bf 64}, 031405(R) (2001).
2349: 
2350: 
2351: 
2352: \bibitem{Nagel2003}
2353: S. B. Nagel, C. E. Simien, S. Laha, P. Gupta, V. S. Ashoka and T.
2354: C. Killian, {\it Phys. Rev. A} {\bf 67}, 011401(R) (2003).
2355: 
2356: 
2357: 
2358: \bibitem{Poli2005}
2359: N. Poli, R. E. Drullinger, G. Ferrari, J. L\'eonard,F. Sorrentino
2360: and G. Tino, {\it Phys. Rev. A} {\bf 71}, 061403(R) (2005).
2361: 
2362: 
2363: 
2364: \bibitem{opticaltrapping}
2365: R. Grimm, M. Weidemüller and Y. B. Ovchinnikov,  {\it Adv. At.
2366: Mol. Opt. Phys.} {\bf 42}, 95 (2000).
2367: 
2368: 
2369: 
2370: \bibitem{Dicke}
2371: R. H. Dicke, {\it  Phys. Rev.} {\bf 89}, 472 (1953).
2372: 
2373: 
2374: 
2375: \bibitem{Ido2003}
2376: T. Ido and H. Katori, {\it Phys. Rev. Lett.} {\bf 91}, 053001
2377: (2003).
2378: 
2379: 
2380: 
2381: \bibitem{Fedichev1996}
2382: P. O. Fedichev, M. W. Reynolds and G. V. Shlyapnikov {\it Phys.
2383: Rev. Lett.} {\bf 77}, 2921 (1996).
2384: 
2385: 
2386: 
2387: \bibitem{Delannoy2001}
2388: G. Delannoy, S. G. Murdoch, V. Boyer, V. Jossel, P. Bouyer and A.
2389: Aspect, {\it Phys. Rev. A} {\bf 63}, 051602(R) (2001).
2390: 
2391: 
2392: 
2393: \bibitem{Yamashita2003}
2394: M. Yamashita and T. Mukai, {\it Phys. Rev. A} {\bf 68}, 063601
2395: (2003).
2396: 
2397: 
2398: 
2399: \bibitem{Monroe1993}
2400: Tc. R. Monroe, E. A. Cornell, C. A. Sackett, C. J. Myatt and C. E.
2401: Wieman, {\it Phys. Rev. Lett.} {\bf 70}, 414 (1993).
2402: 
2403: 
2404: 
2405: \bibitem{Arndt1997}
2406: M. Arndt, M. Ben Dahan, D. Gu\'ery-Odelin, M. W. Reynolds and J.
2407: Dalibard, {\it Phys. Rev. Lett.} {\bf 79}, 625 (1997).
2408: 
2409: 
2410: 
2411: \bibitem{Hopkins2000}
2412: S. A. Hopkins, S. Webster, J. Arlt, P. Bance, S. Cornish, O.
2413: Marago and C. J. Foot, {\it Phys. Rev. A} {\bf 61}, 032707-1
2414: (2000).
2415: 
2416: 
2417: 
2418: \bibitem{Schmidt2003}
2419: P. O. Schmidt, S. Hensler, J. Werner, A. Griesmaier, A.
2420: {G\"orlitz}, T. Pfau and A. Simoni, {\it Phys. Rev. Lett.} {\bf
2421: 91}, 193201 (2003).
2422: 
2423: 
2424: 
2425: \bibitem{Odelin1999}
2426: D. Gu\'ery-Odelin, F. Zabelli, J. Dalibard and S. Stringari, {\it
2427: Phys. Rev. A} {\bf 60}, 4851 (1999).
2428: 
2429: 
2430: \bibitem{Gensemer2001}
2431: S. D. Gensemer and D. S. Jin, {\it Phys. Rev. Lett.} {\bf 87},
2432: 173201 (2001).
2433: 
2434: 
2435: 
2436: 
2437: \bibitem{Kavoulakis2000}
2438: G. M. Kavoulakis, C. J. Pethick and H. Smith, {\it Phys. Rev. A}
2439: {\bf 61}, 053603 (2000).
2440: 
2441: 
2442: 
2443: \bibitem{yasuda2004}
2444: M. Yasuda, T. Kishimoto, M. Takamoto and H. Katori, {\it Phys.
2445: Rev. A} {\bf 73}, 011403(R) (2006).
2446: 
2447: 
2448: 
2449: \bibitem{Goldwin}
2450: J. Goldwin, S. Inouye, M. L. Olsen, B. Newman, B. D. DePaola and
2451: D. S. Jin, {\it Phys. Rev. A} {\bf 70}, 021601 (2004).
2452: 
2453: 
2454: 
2455: \bibitem{K-Rb}
2456: G. Ferrari, M. Inguscio, W. Jastrzebski, G. Modugno, G. Roati and
2457: A. Simoni, {\it Phys. Rev. Lett.} {\bf 89}, 053202 (2002).
2458: 
2459: 
2460: 
2461: \bibitem{Mosk}
2462: A. Mosk, S. Kraft, M. Murdrich, K. Singer, W. Wohlleben, R. Grimm
2463: and M. {Weidem\"uller}, {\it Appl. Phys. B} {\bf 73}, 791 (2004).
2464: 
2465: 
2466: 
2467: \bibitem{Delannoy}
2468: G. Delannoy, S. G. Murdoch, V. Boyer, V. Josse, P. Bouyer and A.
2469: Aspect, {\it Phys. Rev A} {\bf 63}, 051602(R) (2001).
2470: 
2471: 
2472: 
2473: \bibitem{Burt}
2474: E. A. Burt, R. W. Ghrist, C. J. Myatt, M. J. Holland, E. A.
2475: Cornell and C. E. Wieman, {\it Phys. Rev. Lett.} {\bf 79}, 337
2476: (1997).
2477: 
2478: 
2479: \bibitem{Hess1986}
2480: H. F. Hess, {\it Phys. Rev. B} {\bf 34}, (R)3476 (1986).
2481: 
2482: 
2483: 
2484: \bibitem{Takasu2003}
2485: Y. Takasu, K. Maki, K. Komori, T. Takano, K. Honda, M. Kumakura,
2486: T. Yabuzaki and Y. Takahashi, {\it Phys. Rev. Lett.} {\bf 91},
2487: 040404 (2003).
2488: 
2489: 
2490: 
2491: \bibitem{Weber2002}
2492: T. Weber, J. Herbig, M. Mark, H.-C. Naegerl and R. Grimm, {\it
2493: Science} {\bf 299}, 232 (2002).
2494: 
2495: 
2496: 
2497: 
2498: \bibitem{Weiss2003}
2499: T. Kinoshita, T. Wenger and D. S. Weiss, {\it Phys. Rev. A} {\bf
2500: 71}, 011602(R) (2005).
2501: 
2502: 
2503: 
2504: \bibitem{Peters1999}
2505: A. Peters, K. Y. Chung and S. Chu, {\it Nature} {\bf 400}, 849
2506: (1999).
2507: 
2508: 
2509: 
2510: \bibitem{Gustavson2000}
2511: T. L. Gustavson, A. Landragin and M. A. Kasevich, {\it Class.
2512: Quantum Grav.} {\bf 17}, 2385 (2000).
2513: 
2514: 
2515: 
2516: \bibitem{Wicht2002}
2517: A. Wicht, J. M. Hensley, E. Sarajlic and S. Chu, {\it Physica
2518: Scripta} {\bf 102}, 82 (2002).
2519: 
2520: 
2521: 
2522: \bibitem{Clade2006}
2523: P. Clad\'e,  E. de Mirandes, M. Cadoret, S. Guellati-Kh\'elifa, C.
2524: Schwob, F. Nez, L. Julien and F. Biraben1, {\it Phys. Rev. Lett.}
2525: {\bf 96}, 033001 (2006).
2526: 
2527: 
2528: 
2529: \bibitem{Stuhler2003}
2530: J. Stuhler, M. Fattori, T. Petelski and G. M. Tino, {\it J. Opt.
2531: B: Quantum Semiclass. Opt.} {\bf 5}, S75 (2003).
2532: 
2533: 
2534: 
2535: 
2536: 
2537: \bibitem{Fray2004}
2538: S. Fray, C. A. Diez, T. W. H\"ansch and M. Weitz, {\it Phys. Rev.
2539: Lett.} {\bf 93}, 240404 (2004).
2540: 
2541: 
2542: 
2543: 
2544: 
2545: \bibitem{Antezza2005}
2546: M. Antezza, L. P. Pitaevskii and S. Stringari, {\it Phys. Rev.
2547: Lett.} {\bf 95}, 113202 (2005).
2548: 
2549: 
2550: 
2551: \bibitem{Long2003}
2552: J. C. Long, H. W. Chan, A. B. Churnside, E. A. Gulbis, M. C. M.
2553: Varney and J. C. Price, {\it Nature} {\bf 421} (2003).
2554: 
2555: 
2556: \bibitem{Dimopoulos2003}
2557: S. Dimopoulos and A. A. Geraci, {\it Phys. Rev. D} {\bf 68},
2558: 124021 (2003).
2559: 
2560: 
2561: \bibitem{Samullin2005}
2562: S. J. Samullin, A. A. Geraci, D. M. Weld, J. Chiaverini, S. Holmes
2563: and A. Kapitulnik, {\it Phys. Rev. D} {\bf 72}, 122001 (2005).
2564: 
2565: 
2566: 
2567: \bibitem{Bloch2005}
2568: I. Bloch, {\it Nature Phys.} {\bf 1}, 23 (2005).
2569: 
2570: 
2571: 
2572: \bibitem{Bloch1929}
2573: F. Bloch, {\it Z. Phys.} {\bf 52}, 555 (1929).
2574: 
2575: 
2576: 
2577: \bibitem{Raizen1997}
2578: M. Raizen, C. Salomon  and Q. Niu, {\it Physics Today} {\bf 50}
2579: No. 7, 30 (1997).
2580: 
2581: 
2582: 
2583: \bibitem{Anderson1998}
2584: B. P. Anderson, M. A. Kasevich,{\it Science} {\bf 282}, 1686
2585: (1998).
2586: 
2587: 
2588: 
2589: \bibitem{Roati2004}
2590: G. Roati,  E. de Mirandes, F. Ferlaino, H. Ott, G. Modugno and M.
2591: Inguscio, {\it Phys. Rev. Lett.} {\bf 92}, 230402 (2004).
2592: 
2593: 
2594: 
2595: %\bibitem{Petelski}
2596: %T. Petelski, {\it Ph.D. Thesis, Fienze} (2004).
2597: 
2598: 
2599: 
2600: \bibitem{Udem1999}
2601: Th. Udem, J. Reichert, R. Holzwarth and T.W. Ha¨nsch, {\it Opt.
2602: Lett.} {\bf 24}, 881 (1999).
2603: 
2604: 
2605: 
2606: \bibitem{Diddams2000}
2607: S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J.
2608: K. Ranka, R. S. Windeler, R. Holzwarth, T. Udem and T. W. H\"ansch
2609: {\it Phys. Rev. Lett.} {\bf 84}, 5102 (2000).
2610: 
2611: 
2612: 
2613: \bibitem{Udem2002}
2614: Th. Udem, R. Holzwarth and T.W. H\"ansch, {\it Nature} (London)
2615: {\bf 416}, 233 (2002).
2616: 
2617: 
2618: 
2619: \bibitem{Udem2001}
2620: T. Udem, S. A. Diddams, K. R. Vogel, C. W. Oates, E. A. Curtis, W.
2621: D. Lee, W. M. Itano, R. E. Drullinger, J. C. Bergquist and L.
2622: Hollberg, {\it Phys. Rev. Lett.}  {\bf 86}, 4996 (2001) and
2623: references therein.
2624: 
2625: 
2626: 
2627: \bibitem{Gill2005}
2628: P. Gill, {\it Metrologia} {\bf 42}, S125 (2005).
2629: 
2630: 
2631: 
2632: \bibitem{Katori2003}
2633: H. Katori, M. Takamoto, V. G. Pal'chikov and V. D. Ovsiannikov,
2634: {\it Phys. Rev. Lett.} {\bf 91}, 173005 (2003).
2635: 
2636: 
2637: 
2638: \bibitem{Hall1989}
2639: J. L. Hall, M. Zhu and P. Buch, {\it J. Opt. Soc. Am. B} {\bf 6},
2640: 2194 (1989).
2641: 
2642: 
2643: 
2644: \bibitem{Tino1992}
2645: G. M. Tino, M. Barsanti, M. de Angelis, L. Gianfrani and M.
2646: Inguscio, {\it Appl. Phys. B} {\bf 55}, 397 (1992).
2647: 
2648: 
2649: 
2650: 
2651: \bibitem{Ludlow2005}
2652: A. Ludlow, M. M. Boyd, T. Zelevinsky, S. M Foreman, S. Blatt, M.
2653: Notcutt, T. Ido and J. Ye, arXiv:physics/0508041 (2005).
2654: 
2655: 
2656: 
2657: \bibitem{Hoyt2005}
2658: C. W. Hoyt, Z. W. Barber, C. W. Oates, T. M. Fortier, S. A.
2659: Diddams and L. Hollberg, arXiv:physics/0503240 (2005).
2660: 
2661: 
2662: 
2663: \bibitem{Santra2005}
2664: R. Santra, E. Arimondo, T. Ido, C. H. Greene and J. Ye, {\it Phys.
2665: Rev. Lett.} {\bf 94}, 173002 (2005).
2666: 
2667: 
2668: 
2669: \bibitem{Hong2005}
2670: T. Hong, C. Cramer, E. Cook, W. Naugourney and E. N. Fortson, {\it
2671: Phys. Rev. Lett.} {\bf 94}, 050801 (2005).
2672: 
2673: 
2674: 
2675: \bibitem{Taichenachev2005}
2676: A. V. Taichenachev, V. I. Yudin, C. W. Oates, C. W. Hoyt, Z. W.
2677: Barber and L. Hollberg, {\it Phys. Rev. Lett.} {\bf 96}, 083001
2678: (2006).
2679: 
2680: 
2681: 
2682: \bibitem{Barber2005}
2683: Z. W. Barber, C. W. Hoyt, C. W. Oates, L. Hollberg, A. V.
2684: Taichenachev and V. I. Yudin, {\it Phys. Rev. Lett.} {\bf 96},
2685: 083002  (2006).
2686: 
2687: 
2688: 
2689: 
2690: 
2691: \bibitem{Cundiff2001}
2692: S. T. Cundiff, J. Ye, and J. L. Hall, {\it Rev. Sci. Instr.} {\bf
2693: 72}, 3749 (2001).
2694: 
2695: 
2696: 
2697: %\bibitem{Minardi1999}
2698: %F. Minardi et al., {\it Phys. Rev. A} {\bf 60}, 4164 (1999); M.
2699: %Artoni, I. Carusotto and F. Minardi, {\it Phys. Rev. A} {\bf 62},
2700: %023402 (2000).
2701: 
2702: 
2703: 
2704: %\bibitem{Hall1976}
2705: %J. L. Hall and C. J. Borde´, {\it Appl. Phys. Lett.} {\bf 29}, 788
2706: %(1976).
2707: 
2708: 
2709: 
2710: %\bibitem{Crane1994}
2711: %J. K. Crane, M. J. Shaw and R.W. Presta, {\it Phys. Rev. A} {\bf
2712: %49}, 1666 (1994).
2713: 
2714: 
2715: 
2716: \bibitem{Buchinger1985}
2717: F. Buchinger, R. Corriveau, E. B. Ramsay, D. Berdichevsky and D.
2718: W. L. Sprung, {\it Phys. Rev. C} {\bf 32}, 2058 (1985).
2719: 
2720: 
2721: 
2722: %\bibitem{Young1999}
2723: %B. C. Young, F. C. Cruz, W. M. Itano and J. C. Bergquist, {\it
2724: % Phys. Rev. Lett.} {\bf 82}, 3799 (1999).
2725: 
2726: 
2727: 
2728: \bibitem{Courtillot2003} I. Courtillot, A. Quessada, R. P.
2729: Kovacich, J. J. Zondy, A. Landragin, A. Clairon and P. Lemonde,
2730: {\it Opt. Lett.} {\bf 28}, 468 (2003).
2731: 
2732: 
2733: 
2734: \bibitem{WindowSeal}
2735: A. Noble and M. Kasevich, {\it Rev. Sci. Instr.} {\bf 65}, 3042
2736: (1994).
2737: 
2738: 
2739: 
2740: \bibitem{Littrow1991}
2741: C. E. Wieman and L. Hollberg,{\it Rev. Sci. Instr.} {\bf 62}, 1
2742: (1991).
2743: 
2744: 
2745: 
2746: \bibitem{Couillaud81}
2747: B. Couillaud and T.W. H\"{a}nsch, {\it Opt. Comm.} {\bf 35}, 441
2748: (1981).
2749: 
2750: 
2751: 
2752: \bibitem{Pound-Drever-Hall}
2753: R. W. P. Drever, J. L. Hall, F. V. Kowalski, J. Hough, G. M. Ford,
2754: A. J. Munley and H. Ward, {\it App. Phys B} {\bf 31}, 97 (1983).
2755: 
2756: 
2757: \bibitem{Ferrari2004}
2758: G. Ferrari, T. Brzozowski, R. E. Drullinger, N. Poli, M.
2759: Prevedelli, C. Toninelli and G. M. Tino, {\it Proc. SPIE } {\bf
2760: 5478}, 5.3 (2004).
2761: 
2762: 
2763: \bibitem{Poli2006}
2764: N. Poli, G. Ferrari, M. Prevedelli, F. Sorrentino, R. E.
2765: Drullinger and G. M. Tino, {\it Spectrochim. Acta Part A} {\bf
2766: 63}, 981 (2006).
2767: 
2768: 
2769: %\bibitem{ClaironStabilization}
2770: %L.  Hilico, D. Touahri, F. Nez and A. Clairon, {\it Rev. Sci.
2771: %Instr.}{\bf 65}, 3628 (1994).
2772: 
2773: \bibitem{UltraStableCavity}
2774: B. C. Young, F. C. Cruz, W. M. Itano and J. C. Bergquist, {\it
2775: Phys. Rev. Lett.} {\bf 82}, 3799 (1999).
2776: 
2777: 
2778: 
2779: 
2780: 
2781: %\bibitem{ProceedingStPetersburg}
2782: %G. Ferrari, T. Brzozowski, R. E. Drullinger, N. Poli, M.
2783: %Prevedelli, C. Toninelli and G. M. Tino, {\it Laser Optics 2003:
2784: %Solid State Lasers and Nonlinear Frequency Conversion}, V. I.
2785: %Ustugov ed., proceedings of SPIE Vol. 5478, p. 210 (SPIE,
2786: %Bellingham, WA).
2787: 
2788: 
2789: 
2790: 
2791: 
2792: 
2793: 
2794: 
2795: 
2796: \end{thebibliography}
2797: 
2798: 
2799: 
2800: \end{document}
2801: