1: \documentclass[pra,twocolumn,superscriptaddress,footinbib,showpacs]{revtex4}
2:
3: \usepackage{graphicx}
4: \usepackage{epsf}
5:
6: %\setlength{\pdfpagewidth}{\paperwidth}
7: %\setlength{\pdfpageheight}{\paperheight}
8:
9: \newcommand{\bq}{\begin{equation}}
10: \newcommand{\eequ}{\end{equation}}
11: \newcommand{\bqa}{\begin{eqnarray}}
12: \newcommand{\eqa}{\end{eqnarray}}
13: \newcommand{\nn}{\nonumber}
14: \newcommand{\ms}[1]{\mbox{\scriptsize #1}}
15: \newcommand{\dg}{^\dagger}
16: \newcommand{\smallfrac}[2]{\mbox{$\frac{#1}{#2}$}}
17: \newcommand{\la}{\langle}
18: \newcommand{\ra}{\rangle}
19: \newcommand{\ket}[1]{\ensuremath{| {#1} \ra}}
20: \newcommand{\bra}[1]{\la {#1} |}
21: \newcommand{\inprod}[2]{\la {#1} | {#2} \ra}
22: \newcommand{\outprod}[2]{| {#1} \ra \lar {#2} |}
23: \newcommand{\ioh}{-\frac{i}{\hbar}}
24: \newcommand{\oh}{-\frac{1}{\hbar^2}}
25: \newcommand{\sch}{Schr\"odinger }
26: \newcommand{\hei}{Heisenberg }
27: \newcommand{\htn}{Hamiltonian }
28: \newcommand{\half}{\smallfrac{1}{2}}
29: \newcommand{\bl}{{\bigl(}}
30: \newcommand{\br}{{\bigr)}}
31:
32: \def\gt{\ensuremath{\tilde g}}
33: \def\kt{\ensuremath{\tilde k}}
34: \def\kapt{\ensuremath{\tilde \kappa}}
35: \def\gamt{\ensuremath{\tilde \gamma}}
36: \def\ut{\ensuremath{\tilde u}}
37: \def\Delt{\ensuremath{\tilde \Delta}}
38: \def\expct#1{\ensuremath{\langle #1\rangle}}
39: \def\bexpct#1{\ensuremath{\left\langle #1\right\rangle}}
40: \def\paren#1{\ensuremath{\left(#1\right)}}
41: \def\sparen#1{\ensuremath{\left[#1\right]}}
42: \def\cparen#1{\ensuremath{\left\{#1\right\}}}
43: \def\ms#1{\mathrm{#1}}
44: \def\nn{\nonumber}
45: \def\Xe{\ensuremath{\expct{X}_\mathrm{e}}}
46: \def\Pe{\ensuremath{\expct{P}_\mathrm{e}}}
47: \def\dXe{\ensuremath{d\expct{X}_\mathrm{e}}}
48: \def\dPe{\ensuremath{d\expct{P}_\mathrm{e}}}
49: \def\Vxe{\ensuremath{{V_X^\mathrm{e}}}}
50: \def\Vpe{\ensuremath{{V_P^\mathrm{e}}}}
51: \def\dVxe{\ensuremath{dV_X^\mathrm{e}}}
52: \def\dVpe{\ensuremath{dV_P^\mathrm{e}}}
53: \def\Ce{\ensuremath{C_\mathrm{e}}}
54: \def\dCe{\ensuremath{dC_\mathrm{e}}}
55: \def\dWe{\ensuremath{dW_\mathrm{e}}}
56: \def\Ae{\ensuremath{A_\mathrm{e}}}
57: \def\dAe{\ensuremath{dA_\mathrm{e}}}
58: \def\Vcl{\ensuremath{V_\mathrm{cl}(\Xe)}}
59: \def\Veff{\ensuremath{V_\mathrm{eff}}}
60: \def\Heff{\ensuremath{H_\mathrm{eff}}}
61: \def\capsubsz{\scriptscriptstyle}
62: \def\omegaL{\ensuremath{\omega_\mathrm{\capsubsz L}}}
63: \def\omegaA{\ensuremath{\omega_\mathrm{\capsubsz A}}}
64: \def\omegaHO{\ensuremath{\omega_\mathrm{\capsubsz HO}}}
65: \def\prty{\ensuremath{\mathcal{P}}}
66: \def\reduceop{\ensuremath{\mathcal{R}}}
67: \def\prtyr{\ensuremath{\mathcal{R^\dagger PR}}}
68: \def\psiR{\ensuremath{\psi_\mathrm{\capsubsz R}}}
69: \def\psiket{\ensuremath{\ket{\psi}}}
70: \def\Vmax{\ensuremath{{V_\mathrm{max}}}}
71: \def\twoktDeltaXe{2\kt\Delta\!\Xe}
72: \def\DtG{\ensuremath{\Delta t_\mathrm{\capsubsz G}}}
73: \def\taud{\ensuremath{\tau_\mathrm{d}}}
74:
75: \def\workingcomment#1{$\ll\clubsuit$#1$\clubsuit \gg$}
76:
77: \begin{document}
78: %\bibliographystyle{revtex}
79: \pacs{02.30.Yy,32.80.Pj,42.50.-p,03.67.-a}
80: \title{Feedback cooling of atomic motion in cavity QED
81: \vbox to 0pt{\vss
82: \hbox to 0pt{\hskip-37pt\rm LA-UR-04-4687\hss}
83: \vskip 25pt}}
84:
85:
86: \author{Daniel A. Steck}
87: \affiliation{Theoretical Division (T-8), MS B285, Los Alamos National
88: Laboratory, Los Alamos, NM 87545}
89: \affiliation{Oregon Center for Optics and Department of Physics, 1274 University
90: of Oregon, Eugene, OR 97403-1274}
91: \author{Kurt Jacobs}
92: \affiliation{Theoretical Division (T-8), MS B285, Los Alamos National
93: Laboratory, Los Alamos, NM 87545}
94: \affiliation{Centre for Quantum Computer Technology,
95: Centre for Quantum Dynamics, School of Science,
96: Griffith University, Nathan 4111, Australia}
97: \affiliation{Quantum Science and Technologies Group, Hearne Institute for
98: Theoretical Physics, Department of Physics and Astronomy, Louisiana State
99: University, 202 Nicholson Hall, Tower Dr., Baton Rouge, LA 70803}
100: \author{Hideo Mabuchi}
101: \affiliation{Norman Bridge Laboratory of Physics 12-33,
102: California Institute of Technology, Pasadena, CA 91125}
103: \author{Salman Habib}
104: \affiliation{Theoretical Division (T-8), MS B285, Los Alamos National
105: Laboratory, Los Alamos, NM 87545}
106: \author{Tanmoy Bhattacharya}
107: \affiliation{Theoretical Division (T-8), MS B285, Los Alamos National
108: Laboratory, Los Alamos, NM 87545}
109:
110: \begin{abstract}
111: We consider the problem of controlling the motion of an atom trapped
112: in an optical cavity using continuous feedback. In order to realize
113: such a scheme experimentally, one must be able to perform state
114: estimation of the atomic motion in real time. While in theory this
115: estimate may be provided by a stochastic master equation describing
116: the full dynamics of the observed system, integrating this equation
117: in real time is impractical. Here we derive an approximate
118: estimation equation for this purpose, and use it
119: as a drive in a feedback algorithm
120: designed to cool the motion of the atom. We examine the
121: effectiveness of such a procedure using full simulations of the
122: cavity QED system, including the quantized motion of the atom in one
123: dimension.
124: \end{abstract}
125: \maketitle
126:
127: \section{Introduction}
128: The theory of quantum measurement has been mired in controversy for
129: the majority of its history. In a large part this is due to the
130: philosophical difficulties, usually referred to as the ``quantum
131: measurement problem,'' with which it is associated~\cite{Wheeler83}. However,
132: in most areas of physics it has been generally possible to ignore
133: quantum measurement theory, because it was largely irrelevant to real
134: measurements that were being made by experimentalists~\cite{Milburn94}. In
135: particular, if needed at all, an ensemble picture in which quantum
136: measurement theory merely predicted average statistics provided a
137: sufficient description. However, experimental technology has now
138: advanced to the point where repeated measurements may be made on a
139: \textit{single} quantum system as it evolves~\cite{Mabuchi99,
140: Muenstermann99, Hood00, Pinkse00}. As a
141: result, not only does quantum measurement theory become directly
142: relevant to experimental physics, but we can consider testing the
143: predictions of this theory regarding the effects of measurement on
144: the single system itself.
145:
146: While the theory of quantum measurement is now widely accepted, some
147: still remain unconvinced that the conditioned state %predicted wave function collapse
148: describes what is really happening to an \textit{individual} system when
149: it is measured. How might we verify these predictions, and bring at
150: least this part of the controversy to an end? One way to do this is
151: to attempt to perform feedback control on an individual quantum
152: system \cite{Belavkin83,Belavkin87}.
153: Such an experiment involves taking into account the changes in the
154: system due to the measurement results, and using this knowledge to
155: alter inputs to the system in order to control its dynamics.
156: This would therefore provide a direct verification of quantum measurement
157: theory: if
158: the feedback algorithm works, then the theory is correctly predicting
159: the behavior of the system; conversely, if the theory is not giving
160: accurate predictions, the feedback algorithm will be ineffective.
161:
162: Cavity quantum electrodynamics (CQED) is one area in
163: which the motion of a single quantum system, an
164: individual atom in an optical cavity, can be monitored experimentally in real
165: time~\cite{Mabuchi99, Muenstermann99, Hood00, Pinkse00, Sauer04}.
166: Not only does the light emitted from the
167: optical cavity provide a means to track the atomic motion, but the
168: laser driving the cavity provides a means to apply forces to the
169: atom. This system is therefore an excellent candidate for implementing
170: real-time quantum feedback control. It is our purpose here to examine the
171: implementation of such a process in this system, and to show that by
172: using approximate estimation techniques, which will be essential in
173: such experiments for the foreseeable future,
174: a measurable degree of control over the atomic
175: motion can be realized. In particular, we will simulate a control
176: algorithm designed to cool the atomic motion. This might be referred
177: to as ``active'' cooling (control)
178: \cite{Belavkin83, Raizen98, Belavkin99, Doherty99, Doherty00, Armen02, Fischer02, Vitali03, Hopkins03, Geremia03, Geremia04, vanHandel04, Geremia04b, Geremia05, Smith04, Smith02, James04, Reiner04, Hanssen02, Morrow02, Vuletic04, Steck04, Rabl05, Steixner05, Wiseman05},
179: as opposed to passive cooling schemes which
180: have also been the focus of recent theoretical
181: \cite{Horak97, Alge97, Gangl00, Gangl00b, Vuletic00, Domokos01, Gangl01, Horak01, vanEnk01}
182: and experimental \cite{Maunz04} work.
183:
184: In general, feedback control of the atomic motional
185: state requires the real-time estimation of the quantum state from the
186: continuous measurement, thus providing the information needed to decide
187: how the system should be perturbed to bring it to the target state.
188: %To perform feedback control the state of the atomic motion must be
189: %estimated from the continuous measurement in real time.
190: The theoretical tools required for obtaining this continuous estimate are
191: those of continuous quantum measurement~\cite{Belavkin87, Belavkin88,
192: Belavkin89, Chruseinski92, Srinivas81, Barchielli83, Barchielli85, Diosi86, Diosi88, Gisin84, Caves87, Barchielli93, Carmichael93, Wiseman93}.
193: %The observer's ``correct''
194: An estimate of the state of the quantum system (in our case the motional
195: state of the atom), conditioned continuously upon the results of
196: measurement, is given by a stochastic master equation (SME).
197: The concept of an SME was first developed by Belavkin in~\cite{Belavkin87}
198: (as the quantum equivalent of the classical Kushner-Stratonovich equation),
199: and a derivation of an SME in the language of quantum optics may
200: be found in~\cite{Wiseman93}.
201: In principle the observer may calculate this state estimate by
202: integrating the relevant SME using the measurement results, and from
203: this determine the values of the inputs to the system to effect
204: control.
205: However, in practice the complexity of the SME will prevent
206: the observer from calculating the state estimate in real time
207: for systems where the dynamics are sufficiently fast, such as the
208: atomic dynamics we consider here. As a result, a simplified
209: state-estimation procedure is essential.
210:
211: The theory of feedback control for classically observed linear
212: systems is well developed~\cite{Jacobs93, Maybeck82, Whittle96, Zhou95}.
213: This theory considers the optimality and robustness of feedback
214: algorithms for given resources, and in some special cases may be
215: applied directly to the control of quantum systems~\cite{Belavkin87,
216: Doherty99}.
217: Discussions of the application of classical feedback control
218: techniques to quantum systems may be found in Refs.~\cite{Belavkin99, Doherty99, Doherty00}.
219: However, for nonlinear control problems \cite{Doherty00, Wiseman02,
220: Jacobs03},
221: such as that of the atomic
222: motion we consider, no general results regarding the optimality of control
223: algorithms exist as of yet.
224: Here we will be concerned
225: with presenting simple control algorithms for cooling the atoms,
226: and demonstrating
227: that along with an approximate estimation algorithm, a substantial
228: intervention into the dynamics of the atom using real-time feedback
229: control should be achievable experimentally in the near future.
230: In doing so, we build upon previous results
231: \cite{Steck04} by providing a detailed analysis of the
232: cooling performance in this system, including its robustness
233: to parameter variations and real-world limitations on computational
234: speed and detection efficiency. Furthermore, we are able to
235: anticipate several difficulties in an experimental implementation
236: and suggest solutions to overcome them.
237:
238: Owing to the computationally expensive nature of simulating an atom
239: in an optical cavity, including the atomic motion in one dimension,
240: supercomputers are invaluable for the task of evaluating
241: the performance of the control algorithm over many realizations. This
242: is the first time that full quantum simulations of such a CQED
243: system, including the quantized, one-dimensional motion of the atom,
244: have been performed. We are thus able to confirm the results
245: of various approximations which have been used previously to provide
246: a simplified analysis of this system~\cite{Holland91, Storey92, Quadt95,
247: Doherty98, Doherty99}.
248:
249:
250: \section{Description of the system}\label{thesys}
251: The system we consider here is that of a two-level atom in a driven,
252: single-mode optical cavity, where the output from the cavity is
253: monitored using homodyne detection~\cite{Carmichael93, Wiseman93}. Due to the
254: continuous observation, the evolution of the system is described by a
255: stochastic master equation (SME) for the density operator, which
256: can be written in It\^o form as~\cite{noteSME}
257: \begin{equation}
258: \setlength{\arraycolsep}{0ex}
259: \renewcommand{\arraystretch}{1.8}
260: \begin{array}{rcl}
261: d\rho &{}={}& \displaystyle -i[(H/\hbar),\rho]dt
262: + \gamma dt \int_{-\hbar k}^{\hbar k} N(u)
263: \mathcal{D}[\sigma e^{-iux/\hbar}]\rho \, du \\
264: && \displaystyle + \kappa \mathcal{D}[a]\rho\, dt
265: + \sqrt{\eta\kappa} \mathcal{H}[a]\rho\, dW ,
266: \end{array}
267: \label{SME}
268: \end{equation}
269: where
270: the Hamiltonian is~\cite{Doherty98}
271: \begin{equation}
272: \frac{H}{\hbar} = \frac{p^2}{2m\hbar} + g\cos(kx) (\sigma^\dagger
273: a e^{i\Delta t} + \mathrm{H.c.}) - E (a+a^\dagger) ,
274: \end{equation}
275: and we have defined the superoperators $\mathcal{D}[c]$ and $\mathcal{H}[c]$
276: by
277: \begin{equation}
278: \setlength{\arraycolsep}{0ex}
279: \renewcommand{\arraystretch}{1.8}
280: \begin{array}{rcl}
281: \mathcal{D}[c]\rho &{}:={}& c\rho c^\dagger - \frac{1}{2}(c^\dagger c \rho
282: + \rho c^\dagger c)\\
283: \mathcal{H}[c]\rho &{}:={}& c\rho +\rho c^\dagger - \expct{c+c^\dagger}\rho
284: \end{array}
285: \end{equation}
286: for an arbitrary operator $c$.
287: In addition, the measured homodyne signal, given by
288: \begin{equation}
289: dr(t) = \eta\kappa \langle a + a^\dagger \rangle dt
290: + \sqrt{\eta{\kappa}} \, dW,
291: \label{mr}
292: \end{equation}
293: continuously provides information about the optical phase shift due to the
294: atom-cavity system and hence, as we will see below, information about
295: the atomic position.
296: In the above equations, $x$ is an
297: operator representing the atomic position along the length of the
298: cavity, $p$ is the conjugate momentum operator, $a$ is the
299: annihilation operator for the cavity mode, and $\sigma$ is the
300: lowering operator for the atomic internal states. The energy loss rate of
301: the cavity (or equivalently, the full-width-at-half-maximum cavity transmission)
302: is $\kappa$, and $E$ gives the strength of the driving
303: laser, and is related to the laser power $P$ by $E = \sqrt{\kappa
304: P/(\hbar\omegaL)}$, where $\omegaL$ is the angular frequency of the
305: laser light. We take the laser frequency to be resonant with the
306: cavity mode. The laser angular wave number is given by
307: $k=\omegaL/c$, the decay rate of the atomic excited state is denoted
308: by $\gamma$, $g$ is the CQED coupling constant between the atomic
309: internal states and the cavity mode, and $\Delta$ is the detuning
310: between the atom and the cavity mode, given by
311: $\Delta=-(\omegaL-\omegaA)$, where $\omegaA$ is the atomic transition frequency.
312: We will assume a positive (red) detuning, corresponding to a
313: trapping optical potential.
314: Also, $N(u)$ is the probability that the atomic momentum recoil along the
315: $x$-direction is $u$ due to a spontaneous emission event, and
316: $dW$ is the differential of a stochastic Wiener process.
317: Note that we
318: have written the stochastic master equation in the interaction
319: picture where the free oscillation of the light and
320: the dipole rotation of the atom are implicit.
321:
322: The density operator $\rho(t)$ obtained as a solution to the SME
323: condenses our knowledge of the initial condition and the
324: measurement record, allowing a prediction of the statistics of any
325: future measurement on the system. As we show numerically in
326: Section~\ref{sec:estimation-cooling-dynamics}, the solution
327: to the equation is, generically, insensitive
328: to the initial conditions except at very early times.
329: We therefore refer to the solution of the SME as the
330: observer's \textit{best estimate} or \textit{state of knowledge} of the quantum system.
331:
332: In order to perform numerical calculations, it is sensible to choose
333: units for position, momentum and time that are relevant for the corresponding
334: scales of the system. When the detuning of the atom from the cavity
335: is large compared to the cavity decay rate and the CQED coupling
336: constant, both the internal atomic states and the cavity mode may be
337: adiabatically eliminated, and in this case the atom sees an effective
338: potential, given by
339: \begin{equation}
340: \Veff(x) = -\hbar \frac{g^2\alpha^2}{\Delta}\cos^2(kx),
341: \end{equation}
342: where $\alpha := 2E/\kappa$. Since this regime is the one in which we will
343: be primarily interested, convenient units for $x$ and $p$ are
344: given respectively by the width in position and momentum of the
345: ground-state wave function in the harmonic approximation to one of the
346: cosine wells. A convenient unit for time is set by the frequency of
347: oscillation in the same harmonic potential. Thus, given the three
348: scales of the problem,
349: \begin{equation}
350: \begin{array}{rcl}
351: \omegaHO &{}:={}& \alpha g k \sqrt{2\hbar/|m\Delta|} \\
352: w_x &{}:={}& \sqrt{\hbar/(m\omegaHO)} \\
353: w_p &{}:={}& \sqrt{\hbar m\omegaHO} ,
354: \end{array}
355: \end{equation}
356: we will use the
357: scaled variables $X := x/w_x$ and $P := p/w_p$, and we will measure time
358: in units of $2\pi/\omegaHO$.
359: The SME for the density matrix may now be written
360: as
361: \begin{equation}
362: \setlength{\arraycolsep}{0ex}
363: \renewcommand{\arraystretch}{1.8}
364: \begin{array}{rcl}
365: d\rho &{}={}& -i[H,\rho]dt + \tilde{\gamma} \,dt
366: \int_{-\tilde{k}}^{\tilde{k}} N(\tilde{u})
367: \mathcal{D}(\sigma e^{-i\tilde{u}X})\rho \, d\tilde{u} \\
368: &&+ \tilde{\kappa} \mathcal{D}[a]\rho\, dt +
369: \sqrt{\eta\tilde{\kappa}} \mathcal{H}[a]\, dW ,
370: \label{scSME}
371: \end{array}
372: \end{equation}
373: where
374: \begin{equation}
375: H = \pi P^2 + \tilde{g} \cos(\tilde{k}X)
376: (\sigma^\dagger a e^{i\tilde{\Delta} t} + \mbox{H.c}) - \tilde{E}
377: (a+a^\dagger)
378: \end{equation}
379: and
380: \begin{equation}
381: d\tilde{r}(t) = \eta \langle a + a^\dagger \rangle dt
382: + \sqrt{\eta} \, dW,
383: \label{mrscaled}
384: \end{equation}
385: with $d\tilde{r}(t) = dr(t)/\sqrt{\kapt}$.
386: The scaled rate constants $\tilde{\kappa}, \tilde{E}, \tilde{\Delta},
387: \tilde{\gamma}$, and $\tilde g = \sqrt{\pi\Delt}/(\alpha\kt)$
388: are obtained from
389: their absolute counterparts by dividing by $\omegaHO/(2\pi)$. The scaled
390: wave number is naturally $\tilde{k} = k w_x$. Note that in these
391: scaled units we have scaled $\hbar$ out of the problem, so that $[X,P]=i$.
392:
393: While we will solve this equation numerically, it is nevertheless
394: useful to examine the solutions obtained by adiabatically eliminating
395: both the atomic internal states and the cavity mode,
396: which is possible when the detuning
397: $\tilde{\Delta}$ is much larger than the coupling constant
398: $\tilde{g}$ so that the excited atomic state is nearly unpopulated during
399: the evolution \cite{Doherty98}. In
400: this case the Hamiltonian part of the evolution is governed by the
401: effective Hamiltonian
402: \begin{equation}
403: \Heff := \pi P^2 - \frac{\tilde{g}^2}{\tilde{\Delta}}
404: \cos^2(\tilde{k}X) a^\dagger a .
405: \end{equation}
406: Thus, the atom moves in a sinusoidal potential, with period
407: $\pi/\tilde{k}$. The average height of the potential wells
408: is essentially proportional
409: to $\langle a^\dagger a \rangle$. It is
410: therefore instructive to consider the cavity-mode dynamics,
411: which determine the magnitude of the effective potential. Under the
412: effective Hamiltonian, the Heisenberg equations of motion for the
413: relevant cavity mode operators are
414: \begin{equation}
415: \setlength{\arraycolsep}{0ex}
416: \renewcommand{\arraystretch}{1.8}
417: \begin{array}{rcl}
418: \partial_t{a} &{}={}& \displaystyle i\tilde{E} - \frac{\tilde{\kappa}}{2} a +
419: i\frac{\tilde{g}^2}{\tilde{\Delta}}\cos^2(\kt X) a + \sqrt{\tilde{\kappa}}
420: a_\mathrm{in}(t) \\
421: \partial_t{(a^\dagger a)} & = & \displaystyle
422: -\tilde{\kappa}a^\dagger a - i \tilde{E}(a
423: - a^\dagger) \\ &&\displaystyle
424: +\sqrt{\tilde{\kappa}}(a^\dagger a_\mathrm{in}(t) +
425: a_\mathrm{in}^\dagger(t) a) ,
426: \end{array}
427: \end{equation}
428: where $a_{\mathrm{in}}(t)$ is the input quantum noise operator coming
429: from the electromagnetic field outside the cavity, and satisfies
430: $[a_{\mathrm{in}}(t),a_{\mathrm{in}}^\dagger(t+\tau)]=\delta(\tau)$. We can
431: now identify a regime in which $\tilde{\kappa}$ is large compared to
432: $\tilde{g}^2/\tilde{\Delta}$. In this case the cavity mode damps
433: quickly to a steady state, and we can approximate the solution by
434: \begin{equation}
435: \setlength{\arraycolsep}{0ex}
436: \renewcommand{\arraystretch}{2.1}
437: \begin{array}{rcl}
438: a &{}\approx{}& \displaystyle
439: i\alpha \left[ 1 - i2\left(\frac{\tilde{g}^2}{\tilde{\kappa}\tilde{\Delta}}\right)
440: \cos^2(\tilde{k}X) \right]^{-1} , \\
441: &{}\approx{}& \displaystyle i\alpha - \alpha \left(
442: \frac{2\tilde{g}^2}{\tilde{\Delta}\tilde{\kappa}}
443: \right) \cos^2(\tilde{k}X) , \\
444: a^\dagger a &{}\approx{}& \displaystyle\alpha \frac{(a - a^\dagger)}{2i} .
445: \end{array}
446: \end{equation}
447: There will be fluctuations about these values due to the noise
448: with a spectrum with a width of the order of $\tilde{\kappa}$. When
449: $\kappa$ (or $\tilde{\kappa}$) is sufficiently large,
450: then $\langle a^\dagger a \rangle \approx \alpha^2$, and the height
451: of the potential will therefore be determined by the strength of the
452: driving. As a consequence, one can manipulate the potential that the
453: atom sees by changing the strength $\tilde{E}$ of the driving.
454: The measured homodyne signal is given in the adiabatic approximation by
455: \begin{equation}
456: d\tilde{r}(t) = -
457: \sqrt{8\eta^2\Gamma} \langle
458: \cos^2(\kt X)\rangle dt + \sqrt{\eta} \, dW ,
459: \label{mradiab}
460: \end{equation}
461: where $d\tilde{r}(t):= dr(t)/\sqrt{\kapt}$;
462: the position-information content of
463: this signal is more clear in this form.
464: %These
465: %results suggest, therefore, that one can perform feedback control of
466: %the atomic motion by continuously tracking the atomic dynamics using
467: %homodyne detection on the cavity output light, and using this
468: %information to alter the strength of the driving laser.
469:
470: \section{Cooling the atomic motion using feedback}\label{alg}
471: As mentioned in the introduction, in implementing feedback control,
472: two considerations are paramount. The first is the ability to effectively
473: track the evolution of the system in real time (state estimation).
474: The second is to provide an algorithm which determines how this
475: information about the system should be used to alter the inputs to
476: the system to effect control. We now consider these two questions in
477: turn.
478:
479: \subsection{State Estimation in Real Time}\label{section:gest}
480: The SME given in Eq.~(\ref{SME}) continuously provides the observer's
481: best estimate of the state of the system from the information
482: provided by the measurement record $r(t)$ [Eq.~(\ref{mr})]. Therefore,
483: it is possible, in principle, to use the measurement record to integrate the
484: full SME to obtain a continuous state estimate. However, the
485: resources required to do this in real time are prohibitive in practice. This is
486: because the SME, being a partial stochastic differential equation,
487: contains a large effective number of variables, and is almost
488: impossible to integrate in real time, given that the time scale of
489: the atomic motion is typically a few $\mu$s. As a result,
490: we must obtain an estimation equation which contains only a small
491: number of variables, but is nevertheless a good approximation to the
492: full SME.
493:
494: To obtain a compact system of estimation equations, we first note that initially
495: Gaussian states remain close to Gaussian for a fairly large number of
496: oscillations in a sufficiently strong sinusoidal potential.
497: We thus make a
498: Gaussian approximation for the density matrix, which amounts to ignoring
499: all cumulants higher than second order.
500: In the present system, which effectively measures
501: $\expct{\cos^2(\kt X)}$, we expect this approximation to remain valid as long
502: as the wave packet is much narrower than a single well ($\kt \Delta X\ll 1$),
503: and the wave-packet momentum width
504: is much larger than two photon recoils ($\Delta P \gg 2\kt$).
505: To further simplify the estimation equations, we consider
506: the SME after adiabatic elimination of the cavity
507: and internal atomic degrees of freedom, given by
508: \begin{equation}
509: \setlength{\arraycolsep}{0ex}
510: \renewcommand{\arraystretch}{1.5}
511: \begin{array}{rcl}
512: d\rho &{}={}& \displaystyle -i[\Heff,\rho]\, dt
513: + 2 \Gamma\mathcal{D}[\cos^2(\kt X)]\rho\, dt \\
514: &&\displaystyle -\sqrt{2\eta\Gamma}\mathcal{H}[\cos^2(\kt X)]\rho\, dW,
515: \end{array}
516: \label{adiabaticSME}
517: \end{equation}
518: which gives a three-parameter model ($\kt$, $\Gamma$, and
519: $\eta$).
520: Here, we have defined the effective Hamiltonian
521: \begin{equation}
522: \Heff = \pi P^2 - \Vmax\cos^2(\kt X),
523: \end{equation}
524: where the maximum light shift
525: \begin{equation}
526: \Vmax := \frac{\alpha^2\gt^2}{\Delt} = \frac{\pi}{\kt^2}
527: \end{equation}
528: is not an independent parameter in this static problem, but we will
529: treat it as such when considering feedback via a time-dependent potential.
530: We have also defined the effective measurement strength
531: \begin{equation}
532: \Gamma := \frac{2\alpha^2 \tilde{g}^4}{\tilde{\Delta}^2\tilde{\kappa}}.
533: \label{Gammadef}
534: \end{equation}
535: We further simplify the estimation model by making a Gaussian
536: ansatz for the density operator.
537: The resulting estimation equations contain merely 5
538: variables: the estimated mean position and momentum, \Xe\
539: and \Pe, the estimated
540: variances, \Vxe\ and \Vpe, and the estimated
541: symmetrized covariance $\Ce := \langle XP +
542: PX\rangle_{\mathrm{e}}/2 - \Xe\Pe$.
543: (Due to the information content of the measurement process, the two
544: variances and the covariance are algebraically independent, in
545: contrast to a closed system.)
546: Using the equation of motion for the expectation value of
547: an arbitrary operator $A$,
548: \begin{equation}
549: \setlength{\arraycolsep}{0ex}
550: \renewcommand{\arraystretch}{1.4}
551: \begin{array}{rcl}
552: d\expct{A}&{}={}&\mathrm{Tr}[A\,d\rho]\\ &{}={}&\displaystyle
553: -i\expct{[A,\Heff]}dt
554: + 2 \Gamma \expct{D[\cos^2(\kt X)]A}dt\\ && \displaystyle
555: {}- \sqrt{2\eta\Gamma}\expct{\mathcal{H}[\cos^2(\kt X)]A}dW,
556: \end{array}
557: \end{equation}
558: we can calculate the system of five Gaussian estimation equations,
559: which are
560: \begin{equation}
561: \setlength{\arraycolsep}{0ex}
562: \renewcommand{\arraystretch}{1.4}
563: \begin{array}{rcl}
564: \dXe &{}={}& 2\pi \Pe dt \\
565: && + \sqrt{8\eta\Gamma} \kt \Vxe
566: \exp(-2\kt^2\Vxe) \sin(2\kt\Xe) \dWe
567: \\ %\\
568: \dPe &{}={}& \displaystyle -
569: \Vmax %\left( \frac{\alpha^2 \tilde{g}^2}{\Delt} \right)
570: \kt \exp(-2\kt^2\Vxe) \sin(2\kt\Xe) dt \\
571: & & + \sqrt{8\eta\Gamma} \kt \Ce \exp(-2\kt^2\Vxe)\sin(2\kt\Xe) \dWe
572: \\ %\\
573: \dVxe &{}={}& 4\pi\Ce dt \\
574: & & -8\eta\Gamma \kt^2 \Vxe^2 \exp(-4\kt^2\Vxe) \sin^2(2\kt\Xe) dt\\
575: & & +2 \sqrt{8\eta\Gamma} \kt^2 \Vxe^2\exp(-2\kt^2\Vxe) \cos(2\kt\Xe) \dWe
576: \\ %\\
577: \dVpe &{}={}&\displaystyle
578: -4 \Vmax %\left( \frac{\alpha^2 \gt^2}{\Delt} \right)
579: \kt^2\Ce\exp(-2\kt^2\Vxe)\cos(2\kt\Xe) dt\\
580: && + \Gamma \kt^2[1-\exp(-8\kt^2\Vxe)\cos(4\kt\Xe)] dt\\
581: && - 8\eta\Gamma \kt^2 \Ce^2 \exp(-4\kt^2\Vxe)\sin^2(2\kt\Xe) dt\\
582: && - \sqrt{2\eta\Gamma}\kt^2[1-4
583: (\Ce^2+\kt^2\Vxe)\exp(-2\kt^2\Vxe)]\\
584: &&\hspace{6mm}\times\cos(2\kt\Xe)\dWe
585: \\ %\\
586: \dCe &{}={}& \displaystyle
587: 2\pi\Vpe dt\\
588: &&\displaystyle -2 \Vmax %\left( \frac{\alpha^2 \gt^2}{\Delt} \right)
589: \kt^2\Vxe\exp(-2\kt^2\Vxe)\cos(2\kt\Xe) dt\\
590: && - 8\eta\Gamma \kt^2 \Vxe \Ce \exp(-4\kt^2\Vxe)\sin^2(2\kt\Xe) dt\\
591: && +2\sqrt{8\eta\Gamma}\kt^2\Vxe\Ce\exp(-2\kt^2\Vxe)\cos(2\kt\Xe)\dWe .
592: \end{array}
593: \label{gest}
594: \end{equation}
595: Although we can eliminate \Vmax\ from these estimator equations
596: (since $\Vmax\kt^2=\pi$), we keep the optical potential explicit
597: here because both $\Vmax$ and $\Gamma$ will effectively
598: become time-dependent
599: quantities when we consider feedback control by modulating the
600: amplitude of the optical potential.
601: The reconstructed ``Wiener increment'' \dWe\ arises by making the Gaussian
602: approximation to the photocurrent equation (\ref{mradiab}):
603: \begin{equation}
604: \dWe = \frac{d\tilde{r}(t)}{\sqrt{\eta}}
605: + \sqrt{2\eta\Gamma}[1+\exp(-2\kt^2\Vxe)\cos(2\kt\Xe)]dt .
606: \end{equation}
607: These estimation equations resemble stochastic differential equations,
608: but it is important to recognize that \dWe\ does not have the same statistics
609: as the usual Wiener increment $dW$.
610: In particular, the reconstructed increment is the usual Wiener
611: increment with an extra deterministic component,
612: \begin{equation}
613: \setlength{\arraycolsep}{0ex}
614: \begin{array}{rcl}
615: \dWe &{}={}& dW + \sqrt{2\eta\Gamma}\left[
616: \exp(-2\kt^2\Vxe)\cos(2\kt\Xe)\right. \\&&
617: \hspace{40mm}{}- \displaystyle\left.
618: \langle\cos(2\kt X)\rangle\right] dt,
619: \end{array}
620: \label{dWefromdW}
621: \end{equation}
622: which reflects the difference between the actual and estimated
623: quantum states. This is the route by which the measurement
624: information is incorporated into the estimator evolution.
625: Thus, these equations cannot be treated by the usual high-order
626: numerical methods for stochastic differential equations, which assume
627: that the deterministic and stochastic parts of the differential
628: equation can be explicitly separated.
629:
630: Another important issue to note is that because the estimator only has
631: information about $\langle \cos^2(\tilde{k}X) \rangle$, it cannot
632: tell us in which well the atom is located,
633: and therefore, while feedback will allow us to cool an atom within a
634: well, we cannot know within \textit{which} well the atom is confined.
635: Moreover, the estimator cannot tell us on what side of a well the
636: atom is located.
637: Note that this degeneracy is not a consequence of the Gaussian
638: approximation, but also applies to an observer using the
639: full SME for estimation.
640: As a consequence, the estimator will show us the
641: motion of the atom up to a phase factor of $\pi$: the estimated state
642: will either be in phase with the true motion, or completely out of phase,
643: since these two motions are mirror images of one another. Thus, for
644: the feedback algorithm to be effective for every run, it is important
645: that it works for both cases without change. The algorithms that we
646: describe in the next section have this property.
647:
648: It is worth noting that the insensitivity of the feedback algorithm
649: to the side of the well the atom is on has a potential advantage. In
650: simulating the dynamics of the atom, we find that if the initial
651: energy of the atom is large enough for it to move from one well to
652: another, then as it does so the wave function tends to split across
653: the interwell barrier, with part going back down into the original well,
654: and part going down into an adjacent well. In this case the wave function
655: is no longer Gaussian, but consists of two or more ``humps."
656: However, even though the two humps are on opposite sides
657: of their respective wells, they have approximately the same downward
658: motion. Thus, even though the estimator is Gaussian,
659: the fact that the humps are in different wells is
660: immaterial to the measurement of $\langle\cos^2(\kt X)\rangle$,
661: and the estimator effectively tracks the evolution of
662: an ``equivalent'' quantum state localized within a single well in the
663: sense of generating the same measurement record as the fragmented state.
664: As a result, we can
665: expect the Gaussian estimator to continue to work even though the
666: atom is distributed across more than one well.
667:
668: In order to perform the continuous estimation (whether using
669: the full SME or the above Gaussian approximate estimator) one chooses
670: an initial state for the estimator that reflects the initial ignorance
671: of the true quantum state and integrates
672: the estimator equations using the measurement record as this record
673: is obtained. In our case, when the atom is dropped into the cavity we
674: have very little idea of where it is. It is therefore sensible to
675: choose as the initial state for the estimator a Gaussian broad enough
676: to cover one side of a well, and centered somewhere on that side (it
677: is not important where precisely). We only need to include one side of
678: the well in our initial uncertainty, since as mentioned above, both
679: sides are equivalent from the point of view of the estimator.
680: In practice, though, it is sufficient to pick a state with a much
681: smaller uncertainty product,
682: centered on one side of a potential well. Even if this
683: estimator state
684: does not substantially overlap the actual quantum state, as we will
685: see below, it will
686: converge to the ``true'' state in a time of order $1/\Gamma$ as the
687: measurement information is incorporated.
688:
689:
690: \subsection{An Effective Cooling Algorithm}\label{section:alg}
691:
692: As discussed in Section~\ref{thesys}, it is possible to raise and
693: lower the height of the potential by changing the input laser power.
694: Thinking of the atom as a classical particle, one would expect to be
695: able to cool the atom using the following simple algorithm: When the
696: atom is moving toward the center of the well,
697: it is being accelerated by the potential, and so we can lower the
698: laser power to reduce this acceleration. On the other hand, when the
699: atom is moving away from the center of the well, then the potential
700: decelerates the atom, and in this case we can increase the laser
701: power so as to increase the deceleration.
702: Writing the optical field amplitude in the absence of
703: feedback as $E$, we will denote the value of the field during
704: times when it is ``switched high'' as $(1 + \varepsilon_1) E$ and the
705: value when it is ``switched low'' as $(1 - \varepsilon_2) E$.
706: Thus the potential height will have the respective values
707: $(1 + \varepsilon_1)^2 V$ and $(1 - \varepsilon_2)^2 V$, where $V$ is the
708: unmodulated potential depth.
709: In this way, we repeatedly
710: switch the potential between a high value and a low value, which
711: both slows the atom and moves it in towards center of the well,
712: reducing the total energy and hence cooling the atom. This algorithm
713: also has the nice property that all the information it requires is
714: whether the atom is climbing or descending the well. Therefore,
715: the question of whether the estimated state is in phase or completely
716: out of phase with the true motion is immaterial, since in both cases the relevant
717: information is known. However, it turns out that this algorithm is not
718: very effective at cooling the atomic motion, because while it does
719: cool the centroid, it simultaneously feeds energy into
720: the motion by squeezing the atomic wave function. A detailed discussion
721: of this effect is given in Appendix A. The result is that this simple
722: centroid-only cooling algorithm is not adequate to cool the atom to the
723: ground state. To do so we must develop a more sophisticated algorithm
724: which can reduce the motional energy associated with the variances of
725: the wave function.
726:
727: To proceed, we consider the change in the motional energy due to time
728: dependence of the optical-potential amplitude:
729: \begin{equation}
730: \partial_t\langle {E}_\mathrm{eff}\rangle_\mathrm{fb}
731: = -(\partial_t \Vmax)\langle\cos^2(\kt X)\rangle.
732: \label{coolingenergy}
733: \end{equation}
734: It is clear from this expression that the ``bang-bang''
735: feedback strategy from the simple cooling algorithm, where the
736: potential amplitude is switched cyclically and
737: suddenly between two values,
738: maximizes
739: the energy extraction rate if we consider only algorithms
740: with a cyclic modulation of the potential within that range.
741: In fact, we should switch the potential low when the
742: value of $-\langle\cos^2(\kt X)\rangle$ is maximized (i.e., the
743: modulus is minimized), thus
744: maximizing the energy extracted from the atom. Then we should switch
745: the potential high when $-\langle\cos^2(\kt X)\rangle$
746: is minimized, thus minimizing the energy transferred to the
747: atom when completing the modulation cycle.
748:
749: When we must rely on the Gaussian estimator
750: to determine when to switch the potential, we switch the potential
751: low or high when the estimated quantity
752: \begin{equation}
753: y_\mathrm{est} := -\exp(-2\kt^2 \Vxe)\cos(2\kt \Xe)
754: \label{gestswitch}
755: \end{equation}
756: is maximized or minimized, respectively. Doing so involves an
757: additional technical complication, however, since at any given
758: instant, we must predict whether or not $y_\mathrm{est}$ is
759: an extremal value. To do this we dynamically fit a quadratic
760: curve to a history of these values (typically the last several
761: hundred), and then trigger the feedback transitions
762: on the slope of the fitted curve. This procedure is a convenient
763: method for predicting the times when $y_\mathrm{est}$ is
764: maximized or minimized, and helps to reject the residual noise in
765: the time evolution of this quantity.
766: More formally, given a
767: set of estimates $y_{\mathrm{est}, n}$ corresponding to
768: times $t_n$, we implement the curve fit of the
769: function $a_0+a_1 x + a_2 x^2$ to the last $q$ values of
770: $y_{\mathrm{est}, i}$. We will defer the discussion of the
771: computational efficiency of this algorithm until Section
772: \ref{section:speed}.
773:
774: It may seem strange that this algorithm makes use of past state
775: information, when all the necessary information should in principle
776: be contained in the current estimate of the state. Indeed, it is
777: clear from the
778: quantum Bellman equation \cite{Belavkin99, Belavkin87, Doherty00} that
779: the optimal control algorithm is given by only looking forward
780: in time given the current state.
781: In practice, such an algorithm is difficult to realize,
782: and the present algorithm is likely suboptimal but robust and
783: effective.
784:
785: We also note that since we are triggering in this algorithm
786: on the estimated value of $-\langle\cos^2(\kt X)\rangle$, it
787: seems that it could be much simpler to simply trigger on the
788: detector photocurrent $d\tilde{r}(t)$,
789: which from Eq.~(\ref{mradiab}) we know is
790: a direct measure of this quantity plus noise. In fact,
791: we can use the same curve-fitting procedure on $d\tilde{r}(t)$ to reject
792: the noise on the signal. However, as we will see, this procedure
793: does not work nearly as well as with the Gaussian estimator, because
794: the signal $d\tilde{r}(t)$ is noise-dominated, and the Gaussian estimator
795: acts as a nearly optimal filter for the useful information from
796: this signal, in the same sense as a Kalman filter
797: \cite{Maybeck82, Belavkin79, Doherty99}.
798:
799:
800: \subsection{Cooling Limits}\label{section:limits}
801:
802: At this point, we can work out a simple theory of cooling for the
803: improved cooling algorithm developed
804: in the previous section. When we switch
805: the optical potential low, the
806: change in energy from Eq.~(\ref{coolingenergy}) is given by
807: \begin{equation}
808: \Delta\langle {E}_\mathrm{eff}\rangle_\mathrm{fb}
809: = -4\varepsilon \Vmax\langle\cos^2(\kt X)\rangle,
810: \label{downtrans}
811: \end{equation}
812: where we have taken $\varepsilon_1=\varepsilon_2=\varepsilon$, so that the
813: potential amplitude is switched from $(1+\varepsilon)^2\Vmax$ to
814: $(1-\varepsilon)^2\Vmax$. When we switch high, the expression
815: for the energy change is
816: the same except for an overall minus sign.
817:
818: Now we consider these expectation values for a particle very close to the
819: ground state, and we will first consider the case where
820: all the excess motional energy is associated with the motion of
821: the wave-packet centroid.
822: For the ground state itself, an equal part $E_0/2$
823: of the ground-state energy $E_0$ is associated with each of
824: the kinetic and potential parts of the effective Hamiltonian.
825: If we define $\Delta E := \langle {E}_\mathrm{eff}\rangle - E_0$
826: as the small, additional motional energy above the ground state,
827: then as the displaced wave packet (coherent state)
828: evolves in the nearly harmonic
829: potential,
830: the value $\Vmax\langle[1-\cos^2(\kt X)]\rangle$ of the
831: potential part of the atomic energy oscillates between
832: $E_0/2$ and $E_0/2 + \Delta E$.
833: Then the best cooling that can be achieved for this state corresponds
834: to two up and two down transitions per oscillation period (which is
835: unity in our scaled units). Inserting these extremal values of
836: $-\Vmax\langle\cos^2(\kt X)\rangle$ into Eq.~(\ref{downtrans})
837: and the corresponding expression for upward transitions
838: gives the cooling rate of
839: \begin{equation}
840: \Delta\langle {E}_\mathrm{eff}\rangle_\mathrm{fb}
841: = -8\varepsilon(\langle {E}_\mathrm{eff}\rangle - E_0)
842: \label{coolingrate}
843: \end{equation}
844: per unit time.
845:
846: In practice, the cooling algorithm switches more often than in this
847: estimate due to noise on the signal.
848: However, in such cases we might still expect that Eq.~(\ref{coolingrate}) is
849: a reasonable estimate for the cooling rate.
850: Even though the potential is switched more often, the switches are
851: caused by noise and thus do not occur at optimal moments in time.
852: Thus the amount of cooling per switching cycle is reduced, and we
853: observe that the net cooling effect per unit time is near the optimal
854: rate of Eq.~(\ref{coolingrate}), presumably
855: because the cooling algorithm is not too sensitive
856: to the noise.
857: If we then assume that heating due to spontaneous emission is
858: negligible compared to measurement heating, then the steady-state
859: motional energy is obtained when the cooling rate
860: in Eq.~(\ref{coolingrate}) balances the heating rate which we derive in
861: Appendix~\ref{heating} (Eq.~(\ref{measurementheating})).
862: This gives the steady-state energy
863: \begin{equation}
864: \langle {E}_\mathrm{eff}\rangle_\mathrm{ss}
865: = \frac{E_0}{1-\beta},
866: \label{E0ss}
867: \end{equation}
868: where $\beta := \Gamma\kt^4/2\varepsilon$.
869:
870: The other extreme case we will consider is where the motional energy
871: in excess of the ground-state energy is associated purely with
872: squeezing of the wave packet and not with the centroid.
873: In this case, $\Vmax\langle\cos^2(\kt X)\rangle$
874: oscillates between the values
875: $(\langle E_\mathrm{eff}\rangle
876: \pm\sqrt{\langle E_\mathrm{eff}\rangle^2-E_0^2})/2$, giving the
877: cooling rate of
878: \begin{equation}
879: \Delta\langle {E}_\mathrm{eff}\rangle_\mathrm{fb}
880: = -8\varepsilon\sqrt{E_\mathrm{eff}^2-E_0^2}
881: \label{coolingratevar}
882: \end{equation}
883: per unit time. Then the condition that this cooling rate balances the
884: heating rate in Eq.~(\ref{measurementheating}) implies
885: \begin{equation}
886: \langle {E}_\mathrm{eff}\rangle_\mathrm{ss}
887: = \frac{E_0}{\sqrt{1-\beta^2}}
888: \label{E0ssvar}
889: \end{equation}
890: for the steady-state energy. But in the regime where these theories
891: are valid (i.e., $\langle {E}_\mathrm{eff}\rangle_\mathrm{ss}-E_0\ll
892: E_0$ or
893: $\beta \ll 1$),
894: the centroid result of Eq.~(\ref{E0ss}) is always larger than the
895: squeezing result of Eq.~(\ref{E0ssvar}). Thus, we expect that
896: the centroid limit [Eq.~(\ref{E0ss})] is the more appropriate cooling limit.
897:
898: Eq.~(\ref{E0ss}) predicts that the atomic motion will cool essentially
899: to the ground state so long as $\beta \alt 1/2$.
900: Thus, as various parameters are changed, there are ``border'' values
901: that divide the parameter space according to whether or not
902: the atomic motion is cooled to the ground state. For example,
903: the atom will cool to nearly the ground state so long as
904: the ``bang amplitude'' $\varepsilon$ is larger than the border value
905: \begin{equation}
906: \varepsilon_\mathrm{b} = \Gamma\kt^4.
907: \end{equation}
908: Otherwise, the steady-state energy should be substantially higher,
909: and the atomic motion may not even cool at all. However, this simple
910: theory is only valid at small energies, so such behavior should not
911: necessarily be predicted well by this theory. Additionally, this
912: theory does not account for how well the Gaussian estimator
913: tracks the true atomic state, and thus we would expect the actual
914: steady-state temperatures to be higher than predicted by
915: Eq.~(\ref{E0ss}).
916:
917: One final modification to this cooling theory is necessary, since as
918: we discuss below, parity considerations show that half of the atoms
919: can cool at best to the first excited band of the optical potential.
920: To take this effect into account, we simply modify Eq.~(\ref{E0ss})
921: by replacing $E_0$ by the average of the ground and first excited
922: band energies, $(E_0+E_1)/2$. But in the harmonic approximation,
923: $E_0 = \pi$ and $E_1 = 3\pi$, so the cooling limit becomes
924: \begin{equation}
925: \langle {E}_\mathrm{eff}\rangle_\mathrm{ss}
926: = \frac{2\pi}{1-\beta},
927: \label{E01ss}
928: \end{equation}
929: This modification is simple because the centroid argument above only
930: assumed that the atomic state is near the band with the lowest
931: achievable energy. The odd-parity atoms have the first excited state
932: as their lowest achievable state, and the above argument applies just
933: as well to states near this effective ground state.
934: Thus, the ensemble average amounts to simply averaging over the two
935: possible steady-state energies.
936:
937:
938: \section{Simulations: Adiabatic Approximation}\label{sim1}
939:
940: \subsection{Stochastic Schr\"odinger Equation}
941:
942: Simulating the simplified dynamics in the adiabatic approximation according to
943: the SME in Eq.~(\ref{adiabaticSME}) is much less numerically intensive than
944: simulating the full SME given in Eq.~(\ref{scSME}). Because of this, we will
945: perform an extensive analysis of the behavior of the feedback control in this
946: section using simulations in the adiabatic approximation. We will
947: then verify that these simulations indeed provide a good description of the
948: dynamics by performing a limited number of simulations of the full SME in the
949: following section.
950:
951: To perform our simulations, we first note that we may view any SME as being
952: generated by averaging a stochastic Schr\"{o}dinger equation (SSE), where the
953: average is taken over all the signals that the observer fails to
954: measure~\cite{Barchielli93, Wiseman94b}. The unobserved signals in our case are
955: the spontaneous emission from the atom and the part of the cavity output that is not
956: measured due to photodetector inefficiency ($\eta<1$).
957: We therefore replace nature with a formal omniscient observer,
958: unknown to the observer who will effect control.
959: The omniscient observer measures everything that the control observer does not, and also
960: has access to the control observer's measurement record. From
961: the omniscient observer's point of view,
962: the evolution of the system is described by an SSE.
963: We can simulate the system
964: by integrating this SSE, and even though it does not provide us with
965: the control observer's state of knowledge, which would require
966: averaging over all possible measurement results of the omniscient
967: observer, it does generate instances of the control
968: observer's appropriate measurement record.
969: The control observer can then use this measurement record to obtain an
970: estimate of the state of the system by integrating the Gaussian
971: estimator derived above~\cite{Doherty00}. Since the ``correct''
972: state estimate, given by integrating
973: the full SME, is not required,
974: we can obtain a full simulation of the feedback control
975: process by integrating the SSE, which requires far fewer resources
976: than integrating the SME.
977:
978: The normalized Schr\"odinger equation corresponding to the
979: master equation (\ref{adiabaticSME}) that we integrate numerically
980: is
981: \begin{equation}
982: \setlength{\arraycolsep}{0ex}
983: \renewcommand{\arraystretch}{2.1}
984: \begin{array}{rcl}
985: d\psiket &{}={}& \displaystyle -iH\psiket \,dt
986: -\frac{\Gamma}{4}\cos^2(2\kt X)\psiket\,dt\\&&
987: \displaystyle
988: +\frac{\Gamma}{2}\langle\cos(2\kt X)\rangle
989: \cos(2\kt X)\psiket\,dt\\&&
990: \displaystyle
991: +\frac{\Gamma}{2}\langle\cos(2\kt X)\rangle^2 \psiket\,dt \\&&
992: \displaystyle
993: +\sqrt{\frac{\Gamma}{2}} \left[
994: \langle\cos(2\kt X)\rangle - \cos(2\kt
995: X)\right]\psiket\,dW,
996: \end{array}
997: \label{adiabaticSSE}
998: \end{equation}
999: which generates trajectories with the same measure as the SME.
1000: Since we are integrating the SSE in lieu of the SME, there are a few
1001: subtleties related to the measurement noise to be accounted for.
1002: In particular, the Wiener increment $dW$ in the SSE (\ref{adiabaticSSE})
1003: is not equivalent to the Wiener increment in expression (\ref{mradiab})
1004: for the measurement record
1005: if $\eta<1$, because the measured noise does not fully represent the
1006: ``true'' noise. Rather, the quantity $\sqrt{\eta}\,dW$ in the
1007: photocurrent expression
1008: should be replaced by a measured noise $dW_\eta$, given in terms of
1009: the full noise of the SSE by
1010: \begin{equation}
1011: dW_\eta = \eta\, dW + \sqrt{\eta(1-\eta)}\, dW_\mathrm{aux},
1012: \end{equation}
1013: where $dW_\mathrm{aux}$ is an auxiliary Wiener increment, so that
1014: $dW_\eta^2 = \eta dt$, and a fraction
1015: \begin{equation}
1016: dW_\mathrm{unm} = (1-\eta)\, dW - \sqrt{\eta(1-\eta)}\, dW_\mathrm{aux}
1017: \end{equation}
1018: of the true noise is not measured by the control observer (i.e.,
1019: $dW_\eta + dW_\mathrm{unm} = dW$).
1020: The photocurrent in Eq.~(\ref{mradiab}) then becomes
1021: \begin{equation}
1022: d\tilde{r}(t) = - \sqrt{8\eta^2\Gamma}
1023: \langle
1024: \cos^2(\tilde{k}X)\rangle dt + \, dW_\eta ,
1025: \label{mragain}
1026: \end{equation}
1027: and the estimated Wiener increment is given in terms of the measured
1028: noise as
1029: \begin{equation}
1030: \setlength{\arraycolsep}{0ex}
1031: \begin{array}{rcl}
1032: \dWe &{}={}& \displaystyle
1033: \frac{dW_\eta}{\sqrt{\eta}} + \sqrt{2\eta\Gamma}\left[
1034: \exp(-2\kt^2\Vxe)\cos(2\kt\Xe)\right. \\&&
1035: \hspace{40mm}{}- \displaystyle\left.
1036: \langle\cos(2\kt X)\rangle\right] dt,
1037: \end{array}
1038: \label{dWefromdWmod}
1039: \end{equation}
1040: which still reflects the difference between the actual and estimated
1041: quantum states in the same way as before.
1042:
1043:
1044: \subsection{Details of the Simulations}\label{sec:details}
1045:
1046: For the simulations, we choose a ``canonical'' set
1047: of experimentally realistic parameters that will put us in the regime
1048: where adiabatic elimination of the cavity and internal atomic states
1049: should be valid.
1050: Using the Cesium D$_2$ line as the
1051: atomic transition, we have
1052: $m=2.21\times 10^{-25}\;\mathrm{kg}$ and
1053: $\omega_0/2\pi=351.7\;\mathrm{THz}$; we assume that the cavity
1054: subtends a small solid angle and thus that the spontaneous
1055: emission rate is given approximately by the free-space value of
1056: $\gamma/2\pi=5.2\;\mathrm{MHz}$.
1057: For the cavity parameters,
1058: we choose values similar to those used in recent
1059: CQED experiments \cite{Mabuchi99, Hood00, Hood98},
1060: yielding an energy decay rate of $\kappa/2\pi=40$~MHz,
1061: an atom-field coupling constant of $g/2\pi = 120$~MHz,
1062: and a mean intracavity photon number of $\alpha=1$.
1063: However, we will consider a much larger detuning of
1064: $\Delta/2\pi = 4\;\mathrm{GHz}$ to the red of the atomic resonance
1065: than was used in the experiments, in order to work in a dispersive
1066: regime where the adiabatic approximation should work well.
1067: The scaled parameters corresponding to these physical values are
1068: $\Gamma=23.6$ and $\kt = 0.155$. The scaled depth of the
1069: optical potential is $V_\mathrm{max} = \pi/\kt^2 = 131$, which
1070: corresponds to an atomic speed of $14.7\;\mbox{cm s}^{-1}$ and
1071: is of sufficient depth that the lowest two band energies
1072: $E_0 = 3.12$ and $E_1 = 9.33$ are close to the corresponding values
1073: in the harmonic approximation of $\pi$ and $3\pi$, respectively
1074: (with
1075: 27 trapped bands in the optical potential and a ground-band width of
1076: $1.7\times 10^{-12}$, so that tunneling transitions between wells are
1077: highly suppressed).
1078: Also, for the canonical parameter set we choose a feedback
1079: amplitude $\varepsilon=0.1$ (easily accomplished experimentally)
1080: and the idealized detection
1081: efficiency of $\eta = 1$. The goal of our analysis will be to
1082: understand the dynamics of the system under these ideal conditions
1083: and then to understand how variations in these parameters
1084: influence the dynamics.
1085:
1086: The optical potential in the simulations spanned 24 wells,
1087: corresponding to the cavity lengths in the
1088: Caltech experiments \cite{Hood98}. To crudely model the
1089: sticking behavior of an atom when colliding with a mirror,
1090: we implemented absorbing boundary conditions, where the
1091: absorption ramped smoothly on in the last half of each
1092: of the boundary potential wells. We explicitly conditioned
1093: the simulations on the fact that they were not lost, mimicking
1094: the experimental situation where a run with a lost atom would
1095: simply be repeated. However, given the fact that the cavity
1096: is relatively long compared to the optical potential period,
1097: these details do not substantially influence the results of
1098: the calculation.
1099:
1100: To understand the typical dynamics of this problem, we consider
1101: ensemble averages.
1102: We consider the experimental situation where
1103: atoms dropped one at a time
1104: have substantial motional energy after loading into the optical
1105: potential, but only atoms that are well trapped are candidates for
1106: further cooling. Thus, we assume that all atoms in the simulation
1107: ensemble have a centroid energy
1108: of $84.2$, which corresponds to a
1109: stationary atom at a distance of 6 away from the bottom of a well
1110: (where the well period is $\pi/\kt = 20.3$),
1111: and we distribute the initial locations of the atoms uniformly
1112: between the bottom of the well and this maximum distance.
1113: The atoms in the simulation begin
1114: in a coherent state with $V_x = V_p = 1/2$
1115: in the center well of the cavity. We do not expect that the
1116: steady-state results are sensitive to the details of these initial
1117: conditions.
1118: Except where noted otherwise we plot averages over 128 trajectories;
1119: in plots where we inspect the variation of ensemble averages as
1120: a function of a parameter, each ensemble average is computed with
1121: respect to an independent set of random-number seeds.
1122:
1123: We obtained numerical solutions to the SSE (\ref{adiabaticSSE})
1124: using the order 1.5 (strong), implicit,
1125: stochastic Runge-Kutta (SRK) algorithm
1126: in Ref.~\cite{Kloeden92} as part of an operator-splitting
1127: method. To evolve the system over a large step of
1128: $\Delta t = 0.0005$, we first used the SRK evolver to evolve the wave
1129: function according to the SSE without the kinetic $\pi P^2$ term
1130: in substeps of size $\Delta t/8$ over an interval $\Delta t/2$.
1131: Next, we applied the operator $\exp(-i\pi P^2\Delta t)$ after
1132: a Fourier transform to momentum space. Finally, after transforming
1133: back, we again evolved according to the SRK algorithm by another
1134: interval $\Delta t/2$ (in four steps) to complete the full $\Delta t$ step.
1135: This operator splitting preserves the order 1.5 (temporal)
1136: convergence of the SRK approximation.
1137: The spatial grid contained 2048 points. The space and time
1138: discretizations were found to give adequate convergence when
1139: computed in 32-bit precision for most of the trajectories,
1140: although a minority (typically only a few percent) displayed
1141: some sensitivity to these numerical parameters due to their
1142: proximity to unstable hyperbolic points in phase space.
1143: However, for the cooling simulations, the steady-state
1144: energies were well converged and not affected by this sensitivity,
1145: since the atoms only saw the harmonic portions of the potential wells.
1146:
1147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1148: % Basic tracking figure link inserted here.
1149: % This link uses the graphicx package.
1150: \begin{figure}[tb]
1151: \begin{center}
1152: \includegraphics[scale=0.52]{fig1.eps}
1153: \end{center}
1154: \vspace{-5mm}
1155: \caption
1156: {(Color online) Example of locking behavior of the Gaussian estimator
1157: for centroids and variances. The heavy lines (a) are the
1158: expectation values for the ``true'' wave packet, while the
1159: light lines (b) are the corresponding Gaussian estimator
1160: quantities. The improved feedback cooling algorithm
1161: was switched on at $t=2$. The bottom graph shows the
1162: corresponding photodetector signal increments $\Delta\tilde{r}$,
1163: plotted for the time increments of $\DtG = 0.0005$
1164: used by the estimator.
1165: \label{fig:basictracking}}
1166: \end{figure}
1167: % data is hires version of trajectory 18 for the basic
1168: % cooling run
1169: % Note that code puts out dr, but we plot d~r = dr/38.145
1170: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1171:
1172: The Gaussian estimator begins in each case with the impure-state
1173: initial conditions $\Xe = 6$, $\Pe = 0$, $\Vxe = \Vpe = 1/\sqrt{2}$,
1174: and $\Ce = 0$.
1175: As we argued above in Section~\ref{section:gest},
1176: this choice of the initial estimate
1177: is arbitrary but sufficient for our purposes, and we will see
1178: in the next section that the cooling performance is insensitive
1179: to any ``reasonable'' initial estimate.
1180: The estimator is updated with a time step of $\DtG=\Delta t$
1181: using the simple Euler method \cite{Kloeden92}
1182: (both to keep the calculations simple and
1183: because, as we discussed in Section~\ref{section:gest},
1184: $\dWe$ is not the usual Wiener increment and thus violates assumptions
1185: used in constructing higher-order integrators).
1186: Some extra measures are needed to guard against instabilities
1187: in the estimator evolution, which are due
1188: to two possible reasons. First, the
1189: low-order numerical integration of the stochastic estimator equations
1190: with a relatively large time step tends to be unstable, especially
1191: when the occasional large fluctuation occurs. Second,
1192: while Gaussian approximations to Hamiltonian evolution (or even
1193: evolution under an unconditioned Lindblad master equation)
1194: can be thought of as resulting from a variational principle, and
1195: hence inherently stable, the Gaussian approximation to the
1196: conditioned SME cannot, and thus does not inherit these strong
1197: stability properties.
1198: We explicitly detect the four unphysical conditions: negative
1199: position variance ($\Vxe<0$), negative momentum variance ($\Vpe<0$),
1200: phase-space area below the Heisenberg limit
1201: ($\Ae := [\Vxe\Vpe-\Ce^2]^{1/2} < 1/2$),
1202: and large position variance ($\Vxe \gg 1$, indicating the atom is not
1203: localized within the cavity).
1204: If any of these conditions are detected, the variances are reset
1205: according to the initial values $\Vxe = \Vpe = 1/\sqrt{2}$, $\Ce =
1206: 0$, while the means $\Xe$ and $\Pe$ are not modified; the measurement
1207: information will quickly restore the correct estimator values.
1208: For a typical
1209: cooling simulation with the canonical parameters, the estimator
1210: must be reset in about 95\% of the trajectories, with trajectories
1211: on average requiring 3 resets due to the area condition (the other
1212: conditions producing negligible numbers of resets in comparison).
1213: Also, to help alleviate the number of resets caused by this condition,
1214: we instead implement the condition that resets only occur when $\Ae < 1/4$.
1215: This situation may seem curious, as the Gaussian area evolves
1216: according to
1217: \begin{equation}
1218: \setlength{\arraycolsep}{0ex}
1219: \renewcommand{\arraystretch}{1.4}
1220: \begin{array}{rcl}
1221: d(\Ae^2) &{}={}& \Gamma\kt^2\Vxe[1-\exp(-8\kt^2\Vxe)\cos(4\kt\Xe)] dt\\
1222: && -8\eta\Gamma\kt^2\Vxe\Ae\exp(-4\kt^2\Vxe)\sin^2(2\kt\Xe) dt\\
1223: && -8\eta\Gamma\kt^4\Vxe^2[1-4\kt^2\Vxe\exp(-2\kt^2\Vxe)]\\
1224: &&\hspace{2mm}\times \exp(-2\kt^2\Vxe)
1225: \cos^2(2\kt\Xe) dt\\
1226: && -\sqrt{2\eta\Gamma}\kt^2\Vxe[1-4(\kt^2\Vxe+\Ae)\exp(-2\kt^2\Vxe)]\\
1227: &&\hspace{2mm} \times \cos^2(2\kt\Xe) \dWe ,
1228: \end{array}
1229: \end{equation}
1230: whence it follows after some examination that $\Ae \ge 1/2$ for all
1231: time if this is true initially.
1232: Thus, these resets are numerical artifacts of evolving the estimator using
1233: a finite time step and a low-order numerical method;
1234: in the simulations, reducing the time step $\DtG$ by a factor of 10
1235: reduces the average number of resets due to the area condition to 1.1
1236: (with 0.8 resets on average due to the $\Vxe<0$ condition). However, we
1237: continue to use the larger step size,
1238: which is more reasonable to
1239: implement in practice given technological constraints on the speed of
1240: calculations for the forseeable future.
1241: For somewhat smaller measurement efficiencies,
1242: the estimator requires fewer resets (e.g., about 12\% of the
1243: trajectories for $\eta = 0.8$, and only 1.7 resets on average per
1244: trajectory), since the steady-state area is
1245: larger (corresponding to a state with lower purity, as
1246: $\mathrm{Tr}[\rho_\mathrm{e}^2] = 1/(2\Ae)$ \cite{Zurek93}, where
1247: $\rho_\mathrm{e}$ is the estimated density operator).
1248:
1249: Because the
1250: estimator locks on quickly with the canonical parameters, the feedback
1251: cooling algorithm is turned on at time $t=2$. For the improved
1252: cooling algorithm, we find that fitting to the last 300 observations
1253: (spaced apart in time by $\DtG$) is optimal.
1254:
1255:
1256: \subsection{Estimation and Cooling Dynamics}
1257: \label{sec:estimation-cooling-dynamics}
1258:
1259: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1260: % Basic cooling figure link inserted here.
1261: % This link uses the graphicx package.
1262: \begin{figure}[tb]
1263: \begin{center}
1264: \includegraphics[scale=0.41]{fig2.eps}
1265: \end{center}
1266: \vspace{-5mm}
1267: \caption
1268: {(Color online) Evolution of the mean energy
1269: $\langle H_\mathrm{eff}\rangle$ (relative to the
1270: minimum potential energy), comparing the performance
1271: of different cooling algorithms.
1272: Heavy solid line (a): improved cooling algorithm.
1273: Solid line (b): centroid-only cooling algorithm.
1274: Dashed line (c): no cooling algorithm. Each line is an
1275: average of 128 trajectories; parameters are for the
1276: canonical set.
1277: \label{fig:basiccooling}}
1278: \end{figure}
1279: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1280:
1281:
1282: Now we will evaluate the performance of the two main components of the
1283: quantum feedback cooling system, the estimator and the cooling
1284: algorithm. Henceforth, we will only study the performance of the
1285: improved cooling algorithm, except where noted otherwise.
1286: Fig.~{\ref{fig:basictracking} shows a typical trajectory
1287: undergoing cooling}
1288: and compares some of the ``true'' moments computed from the atomic
1289: wave function (i.e., those computed with respect to the result of
1290: integrating the SME of the omniscient observer)
1291: with the moments from the estimator. The centroids
1292: and variances of the estimator lock on by the time that the cooling
1293: algorithm is switched on ($t=2$), and the estimator provides a
1294: reasonably faithful representation of the atomic dynamics.
1295: It is especially important that the variances track the atomic
1296: wave packet accurately in order to make the improved cooling
1297: work effectively.
1298: Fig.~\ref{fig:basictracking} also shows the measurement record
1299: increments
1300: \begin{equation}
1301: \Delta\tilde{r}(t) := \int_t^{t+\DtG} d\tilde{r}(t)
1302: \end{equation}
1303: corresponding to the time increments $\DtG = 0.0005$ used in
1304: the Gaussian-estimator integration. The measurement record
1305: appears to be noise-dominated, but nevertheless contains
1306: information about the quantum state in the time-dependent offset
1307: level of the noise.
1308: The action of the estimator is such that it ``demultiplexes'' the
1309: information about the motions of the centroids and variances, even though they
1310: are encoded in the same frequency range in the photocurrent signal.
1311: One can get a sense for how this works from the form of Eqs.~(\ref{gest}).
1312: For example, the measurement (\dWe) terms in the $d\Xe$ and $d\Vxe$
1313: equations are weighted by $\sin(2\kt\Xe)$ and $\cos(2\kt\Xe)$,
1314: respectively. Thus, the incoming measurement information
1315: influences $\Xe$ or $\Vxe$, depending on the state of the estimator:
1316: if the estimator is centered in a well, variations in
1317: $\langle\cos^2\kt X\rangle$ (i.e., the photocurrent) are presumed
1318: to be caused by variations in $V_x$, while such variations are
1319: attributed to $\langle X\rangle$ if the estimator is centered on the
1320: side of a well.
1321: The estimator can also gain information from the
1322: absolute value of the measurement record---as we discussed
1323: in Section~\ref{section:limits}, energy in only the variance
1324: degree of freedom can produce transient values of $\Vmax\langle[1-
1325: \cos^2 (\kt X)]\rangle$ below $E_0/2$, whereas energy
1326: in only the centroid cannot.
1327: This subtle unraveling of the measurement information into the
1328: estimated state is automatic in the measurement formulation we
1329: have used here.
1330:
1331: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1332: % Vary-the-Gaussian-estimator figure link inserted here.
1333: % This link uses the graphicx package.
1334: \begin{figure}[tb]
1335: \begin{center}
1336: \includegraphics[scale=0.41]{fig3.eps}
1337: \end{center}
1338: \vspace{-5mm}
1339: \caption
1340: {(Color online) Evolution of the mean energy
1341: $\langle H_\mathrm{eff}\rangle$, comparing the
1342: performance of the improved cooling algorithm with
1343: different initial (Gaussian) state estimates.
1344: Heavy solid line (a): estimate matches the canonical parameter
1345: set.
1346: Solid line (b): initial estimate is much more impure and
1347: substantially overlaps $(x,p)=(0,0)$.
1348: Dashed line (c): same as solid line, but centered on $(x,p)=(0,0)$.
1349: \label{fig:initcondcheck}}
1350: \end{figure}
1351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1352:
1353: Fig.~\ref{fig:basiccooling} compares the ensemble-averaged
1354: evolution of the energy
1355: $\langle E_\mathrm{eff}\rangle$ in the case where the atoms are
1356: not cooled to the cases where the atoms are cooled by the centroid-only
1357: and improved cooling algorithms. When the atoms are not actively
1358: cooled, the energy simply increases linearly in time.
1359: The initial heating rate of $\partial_t\langle
1360: E_\mathrm{eff}\rangle=1.78$
1361: predicted by Eq.~(\ref{uniformheating}) for measurement-backaction
1362: is substantially smaller than the initial simulated heating rate of
1363: 2.1(1). We can understand this discrepancy, since the initial
1364: conditions in the simulation are uniformly distributed in phase
1365: space along a constant-energy surface, weighting more
1366: heavily the gradients of the
1367: potential where heating is maximized; Eq.~(\ref{uniformheating})
1368: underestimates the heating rate by assuming a uniform distribution in
1369: configuration space.
1370:
1371: The simple, centroid-based cooling algorithm cools the atoms rapidly
1372: for short times, but around $t=10$, the cooling stops and
1373: the atoms begin to heat back up, apparently without bound.
1374: Clearly, we expect some heating from the squeezing effects
1375: discussed above. Furthermore, we expect that when the centroid
1376: cools and the variance becomes much larger than the mean-square
1377: amplitude of centroid motion, the centroid evolution will be dominated
1378: by measurement (projection) noise, and thus the feedback signal will
1379: also be determined mostly by random noise from the measurement.
1380: In this state, the cooling algorithm is ineffective and leads to
1381: heating.
1382: This process runs away, since there
1383: is no mechanism to cool the energy associated with the variances
1384: (squeezing).
1385:
1386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1387: % Basic direct cooling figure link inserted here.
1388: % This link uses the graphicx package.
1389: \begin{figure}[tb]
1390: \begin{center}
1391: \includegraphics[scale=0.41]{fig4.eps}
1392: \end{center}
1393: \vspace{-5mm}
1394: \caption
1395: {(Color online) Evolution of the mean energy
1396: $\langle H_\mathrm{eff}\rangle$, comparing the
1397: performance of the improved cooling algorithm
1398: (heavy solid line, a) with that of cooling based
1399: directly on the homodyne signal (solid line, b).
1400: Heavy solid line: improved cooling algorithm.
1401: Each line is an
1402: average of 128 trajectories; parameters are for the
1403: canonical set.
1404: \label{fig:directcooling}}
1405: \end{figure}
1406: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1407:
1408: On the other hand, the improved cooling algorithm also rapidly
1409: cools the atoms and the atomic energies settle to a well-defined
1410: steady state,
1411: which already demonstrates the potential utility of
1412: this cooling algorithm for long-term storage of the atoms.
1413: For the moment, we will defer the question of how closely the
1414: atoms are cooling to the lowest energy band of the optical
1415: potential until we take a closer look at the atomic dynamics
1416: in the next section.
1417:
1418: The cooling behavior that we see is also insensitive to the
1419: exact choice of the initial (Gaussian) state estimate, so long
1420: as we have made a reasonable choice. To illustrate this statement,
1421: Fig.~\ref{fig:initcondcheck} shows the ensemble-averaged cooling
1422: dynamics for two initial state estimates in addition to the
1423: canonical estimate that we use for the rest of the simulations.
1424: In the first modified initial estimate, we have selected
1425: $\Xe = 0.1$, $\Pe = 0$, $\Vxe = 1$, $\Vpe = 1$, and $\Ce = 0$
1426: for a much larger uncertainty product (lower purity $\mathrm{Tr}[\rho^2]$)
1427: and a substantial overlap with the other side of the potential well.
1428: We see that although the short-time cooling performance is slightly
1429: worse, the long-time cooling performance is the same as for the
1430: original initial estimator. This implies that the tracking behavior
1431: is initially worse for this initial estimator, but at sufficiently
1432: long times the estimator ``forgets'' its initial state and does a
1433: better job of tracking. The other initial estimator is the same
1434: as the previous case, but with $\Xe=0$, so that the estimator is
1435: centered on the origin in phase space. Examination of the
1436: estimator equations (\ref{gest}) reveals that in this case the
1437: centroids $\Xe$ and $\Pe$ will remain zero for all time, effectively
1438: ``freezing out'' the estimator's centroid degree of freedom.
1439: The cooling behavior here is much worse than in the other two cases,
1440: demonstrating both that the estimator's variance degree of freedom is
1441: insufficient to mimic the dynamics of the true state and that it is
1442: important to exercise some caution when selecting an initial state
1443: estimate.
1444:
1445: Finally, we return to the question raised in
1446: Section~\ref{section:alg} regarding the necessity of using
1447: the Gaussian estimator at all, in light of the fact that the
1448: measurement record itself provides a direct, albeit very noisy,
1449: measurement of the quantity $\langle \cos^2\kt X\rangle$ used
1450: to determine the control switching times in the improved
1451: cooling algorithm. Fig.~\ref{fig:directcooling}
1452: address this question by comparing the energy evolution under the
1453: improved cooling algorithm using the Gaussian estimator with
1454: the same algorithm but based directly on the measurement record.
1455: In both cases, the quadratic curve was fitted on-the-fly
1456: to the last 300 points in the measurement record to reduce noise
1457: effects.
1458: We found this number of points to be optimal for the canonical
1459: parameter set, and this optimal time interval over which the data
1460: are fit is tied directly to the time scales of the atomic motion.
1461: Although the direct-cooling case cools rapidly and settles to a lower
1462: temperature, it is clear that the cooling performance
1463: is greatly improved by the Gaussian estimator. The estimator
1464: %,
1465: %although giving at best only an approximate picture of the
1466: %true quantum state,
1467: acts as a nearly optimal noise filter,
1468: efficiently extracting the relevant information from the
1469: noisy measurement record.
1470:
1471:
1472:
1473: \subsection{Parity Dynamics}
1474:
1475: We now consider the evolution of the
1476: parity of the conditioned atomic state due to the measurement.
1477: For this discussion, we will focus on the expectation values
1478: of the usual parity operator
1479: \prty\ (where $\prty\psi(x) = \psi(-x)$). For numerical purposes,
1480: though, we will instead use the \textit{reduced} parity operator,
1481: $\prty':=\prtyr$, where the reduction operator \reduceop\ is defined by
1482: \begin{equation}
1483: \reduceop\psi(x) := \sum_{j=-\infty}^\infty \psi\left(x-\frac{j\pi}{\kt}\right),\;
1484: x\in\left[-\frac{\pi}{2\kt},\frac{\pi}{2\kt}\right),
1485: \end{equation}
1486: which allows us to detect the
1487: parity of the atomic state with respect to $any$ potential well in the
1488: cavity, rather than just the center well.
1489: We will simply use the notation ``\prty'' for both operators
1490: in the discussion, though, as the discussion holds for either operator.
1491: The initial conditions for the simulations, which correspond to
1492: Gaussian wave packets displaced in phase space from the origin
1493: in phase space, represent equal mixtures of odd and even parity
1494: ($\langle \prty\rangle = 0$) to a good approximation.
1495: As time evolves, however, the parity of the conditioned
1496: atomic state evolves (in the adiabatic approximation) according to
1497: \begin{equation}
1498: d\langle\prty\rangle = -\sqrt{8\eta\Gamma}
1499: \left[\langle\prty\cos(2\kt X)\rangle - \langle\prty\rangle
1500: \langle\cos(2\kt X)\rangle\right]\,dW .
1501: \label{dparity}
1502: \end{equation}
1503: This effect is due only to the nonlinearity of the measurement
1504: term in the master equation (\ref{adiabaticSME}); the parity
1505: under unconditioned evolution is invariant.
1506:
1507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1508: % Parity figure link inserted here.
1509: % This link uses the graphicx package.
1510: \begin{figure}[tb]
1511: \begin{center}
1512: \includegraphics[scale=0.41]{fig5.eps}
1513: \end{center}
1514: \vspace{-5mm}
1515: \caption
1516: {(Color online) Histograms of the reduced parity $\langle\prty'\rangle$
1517: for an ensemble of 2048 trajectories undergoing
1518: cooling, showing how parities evolve towards
1519: the extreme states of pure parity at late times.
1520: \label{fig:parity}}
1521: \end{figure}
1522: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1523:
1524: The reduced parity evolution for a simulated ensemble of trajectories
1525: undergoing measurement and cooling is shown in Fig.~\ref{fig:parity}.
1526: It is clear that $\langle\prty\rangle$ evolves towards the extreme
1527: values $\pm 1$ of parity for the simulated trajectories.
1528: This ``parity purification'' behavior is not immediately clear
1529: from the evolution equation (\ref{dparity}), which states that
1530: $\langle\prty\rangle$ evolves according to a diffusion process.
1531: The key point is that this diffusion process is nonstationary,
1532: since the diffusion rate depends on the parity, and indeed vanishes
1533: for states with pure parity. To see the final steady state directly,
1534: we consider the projectors $\prty_+$ and $\prty_-$ for even and
1535: odd parity, respectively, defined by
1536: \begin{equation}
1537: \prty_\pm := \frac{1\pm\prty}{2}.
1538: \end{equation}
1539: Then the product of the expectation values of these projectors
1540: evolves according to
1541: \begin{equation}
1542: d\left[\langle\prty_+\rangle\langle\prty_-\rangle\right] =
1543: -8\eta\Gamma
1544: \left(\langle\prty_+\rangle\langle\prty_-\rangle\wp\right)^2
1545: dt,
1546: \label{dampparity}
1547: \end{equation}
1548: where the parity difference $\wp$ is given for states with
1549: $\langle\prty_\pm\rangle\neq 0$ by
1550: \begin{equation}
1551: \wp := \frac{\langle\prty_+\cos^2\kt X\rangle}{\langle\prty_+\rangle}
1552: - \frac{\langle\prty_-\cos^2\kt X\rangle}{\langle\prty_-\rangle}.
1553: \end{equation}
1554: The magnitude of $\wp$
1555: represents how well the measurement of $\cos^2\kt X$ can
1556: resolve the parity of the atomic state, and it is nonzero for
1557: generic states.
1558: Thus, the quantity $\langle\prty_+\rangle\langle\prty_-\rangle$
1559: damps monotonically to zero, which is only possible if
1560: the parity of the state becomes either pure even or pure odd. Given
1561: the initial conditions in the simulations, and the fact that the
1562: diffusion process in Eq.~(\ref{dparity}) is unbiased, we can expect
1563: that half of the trajectories will become even and half will become
1564: odd. As we mentioned above, this argument also applies to both the
1565: standard and reduced parity operators.
1566: However, the damping rate in Eq.~(\ref{dampparity}) is greatly reduced
1567: for the standard parity operator when considering
1568: states localized in potential wells not centered on $X=0$.
1569:
1570: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1571: % Band populations figure link inserted here.
1572: % This link uses the graphicx package.
1573: \begin{figure}[tb]
1574: \begin{center}
1575: \includegraphics[scale=0.41]{fig6.eps}
1576: \end{center}
1577: \vspace{-5mm}
1578: \caption
1579: {(Color online) Evolution of the band populations for the lowest
1580: two (0 and 1) and next two (2 and 3) energy bands
1581: of the optical lattice, averaged over 128 trajectories.
1582: Top: cooling based on the Gaussian estimator. Bottom:
1583: cooling based on perfect knowledge of the actual
1584: wave function.
1585: \label{fig:bandpops}}
1586: \end{figure}
1587: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1588:
1589: This purification effect obviously has consequences for cooling the
1590: atomic motion. Even if the cooling algorithm can, in principle, cool
1591: the atomic motion faster than it is heated by the measurement, the
1592: cooling algorithm cannot affect the parity of the atomic state, and
1593: thus odd-parity states will not be cooled to the ground state of the
1594: optical potential.
1595: At best, then, the cooling algorithm can cool the atom to the
1596: ground energy band with probability $1/2$ and the first excited
1597: energy band with probability $1/2$.
1598: This cooling behavior is illustrated in Fig.~\ref{fig:bandpops},
1599: where the evolutions of the band populations are plotted for the
1600: two
1601: cases where cooling is based on the Gaussian estimator and
1602: directly on the true atomic wave function using the improved
1603: algorithm in both cases. We can clearly see that in steady state
1604: the lowest and first excited bands are about equally
1605: populated, with these bands accounting for 94\% %(\pm 5%)
1606: of the population
1607: in the Gaussian-estimator case and 98\% of the population in the
1608: perfect-knowledge case. Thus, this parity effect is the
1609: dominant limit to the atomic motional energy for the ensemble.
1610: However, this situation is essentially
1611: as good as cooling fully to the ground band, because the measurement
1612: can distinguish between the two possible outcomes (i.e., by resolving
1613: the two different possible steady-state potential energies). Experimentally,
1614: if the odd-parity outcome is detected, the atom can be transferred to the
1615: ground band by a coherent process (e.g., resonant modulation
1616: of the optical potential by an external field or two-photon, stimulated
1617: Raman transitions), or the state can simply be rejected, in hopes of
1618: obtaining the even-parity outcome on the next trial.
1619:
1620: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1621: % Parity expectations figure link inserted here.
1622: % This link uses the graphicx package.
1623: %\begin{figure}[tb]
1624: % \begin{center}
1625: % \includegraphics[scale=0.5]{parity_thangr.eps}
1626: % \end{center}
1627: % \vspace{-5mm}
1628: % \caption
1629: % {Evolution of the normalized parity expectation values
1630: % ${\langle\prty_+\cos^2\kt X\rangle}/{\langle\prty_+\rangle}$
1631: % (solid line) and
1632: % ${\langle\prty_-\cos^2\kt X\rangle}/{\langle\prty_-\rangle}$
1633: % (dashed line), averaged over 128 trajectories.
1634: % \label{fig:paritythang}}
1635: %\end{figure}
1636: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1637:
1638: Finally, we comment on one odd aspect of the
1639: intermediate-time histogram ($t=40$) in Fig.~\ref{fig:parity}, which
1640: shows a marked asymmetry, even though
1641: the \textit{median} value of $\langle\prty\rangle$ is zero.
1642: Because we can write
1643: Eq.~(\ref{dparity}) in the form
1644: \begin{equation}
1645: d\langle\prty\rangle = -\sqrt{8\eta\Gamma}
1646: \langle\prty_+\rangle\langle\prty_-\rangle\wp\,dW ,
1647: \label{dparityagain}
1648: \end{equation}
1649: all of the asymmetry is due to the $\wp$ factor. Since this
1650: diffusion process mimics a quantum nondemolition measurement of the
1651: parity, we can interpret $\wp$ as an effective measurement strength
1652: that is larger for even-parity states than for odd states. This
1653: asymmetry is reasonable since the measurement and cooling processes produce
1654: even-parity states that are more localized than the odd ones. Thus,
1655: the even-parity states, being tightly localized at the field antinodes,
1656: produce more extreme values of the measurement record and are hence
1657: more easily distinguished from mixed-parity states.
1658: Again, however, this effectively asymmetric parity-measurement process
1659: does not affect the long-time outcome that either parity will be
1660: selected with equal probability.
1661:
1662: %This form shows that the diffusion rate is a symmetric function about
1663: %$\langle\prty\rangle = 0$ except for the dependence of $\wp$ on the
1664: %parity. It is through this factor where the cooling can influence the
1665: %parity purification, even though the feedback modulation of the
1666: %optical potential cannot directly modify the atomic parity.
1667: %The two parity expectation values
1668: %${\langle\prty_\pm\cos^2\kt X\rangle}/{\langle\prty_\pm\rangle}$
1669: %comprised in $\wp$ are plotted in Fig.~\ref{fig:paritythang}
1670: %for the case of improved cooling with the Gaussian estimator,
1671: %which shows that the magnitude of $\wp$ (the difference between the
1672: %two curves) is initially very small
1673: %but becomes much larger when the atoms become very cold.
1674: %This is because the quantities
1675: %${\langle\prty_\pm\cos^2\kt X\rangle}/{\langle\prty_\pm\rangle}$
1676: %are given by the energies of the even/odd components of
1677: %the wave function, up to a constant factor.
1678: %Initially, when many energy eigenstates are
1679: %excited, the even- and odd-component energies are similar, and
1680: %thus $\wp$ is small. But
1681: %when the atoms are cooled to the lowest two bands, the even-
1682: %and odd-component energies are just the energies of the ground
1683: %and first-excited bands, respectively, so $\wp$ in steady state
1684: %is approximately $(E_0-E_1)/2 \Vmax$.
1685:
1686:
1687: %\workingcomment{But now we need a better explanation.
1688: %From Fig.~\ref{fig:bandpops}
1689: %we know that the even parity side is cooling more slowly than the
1690: %odd parity side, so $\wp$ should be bigger for the odd states
1691: %at $t=40$. So why does the histogram show more states piled up
1692: %at $\mathcal{P} = +1$?}
1693:
1694:
1695:
1696: \subsection{Tests of Cooling-Limit Theories}
1697:
1698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1699: % Epsilon sweep figure link inserted here.
1700: % This link uses the graphicx package.
1701: \begin{figure}[tb]
1702: \begin{center}
1703: \includegraphics[scale=0.41]{fig7.eps}
1704: \end{center}
1705: \vspace{-5mm}
1706: \caption
1707: {(Color online) Final energies, measured over the interval from $t=90$
1708: to $100$, as the switching amplitude $\varepsilon$
1709: varies; other parameters match the canonical set.
1710: Circles: cooling based on the Gaussian estimator.
1711: Squares: cooling based on perfect knowledge of the
1712: actual wave function. Dashed line: simple cooling
1713: theory, Eq.~(\ref{E01ss}). Inset: magnified view of the
1714: same data. Error bars reflect
1715: standard errors from averages over 128 trajectories.
1716: \label{fig:epsilonsweep}}
1717: \end{figure}
1718: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1719:
1720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1721: % Vmax sweep figure link inserted here.
1722: % This link uses the graphicx package.
1723: \begin{figure}[tb]
1724: \begin{center}
1725: \includegraphics[scale=0.41]{fig8.eps}
1726: \end{center}
1727: \vspace{-5mm}
1728: \caption
1729: {(Color online) Final energies, measured over the interval from $t=90$
1730: to $100$, as the unmodulated optical potential depth
1731: $\Vmax = \pi/\kt^2$
1732: varies; other parameters match the canonical set.
1733: Circles: cooling based on the Gaussian estimator.
1734: Squares: cooling based on perfect knowledge of the
1735: actual wave function. Dashed line: simple cooling
1736: theory, Eq.~(\ref{E01ss}). Error bars reflect
1737: standard errors from averages over 128 trajectories.
1738: \label{fig:vmaxsweep}}
1739: \end{figure}
1740: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1741:
1742: Now we can examine the dependence of the steady-state cooling
1743: energies on the system parameters and test the validity of the
1744: cooling-limit theories in Section~\ref{section:limits}.
1745: Fig.~\ref{fig:epsilonsweep} shows the dependence of the
1746: simulated steady-state temperatures on the control-switching
1747: amplitude $\varepsilon$ in both the Gaussian-estimator
1748: and perfect-knowledge cases. The simple cooling prediction of
1749: Eq.~(\ref{E01ss}) is also shown here for comparison.
1750: The agreement between the theory and the perfect-knowledge
1751: simulations is much better than between the theory and
1752: the Gaussian-estimator simulations, which is sensible because
1753: the simple theory does not attempt to account for the
1754: imperfections in how the estimator tracks the true state.
1755: While the (perfect-knowledge) steady-state energies consistently
1756: lie slightly below the theoretical prediction, the agreement
1757: is overall quite good, and the theory correctly predicts
1758: a sharp transition from efficient to inefficient cooling
1759: near the transition border
1760: $\varepsilon_\mathrm{b} = 0.014$
1761: (corresponding to the point $\beta\sim 1/2$ where the predicted
1762: steady-state energy is twice the smallest predicted value).
1763: Simply put, when the control transitions of the optical
1764: potential are sufficiently weak, the cooling algorithm does
1765: not cool the atoms at a rate sufficient to counteract the measurement
1766: heating (the system is no longer ``controllable''), and so the atoms do not equilibrate at a
1767: very low temperature.
1768: The cooling and transition behaviors are very similar in the
1769: Gaussian-estimator simulations, but these energies are
1770: overall slightly higher, again probably due to the inability of the
1771: Gaussian estimator to perfectly track the atomic state, even
1772: with $\eta = 1$.
1773:
1774: Similar behavior appears as the (unmodulated) optical potential
1775: depth $\Vmax = \pi/\kt^2$ varies, as shown in Fig.~\ref{fig:vmaxsweep}.
1776: For sufficiently large potential depth, the temperature is essentially
1777: constant, but if the potential depth is too weak, we again lose
1778: controllability and the cooling algorithm fails to be effective.
1779: There is a much more substantial difference between the
1780: Gaussian-estimator simulations and the simple cooling theory for
1781: small $\Vmax$ compared
1782: to Fig.~\ref{fig:epsilonsweep}. This discrepancy is most likely
1783: due to the reduced confinement of the wave packet, which produces
1784: stronger anharmonic effects and thus a breakdown of the Gaussian
1785: approximation.
1786: The simulation data are still in good agreement with the simple
1787: theory.
1788:
1789: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1790: % Gamma sweep figure link inserted here.
1791: % This link uses the graphicx package.
1792: \begin{figure}[tb]
1793: \begin{center}
1794: \includegraphics[scale=0.41]{fig9.eps}
1795: \end{center}
1796: \vspace{-5mm}
1797: \caption
1798: {(Color online) Final energies, measured over the interval from $t=90$
1799: to $100$, as the effective measurement strength $\Gamma$
1800: varies; other parameters match the canonical set.
1801: Circles: cooling based on the Gaussian estimator.
1802: Squares: cooling based on perfect knowledge of the
1803: actual wave function. Dashed line: simple cooling
1804: theory, Eq.~(\ref{E01ss}). Error bars reflect
1805: standard errors from averages over 128 trajectories.
1806: \label{fig:gammasweep}}
1807: \end{figure}
1808: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1809:
1810: The dependence of the steady-state energy on the
1811: measurement strength $\Gamma$ is shown in Fig.~\ref{fig:gammasweep}.
1812: In this case, the theory again matches the perfect-knowledge
1813: situation quite well. The cooling is best for moderately small
1814: values of $\Gamma$, and the simple theory predicts a transition
1815: to an uncooled state if $\Gamma$ exceeds the border value
1816: $\Gamma_\mathrm{b} = \varepsilon/\kt^4 = 173$, since $\Gamma$
1817: controls the rate of heating in the system. However, due to the
1818: functional form of Eq.~(\ref{E01ss}), the ``transition'' here is
1819: much more gradual than the transition observed when varying
1820: $\varepsilon$.
1821: The Gaussian-estimator simulations again result in values that
1822: are consistently slightly above the perfect-knowledge energies.
1823: The difference is very pronounced for small $\Gamma$, where we
1824: might expect that the
1825: estimator is not gaining sufficient information from the
1826: measurement and cannot track the true state well enough
1827: to produce effective cooling (i.e., the system is no longer
1828: ``observable'').
1829: The perfect-knowledge energy data also show a slight upturn
1830: for small $\Gamma$ that is not predicted by the simple theory.
1831: This departure indicates the direct importance of the localization
1832: produced by the measurement in the cooling process:
1833: a localized phase-space distribution
1834: is responsible for larger fluctuations in $\langle\cos^2\kt X\rangle$
1835: than a delocalized distribution with the same energy, and from the
1836: discussion in Section~\ref{section:alg}, the
1837: cooling rate is directly proportional to the magnitude of these
1838: fluctuations. Thus, a sufficiently large value of $\Gamma$ must
1839: be chosen to maintain a localized atomic distribution and hence
1840: good cooling.
1841:
1842:
1843: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1844: % Epsilon (centered) sweep figure link inserted here.
1845: % This link uses the graphicx package.
1846: \begin{figure}[tb]
1847: \begin{center}
1848: \includegraphics[scale=0.41]{fig10.eps}
1849: \end{center}
1850: \vspace{-5mm}
1851: \caption
1852: {(Color online) Final energies as in Fig.~\ref{fig:epsilonsweep}, but
1853: with initial position and momentum centroids (for both
1854: the atomic wave packet and Gaussian estimator) at zero.
1855: The dashed line now refers to the variance-dominated,
1856: even-parity theory of Eq.~(\ref{E0ssvar}).
1857: \label{fig:epsilonctrsweep}}
1858: \end{figure}
1859: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1860:
1861: Thus far, we have focused on testing the theory of Eq.~(\ref{E01ss}),
1862: which is the most relevant theory for a physical implementation of
1863: the cavity QED feedback system.
1864: However, we can also test the variance-only theory of
1865: Eq.~(\ref{E0ssvar}) in the simulations, thereby giving indirect
1866: support to the more physically relevant theory.
1867: We can do this by changing the initial condition of the atomic
1868: wave packets to be a coherent (Gaussian) state centered on
1869: one of the potential wells with zero centroid momentum.
1870: From the form of the Gaussian
1871: estimator equations (\ref{E0ssvar}) we see that the wave-packet
1872: centroid will remain fixed at zero position and momentum, and
1873: thus all the energy will be associated with the wave-packet
1874: variances (in the harmonic approximation). The results of simulations
1875: of this type, where we also modify the initial condition of the
1876: Gaussian estimator to have the same centroid, are plotted
1877: in Fig.~\ref{fig:epsilonctrsweep}. The theory of Eq.~(\ref{E0ssvar})
1878: is also shown; note that we do not need to replace the $E_0$
1879: in this expression with the average of $E_0$ and $E_1$, since
1880: the initial conditions have exactly even parity, which is then
1881: time-invariant. The theory is again in good agreement with the
1882: perfect-knowledge data, both in the lower temperatures for
1883: large $\varepsilon$ and the sharper transition and
1884: lower transition border of
1885: $\varepsilon_\mathrm{b} = 0.0079$ (corresponding to
1886: $\beta\sim\sqrt{3}/2$, again when the predicted energy is twice
1887: the minimum predicted value).
1888:
1889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1890: % Band populations for 3 trajectories figure link inserted here.
1891: % This link uses the graphicx package.
1892: \begin{figure}[tb]
1893: \begin{center}
1894: \includegraphics[scale=0.41]{fig11.eps}
1895: \end{center}
1896: \vspace{-5mm}
1897: \caption
1898: {(Color online) Evolution of the ground and first-excited band populations
1899: for three trajectories,
1900: illustrating typical behavior.
1901: Trajectories that purify to even parity (top) tend to
1902: settle to a quiet equilibrium in the ground state.
1903: Trajectories purifying to odd parity (middle) are
1904: characterized by dynamical equilibria, where periods
1905: in the first excited state are interrupted by
1906: episodes of higher energy. Trajectories that
1907: take longer to purify in parity (bottom) cool
1908: rapidly to the lowest two states, after which
1909: population is transferred between these two states
1910: as the parity diffuses.
1911: Parameters here correspond to the canonical set.
1912: \label{fig:bandpoptrajs}}
1913: \end{figure}
1914: % trajectories are #3, 16, 104 from basic_dist_hiresr run
1915: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1916:
1917: Finally, a note about the steady-state temperatures is in order
1918: for the cases where the Gaussian estimator is used for cooling.
1919: The ensemble-averaged, late-time energies in
1920: Figs.~\ref{fig:epsilonsweep}-\ref{fig:epsilonctrsweep}
1921: for Gaussian-estimator cooling are slightly but
1922: consistently higher than the corresponding perfect-knowledge
1923: energies, suggesting that the cooling is somewhat less effective
1924: when the Gaussian estimator is used. However, these averages
1925: mask complicated dynamics that are illustrated in
1926: Fig.~\ref{fig:bandpoptrajs}.
1927: Trajectories that have even parity at late times settle into
1928: a quiet equilibrium with nearly all the population in the ground
1929: energy band. On the other hand, trajectories that settle to odd parity
1930: have a more complicated equilibrium, where the atom is mostly in the
1931: first excited band but the evolution is punctuated by ``bursts'' of
1932: heating. This is an indication that in steady state,
1933: the Gaussian estimator is, unsurprisingly, a much better approximation
1934: for a wave packet close to the ground state than a wave packet close to the
1935: first excited state. As seen in Fig.~\ref{fig:bandpoptrajs}, these
1936: heating episodes correspond to population being transferred to
1937: higher energy states of odd parity, a process distinct from the parity
1938: drift which, after the initial cooling, transfers population between
1939: the ground and first excited states.
1940: Again, if the goal of the cooling process is to prepare an atom in
1941: the ground state, the rate of success is even better than suggested
1942: by the ensemble-averaged energies, since cooling in the even-parity
1943: cases is essentially as good as if one had perfect knowledge of the
1944: actual wave function.
1945:
1946: %Here, two typical trajectories
1947: %being cooled with the Gaussian estimator are shown as they
1948: %cool separately into the ground and first-excited bands.
1949: %The steady state is quite dynamical, though, showing a
1950: %``bursting'' behavior in the case of the odd-parity outcome.
1951: %The even-parity outcome is much quieter, indicating that
1952: %the odd-parity state is much more difficult for the Gaussian estimator
1953: %to track. The trajectories settle predominantly
1954: %into the lowest possible band (corresponding to the selected
1955: %parity), but occasionally, the estimator fails to track
1956: %the atomic behavior, causing the atom to heat up. Thus,
1957: %the band populations are punctuated by intermittent dips corresponding
1958: %to periods when the atom is temporarily heated. The important lesson
1959: %here is that if one wishes to use this cooling method to prepare
1960: %ground-state atoms for another purpose,
1961: %it is possible to prepare them in the
1962: %ground band (or equivalently, first-excited band) with higher
1963: %probability than would be suggested by the ensemble averages,
1964: %if care is taken to begin using the atoms at the proper time.
1965: %This could be done, for example, by making a separate, time-averaged
1966: %measurement of the photocurrent to monitor the average potential
1967: %energy, and thus the total energy, of the atoms in order to
1968: %determine when to ``accept'' the atoms as being prepared.
1969: %Even in the presence of of these population dips,
1970: %the atoms spend substantial blocks of
1971: %time with a population of well over 90\% in their respective
1972: %``ground'' states.
1973:
1974:
1975: \section{Simulations: Full Atom--Cavity Dynamics}\label{sim2}
1976:
1977: To verify the validity of the results of this paper, it is
1978: important to check the adiabatic approximation that leads to the
1979: reduced master equation (\ref{adiabaticSME}). We thus carried
1980: out simulations of the full, coupled atom--cavity dynamics described
1981: by Eq.~(\ref{scSME}).
1982:
1983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1984: % Small cavity test figure link inserted here.
1985: % This link uses the graphicx package.
1986: \begin{figure}[tb]
1987: \begin{center}
1988: \includegraphics[scale=0.41]{fig12.eps}
1989: \end{center}
1990: \vspace{-5mm}
1991: \caption
1992: {(Color online) Ensemble energy evolution comparing the effect of the spatial
1993: grid size (cavity mirror separation) on the atomic evolution.
1994: Heavy solid line (a): canonical parameters (adiabatic
1995: approximation) spanning 24 optical potential wells. Light solid
1996: line (b):
1997: 6 potential wells.
1998: Each curve represents an average over 128 trajectories.
1999: \label{fig:smcavtest}}
2000: \end{figure}
2001: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2002:
2003: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2004: % Full simulation figure link inserted here.
2005: % This link uses the graphicx package.
2006: \begin{figure}[tb]
2007: \begin{center}
2008: \includegraphics[scale=0.41]{fig13.eps}
2009: \end{center}
2010: \vspace{-5mm}
2011: \caption
2012: {(Color online) Ensemble energy evolution comparing evolution with and
2013: without invoking the adiabatic approximation.
2014: Heavy solid line (a): adiabatic approximation. Light solid
2015: line (b):
2016: full atom-cavity dynamics at 4 GHz detuning from atomic
2017: resonance. Dashed line: full atom-cavity dynamics at 2 GHz
2018: detuning from atomic resonance.
2019: Each curve represents an average over 128 trajectories.
2020: \label{fig:longrun}}
2021: \end{figure}
2022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2023:
2024: The simulation procedure is the same as before, except that we used
2025: the normalized Schr\"odinger equation corresponding to the
2026: master equation (\ref{scSME}),
2027: \begin{equation}
2028: \setlength{\arraycolsep}{0ex}
2029: \begin{array}{rcl}
2030: d\psiket &{}={}& \displaystyle -iH\psiket \,dt
2031: -\frac{\kapt}{2}a^\dagger a\psiket\,dt
2032: +\frac{\kapt}{2}\langle a+a^\dagger\rangle a\psiket\,dt\\&&
2033: \displaystyle
2034: -\frac{\kapt}{8}\langle a+a^\dagger\rangle^2\psiket\,dt %\\&&
2035: +\frac{\gamt}{2}\left(\langle \sigma^\dagger\sigma\rangle
2036: -\sigma^\dagger\sigma\right)\psiket\,dt\\&&
2037: \displaystyle
2038: +\sqrt{\kapt}\, a\psiket\,dW
2039: -\frac{\sqrt{\kapt}}{2}\langle a+a^\dagger\rangle\psiket\,dW
2040: \\&&\displaystyle
2041: +\left(\frac{\sigma e^{i\ut X}}
2042: {\sqrt{\langle\sigma^\dagger\sigma\rangle}}-1\right)
2043: \psiket\,dN.
2044: \end{array}
2045: \end{equation}
2046: We used the canonical parameter set described before.
2047: This set leads to
2048: scaled parameter values of $\gamt=190$ and $\kapt=1460$ in addition to
2049: those used in the adiabatic-approximation simulations as described in
2050: Section~\ref{sec:details}.
2051: The detuning
2052: $\Delta/2\pi = 4\;\mathrm{GHz}$ from atomic resonance implies a wide
2053: separation of timescales in the problem.
2054: Thus we again evolved the system over a large step of
2055: $\Delta t = 0.0005$ but in much smaller substeps of $\Delta t/4096$, applying the
2056: kinetic evolution operator after half of the substeps.
2057: The spatial grid contained 512 points, spanning only 6 wells to make
2058: the computation more tractable. The impact of the smaller grid on the
2059: cooling was to slightly accelerate the cooling process, as
2060: illustrated in Fig.~\ref{fig:smcavtest}. This effect due to an effective
2061: evaporation of hot atoms, since we treat the mirrors as absorbing
2062: boundary condition as discussed above.
2063: We also used 7 cavity states and 2 internal atomic states.
2064: Due to the small time step, we found it necessary to
2065: use 64-bit precision for these simulations.
2066:
2067: The simulation results are shown in Fig.~\ref{fig:longrun}, where we
2068: see that there is only a small difference between the cooling
2069: performance in the adiabatic and full simulations. Also shown is
2070: a simulation with a detuning $\Delta/2\pi = 2\;\mathrm{GHz}$, evolved
2071: with substeps of $\Delta t/2048$, to increase the nonadiabatic effects.
2072: The parameters were scaled to compensate. Even with this relatively
2073: close detuning, we see that the cooling dynamics are well
2074: described in the adiabatic approximation.
2075: Other ensemble-averaged quantities corroborate this point.
2076: For the 4 GHz detuning case, the atomic excited-state population
2077: $\langle \sigma^\dagger \sigma\rangle$ is
2078: $8.4\times 10^{-3}$ (corresponding to a spontaneous emission rate of
2079: 0.16); the adiabatic-approximation expression $\tilde{g}^2|\alpha|^2/(\tilde{\Delta}^2 +
2080: \tilde{\kappa}^2)$ predicts a similar value of $8.6\times 10^{-3}$ if we include a factor
2081: of $1-\langle E_\mathrm{eff}\rangle/2\Vmax = 0.98$ to account for the
2082: ensemble-averaged spatial distribution of the atoms.
2083: For the 2 GHz case, the excited-state population is $3.4\times
2084: 10^{-2}$ (spontaneous emission rate of 0.68),
2085: agreeing well with the adiabatic prediction of $3.5\times 10^{-2}$.
2086: In both cases, the cavity excitation $\langle a^\dagger a\rangle =
2087: 0.98$, in good agreement with the adiabatic value of 1.
2088:
2089: %longrun pop: 0.00084 expect: 0.00086 SE rate: 0.16 <ada> 0.98 (expect 1)
2090: %longrunn pop: 0.0034 expect: 0.0035 SE rate: 0.68 <ada> 0.98 (expect 1)
2091:
2092:
2093:
2094: \section{Conclusion}\label{conc}
2095: Real-time quantum feedback control is worth investigating not only
2096: because of potential future applications, but because it will provide
2097: an undeniable litmus test of quantum measurement theory at the level
2098: of individual measurement records produced by measuring single
2099: systems. In this work we have shown that, by using a simplified
2100: estimation technique coupled with a relatively simple feedback algorithm,
2101: a single atom may be cooled close to its ground state in a realistic
2102: optical cavity. We have also explored the effectiveness of the feedback
2103: algorithm in various parameter regimes and we hope that this will help
2104: to guide future experiments.
2105:
2106: \section*{Acknowledgments}
2107: The authors would like to thank Andrew Doherty
2108: and Sze Tan for helpful discussions.
2109: This research was performed in
2110: part using the resources of the Advanced Computing Laboratory,
2111: Institutional Computing Initiative, and LDRD program
2112: of Los Alamos National Laboratory.
2113:
2114: \appendix
2115:
2116: \section{Simple Feedback Cooling and Squeezing}
2117:
2118: Here we analyze the effect of the simple centroid-only feedback cooling algorithm
2119: described in Section~\ref{section:alg}.
2120: This involves switching the optical potential to a high value when
2121: the atom is climbing a potential well, and switching it to a low value
2122: when it is falling.
2123: Recall that the value of the ``switched high'' field is
2124: $(1 + \varepsilon_1) E$ and the ``switched low'' field is
2125: $(1 - \varepsilon_2) E$, so that the
2126: potential height has the corresponding values
2127: $(1 + \varepsilon_1)^2 V$ and $(1 - \varepsilon_2)^2 V$ ($E$ and $V$
2128: are the respective unmodulated values).
2129:
2130: Since the feedback algorithm involves more than
2131: simply driving the system with an external force, %and since the
2132: %dynamics of the system are fairly complex, we cannot expect to be
2133: %able to predict the behavior accurately without the use of full
2134: %simulations. Nevertheless,
2135: to understand the action of the algorithm one must analyze its
2136: effect both on the means (the centroid) and the variances of the atomic position and
2137: momentum. The effect on the centroid is to damp the motion. In one cooling
2138: cycle, the energy of the centroid is multiplied by a ``cooling
2139: factor,''
2140: \begin{equation}
2141: C_{\mbox{\scriptsize fb}} = \left(\frac{1 - \varepsilon_2}{1 +
2142: \varepsilon_1}\right)^2,
2143: \end{equation}
2144: which is valid in the harmonic approximation.
2145:
2146: The effect on the variances is, however, to squeeze the atomic wave function.
2147: To understand why this is the case, consider what happens when the
2148: height of the potential is changed. This changes the frequency of the
2149: effective harmonic oscillator governing the atomic motion, and as a
2150: consequence, a wave function that was not squeezed in the phase space
2151: of the old harmonic oscillator, is squeezed in the space of the new
2152: one. In particular, raising the height of the potential scales phase
2153: space so that momentum is squeezed, and conversely lowering the
2154: potential squeezes position. To determine the rate of squeezing, it
2155: is useful to consider what happens in one ``cooling cycle.'' For
2156: example, we may define the cooling cycle to start when the atom
2157: crosses the bottom of the well, and end when the atom again crosses
2158: the bottom of the well traveling in the original direction
2159: (essentially one period of oscillation). Examining the effect of each
2160: of the changes in the potential during a cooling cycle, we find that
2161: the result of a single cycle is to multiply the momentum variance of
2162: the atomic wave function by a ``magnification factor,''
2163: \begin{equation}
2164: M_\mathrm{fb} = \left(\frac{1+\varepsilon_1}{1 -
2165: \varepsilon_2}\right)^2,
2166: \end{equation}
2167: in the harmonic approximation. This factor is precisely the
2168: inverse of the cooling factor $C_\mathrm{fb}$.
2169:
2170: This situation is analogous to an attempt to cool a classical
2171: ensemble of particles using a control algorithm designed for a single
2172: particle; although the algorithm will cool a single particle, it
2173: will almost certainly heat the other members of the
2174: ensemble. This will be an important consideration in designing an
2175: improved feedback algorithm. A simulation of this simple feedback
2176: algorithm is given in Figure~\ref{fig:basiccooling}. This shows that after
2177: a few cooling cycles in which the total energy is initially reduced, the
2178: resulting squeezing overtakes the cooling of the centroid, and the
2179: algorithm actually heats the atomic motion.
2180:
2181:
2182: \section{Heating Mechanisms}\label{heating}
2183:
2184: \noindent\textsl{\bfseries Spontaneous Emission.}
2185: Spontaneous emission events kick the atom in both directions, and
2186: therefore cause momentum diffusion. The effect of this is
2187: to increase the energy, on average, linearly with time.
2188: This rate depends both upon the
2189: spontaneous-emission rate as well as the magnitude of the associated kicks.
2190: One point to note is that the magnitude of each
2191: kick is not merely determined by the momentum of the emitted photon,
2192: but also by the difference between the momentum of the excited and
2193: ground state wave functions.
2194: More precisely, we can factor the atomic state vector \ket\psi\
2195: just before a spontaneous-emission event
2196: into a product of internal and motional states,
2197: \begin{equation}
2198: \ket\psi = \ket{\psi_\mathrm{e}}\ket{\mathrm{e}} +
2199: \ket{\psi_\mathrm{g}}\ket{\mathrm{g}},
2200: \end{equation}
2201: where \ket{\psi_\mathrm{\alpha}}\ are states in the center-of-mass
2202: space of the atom, and ``e'' and ``g'' denote excited and ground atomic
2203: energy levels, respectively. Then in an unraveling of the
2204: master equation (\ref{SME}) where the spontaneously emitted photons
2205: are detected but do not give position information about the atom,
2206: the one-dimensional atomic state just after a
2207: spontaneous-emission event is just the previous state with the
2208: internal state lowered and a momentum-recoil factor,
2209: \begin{equation}
2210: \ket\psi = \ket{\psi_\mathrm{e}}\ket{\mathrm{g}}e^{-i\tilde u X},
2211: \end{equation}
2212: where the random variable $\tilde u$ is chosen from the projected
2213: angular distribution of the resonance fluorescence $N(\tilde u)$
2214: referred to in the SME (\ref{scSME}).
2215: Thus, in addition to the emission momentum recoil, the
2216: spontaneous emission effectively converts the atomic state from the ground
2217: state wave function to the excited state wave function. In the
2218: regime where the adiabatic approximation is valid, we have the
2219: relation
2220: \begin{equation}
2221: \ket{\psi_\mathrm{e}} = \frac{\alpha g}{\Delta}\cos(kx)
2222: \ket{\psi_\mathrm{g}},
2223: \end{equation}
2224: and thus we can interpret this extra
2225: process as being simply another momentum kick due to the absorption of a cavity
2226: photon (i.e., a superposition of a photon-recoil kick in either
2227: direction along the cavity axis).
2228:
2229: The actual rate of spontaneous emissions may be estimated by multiplying the
2230: spontaneous emission rate by the average excited state population,
2231: giving $\tilde{\gamma}\tilde{g}^2|\alpha|^2/(\tilde{\Delta}^2 +
2232: \tilde{\kappa}^2)$ for an atom localized at a field antinode.
2233: Calculating this for the conditions
2234: that we use in the numerical simulations in the body of the text, we find that the
2235: rate is around 0.17. In the far-detuned regime that we study here,
2236: the spontaneous-emission heating is a minor effect compared to
2237: the other heating mechanisms.
2238:
2239: \noindent\textsl{\bfseries Measurement Backaction.}
2240: The rate of heating from the back-action due to the continuous
2241: measurement process may be estimated by examining the effective
2242: master equation for the system in
2243: %the case that the cavity mode and
2244: %the internal atomic states have been adiabatically eliminated, as in
2245: Eq.~(\ref{adiabaticSME}).
2246: The energy of the atomic motion in the effective
2247: potential, defined with respect to the minimum
2248: potential energy, is simply
2249: \begin{equation}
2250: E_\mathrm{eff} = \pi P^2 +
2251: \frac{\pi}{\kt^2}\left[1-\cos^2(\tilde{k}X)\right] .
2252: \end{equation}
2253: The rate of increase of this energy, averaged over all
2254: possible trajectories, due to the
2255: measurement is determined by the term in the SME
2256: proportional to $\Gamma$, which gives
2257: \begin{equation}
2258: \setlength{\arraycolsep}{0ex}
2259: \renewcommand{\arraystretch}{1.8}
2260: \begin{array}{rcl}
2261: \partial_t\langle {E}_\mathrm{eff}\rangle_\mathrm{meas} &{}={}& 2\Gamma
2262: \mathrm{Tr}\{E_\mathrm{eff}{\cal D}[\cos^2(\kt x)]\rho\} \\
2263: &{}={}& 8\pi\Gamma \langle \cos^2(\kt X)\sin^2(\kt X)\rangle .
2264: \end{array}
2265: \end{equation}
2266: Note that the $\Gamma$ here is actually a time-averaged quantity when
2267: feedback is applied, which
2268: we assume to be well approximated by the unmodulated value of $\Gamma$
2269: in Eq.~(\ref{Gammadef}), at least to $O(\varepsilon)$.
2270: The simplest limit for this expression is if we assume the particle
2271: to be randomly distributed along the potential,
2272: which is a reasonably good approximation to the initial heating rate:
2273: \begin{equation}
2274: \partial_t\langle {E}_\mathrm{eff}\rangle_\mathrm{meas} = \pi\Gamma\kt^2 .
2275: \label{uniformheating}
2276: \end{equation}
2277: A more appropriate limit for this heating rate after cooling has taken
2278: place is the
2279: low-temperature regime, where we can make the harmonic approximation
2280: so that the rate becomes
2281: \begin{equation}
2282: \partial_t\langle {E}_\mathrm{eff}\rangle_\mathrm{meas}
2283: \approx 8\Gamma\kt^2 \langle \kt^2 X^2\rangle
2284: \approx 4\Gamma\kt^4
2285: \langle E_\mathrm{eff}\rangle,
2286: \label{measurementheating}
2287: \end{equation}
2288: which is valid for small $\langle E_\mathrm{eff}\rangle$.
2289:
2290: Note that although we attribute this heating to measurement
2291: back-action, an equivalent, measurement-independent picture for this
2292: heating process is the cavity decay process. Because light escapes the
2293: cavity in discrete units at random times, there is a stochastic
2294: component to the cavity field intensity. Thus, there is a
2295: stochastic, heating force on the atoms.
2296: The rate obtained by this argument must be the same as the
2297: rate that we just obtained, since the heating rate obtained by
2298: averaging over all possible trajectories in a ``quantum jump''
2299: unraveling of the master equation must be the same as in the ``quantum
2300: diffusion'' unraveling that we use here.
2301:
2302: \section{Effects of Detection Inefficiency and Feedback Delays}
2303:
2304: \subsection{Detection Efficiency}
2305:
2306: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2307: % Eta sweep figure link inserted here.
2308: % This link uses the graphicx package.
2309: \begin{figure}[tb]
2310: \begin{center}
2311: \includegraphics[scale=0.41]{fig14.eps}
2312: \end{center}
2313: \vspace{-5mm}
2314: \caption
2315: {(Color online) Final energies, measured over the interval from $t=90$
2316: to $100$, as the detection efficiency $\eta$
2317: varies; other parameters match the canonical set.
2318: Cooling is according to the improved algorithm based
2319: on the Gaussian estimator; error bars reflect standard
2320: errors from averages over 128 trajectories.
2321: \label{fig:etasweep}}
2322: \end{figure}
2323: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2324:
2325: Now we come to one of the two primary limitations in an experimental
2326: implementation of the present system, that of limited efficiency
2327: of the optical detectors. As this limits the information incorporated
2328: by the estimator, it is natural to expect that the best cooling
2329: is achieved for unit efficiency, and that the cooling simply gets
2330: worse for inefficient measurements. Simulated final energies
2331: using the Gaussian estimator are shown in Fig.~\ref{fig:etasweep},
2332: which for the most part bear out this expectation. One remarkable
2333: feature of these results is that the cooling is extremely robust
2334: to detector inefficiency. It is only down around 50\% detection efficiency
2335: that a few trajectories begin to pull up the average energies,
2336: and below about 35\% efficiency that there is a clear trend
2337: towards large steady-state temperatures. This robustness is
2338: very encouraging that an experimental demonstration of
2339: quantum feedback cooling will work well in spite of imperfect
2340: detectors.
2341:
2342:
2343: \subsection{Speed and Feedback Delay Issues}\label{section:speed}
2344:
2345: In the simulations we have discussed so far, the Gaussian estimator
2346: was updated every time interval of $\DtG = 0.0005$. For the
2347: canonical parameter set, this update interval corresponds to about 3
2348: ns in physical time. This is obviously a very short time in which to
2349: evolve the estimation equations (\ref{gest}) as well as the curve fit
2350: outlined in Section~\ref{section:alg}.
2351: Therefore, the first issue we will tackle is the speed with which
2352: one can iterate the control algorithm.
2353:
2354: \noindent\textsl{\bfseries Curve Fit.}
2355: Recall that, given a set of estimates $y_{\mathrm{est}, n}$
2356: corresponding to
2357: times $t_n$, we wish to implement the curve fit of the
2358: function $a_0+a_1 x + a_2 x^2$ to the last $q$ values of
2359: $y_{\mathrm{est}, i}$. The curve-fit coefficients are
2360: given by the solution of the normal equations,
2361: \begin{equation}
2362: \left(
2363: \begin{array}{ccc}
2364: \Sigma_0 & \Sigma_1 & \Sigma_2 \\
2365: \Sigma_1 & \Sigma_2 & \Sigma_3 \\
2366: \Sigma_2 & \Sigma_3 & \Sigma_4
2367: \end{array}
2368: \right)
2369: \left(
2370: \begin{array}{c}
2371: a_0 \\ a_1 \\ a_2
2372: \end{array}
2373: \right) =
2374: \left(
2375: \begin{array}{c}
2376: S_0 \\ S_1 \\ S_2
2377: \end{array}
2378: \right) ,
2379: \end{equation}
2380: where $\Sigma_n := \sum_{j=1}^{q} (x_j)^n$ and
2381: $S_n := \sum_{j=1}^{q} y_j(x_j)^n$.
2382: However, recomputing the sums in this system of equations
2383: at each iteration is not computationally efficient.
2384: To make the on-the-fly
2385: calculation of this curve fit feasible, we can instead
2386: implement the coupled recurrence relations
2387: \begin{equation}
2388: \setlength{\arraycolsep}{0ex}
2389: \renewcommand{\arraystretch}{1.4}
2390: \begin{array}{rcl}
2391: S_0^{(n)} &{}={}& S_0^{(n-1)} + y_{\mathrm{est}, n} +
2392: y_{\mathrm{est}, n-m}\\
2393: S_1^{(n)} &{}={}& S_1^{(n-1)} - S_0^{(n-1)} +
2394: m y_{\mathrm{est}, n-m}\\
2395: S_2^{(n)} &{}={}& S_2^{(n-1)} -2 S_1^{(n-1)} + S_0^{(n-1)} -
2396: m^2 y_{\mathrm{est}, n-m}
2397: \end{array}
2398: \end{equation}
2399: and then compute the fitted slope coefficient at time $t_n$ according to
2400: \begin{equation}
2401: \begin{array}{l}
2402: \displaystyle
2403: a_1 = \frac{18(2q-1)}{q(q+1)(q+2)}\\
2404: \hspace{5mm}\displaystyle\times\left(
2405: S_0 + \frac{2(8q-11)}{3(q-1)(q-2)}S_1 +
2406: \frac{10}{(q-1)(2q-1)}S_2\right).
2407: \end{array}
2408: \end{equation}
2409: For notational convenience, we have defined the times
2410: corresponding to the last $q$ measurements to be
2411: $t_n = 1-n$, so that the fitted slope at the current time
2412: is simply $a_1$. Thus, we can simply trigger the feedback
2413: cooling algorithm on the sign of $a_1$, switching the potential
2414: high or low for a positive or negative sign of $a_1$, respectively.
2415: The speed of this algorithm is largely independent of the number of
2416: samples used in the curve fit, assuming that the table of the
2417: last $q$ values can be accessed quickly (i.e., it can fit into
2418: fast memory).
2419:
2420: Later we will examine the effects of computation and signal propagation
2421: delays on the cooling performance. Such delays are easily accounted
2422: for in the present algorithm by using the curve fit to extrapolate
2423: forward in time.
2424: To do this, we require the quadratic fit coefficient,
2425: \begin{equation}
2426: \begin{array}{l}
2427: \displaystyle
2428: a_2 = \frac{30}{q(q+1)(q+2)}\\
2429: \hspace{5mm}\displaystyle\times\left(
2430: S_0 + \frac{6}{q-2}S_1 +
2431: \frac{6}{(q-1)(q-2)}S_2\right),
2432: \end{array}
2433: \end{equation}
2434: so that we trigger on the value of the current-time slope,
2435: \begin{equation}
2436: a_1 + 2 a_2 d,
2437: \end{equation}
2438: where $d$ is the delay in multiples of $\DtG$ from the time of the
2439: most recent data used in the fit to the current time.
2440: Due to the predictive nature of
2441: this fitting scheme, we expect that the cooling should be robust
2442: to small delays.
2443:
2444:
2445: \noindent\textsl{\bfseries Gaussian Estimator.}
2446: The other issue to be resolved is to maximize the efficiency of
2447: updating the Gaussian estimator. One computationally expensive
2448: aspect of the form of the equations (\ref{gest}) is the need to
2449: evaluate transcendental functions. A substantial speed improvement
2450: results from introducing the variables
2451: $x_1 := \sin(2\kt\Xe)$,
2452: $x_2 := \cos(2\kt\Xe)$,
2453: $x_3 := \Pe/\kt$,
2454: $x_4 := 2\kt^2\Vxe$,
2455: $x_5 := \exp(-2\kt^2\Vxe)$,
2456: $x_6 := \Vpe/\kt^2$, and
2457: $x_7 := \exp(-2\kt^2\Vxe)$,
2458: we can rewrite the estimator equations in the form
2459: \begin{equation}
2460: \setlength{\arraycolsep}{0ex}
2461: \renewcommand{\arraystretch}{1.4}
2462: \begin{array}{rcl}
2463: dx_1 &{}={}& (4\pi \kt^2 x_2 x_3 -4\eta\Gamma x_1^{\,3} x_4^{\,2} x_5^{\,2})dt\\
2464: && + \sqrt{8\eta\Gamma} x_1 x_2 x_4 x_5\dWe
2465: \\
2466: dx_2 &{}={}& -(4\pi \kt^2 x_1 x_3 +4\eta\Gamma x_1^{\,2}x_2 x_4^{\,2} x_5^{\,2})dt\\
2467: && - \sqrt{8\eta\Gamma} x_1^{\,2} x_4 x_5\dWe
2468: \\
2469: dx_3 &{}={}& -\Vmax x_1 x_5 dt + \sqrt{8\eta\Gamma}x_1x_5x_7\dWe
2470: \\
2471: dx_4 &{}={}& (8\pi \kt^2 x_7 -4\eta\Gamma x_1^{\,2} x_4^{\,2} x_5^{\,2})dt\\
2472: && + \sqrt{8\eta\Gamma} x_2x_4^{\,2} x_5\dWe
2473: \\
2474: dx_5 &{}={}& [-8\pi \kt^2 x_5x_7 +4\eta\Gamma
2475: (x_1^{\,2}x_5-x_2^{\,2}x_4^{\,2}) x_4^{\,2} x_5^{\,2}]dt\\
2476: && - \sqrt{8\eta\Gamma} x_2x_4^{\,2} x_5^{\,2}\dWe
2477: \\
2478: dx_6 &{}={}& (-4\Vmax x_2x_5x_7 +\Gamma[1+x_5^{\,4}(1-2x_2^{\,2})]\\
2479: &&-8\eta\Gamma x_1^{\,2}x_5^{\,2}x_7^{\,2})dt\\
2480: && - \sqrt{2\eta\Gamma}[1-2( x_4+2x_7^{\,2}) x_5] x_2\dWe
2481: \\
2482: dx_7 &{}={}& (2\pi \kt^2 x_6 -\Vmax x_2 x_4x_5
2483: -4\eta\Gamma x_1^{\,2} x_4 x_5^{\,2} x_7)dt\\
2484: && + \sqrt{8\eta\Gamma} x_2 x_4 x_5 x_7\dWe ,
2485: \\
2486: \end{array}
2487: \label{fastgest}
2488: \end{equation}
2489: with
2490: \begin{equation}
2491: \dWe = \frac{dr}{\sqrt{\eta\kapt}} + \sqrt{2\eta\Gamma}(1+x_2 x_5) dt ,
2492: \end{equation}
2493: thus obviating any need for transcendental-function evaluation.
2494: However, it turns out that the evolution of these equations via
2495: the Euler method is much less stable than the evolution of the
2496: original equations. This is due to the fact that there are pairs of
2497: variables [$(x_1, x_2)$ and $(x_4, x_5)$] that evolve separately
2498: but must satisfy consistency conditions [$x_1^2+x_2^2 = 1$ and
2499: $x_5 = \exp(-x_4)$]; also a large stochastic fluctuation can
2500: push these values into unphysical ranges (e.g., $x_1>1$).
2501: A better strategy is to emulate the results of an Euler solution
2502: to the original estimator equations, and hence
2503: update $x_1$, $x_2$, and $x_5$ using the increments
2504: \begin{equation}
2505: \setlength{\arraycolsep}{0ex}
2506: \renewcommand{\arraystretch}{1.4}
2507: \begin{array}{rcl}
2508: \Delta x_1 &{}={}& x_1[\cos(\twoktDeltaXe) - 1] + x_2\sin(\twoktDeltaXe)
2509: \\
2510: \Delta x_2 &{}={}& x_2[\cos(\twoktDeltaXe) - 1] - x_1\sin(\twoktDeltaXe)
2511: \\
2512: \Delta x_5 &{}={}& x_5\exp(-\Delta x_4),
2513: \end{array}
2514: \label{fastgestmod}
2515: \end{equation}
2516: where
2517: \begin{equation}
2518: \twoktDeltaXe = 4\pi\kt^2 x_3\DtG + \sqrt{8\eta\Gamma}x_1x_4x_5
2519: \Delta W_\mathrm{e},
2520: \end{equation}
2521: $\Delta W_\mathrm{e}(t) := \int_{t}^{t+\DtG}\dWe$,
2522: and $\Delta x_4$ is the Euler-update increment from
2523: Eqs.~(\ref{fastgest}). These equations can then be expanded in powers
2524: of $2\kt\Delta\Xe$ to avoid computing the transcendental functions.
2525: A final trick to help stabilize the equations is to multiply the
2526: $x_1$ and $x_2$ variables at each time step by the normalization quantity
2527: $[1 - (x_1^{\,2} + x_2^{\,2})]/2$, which explicitly maintains the
2528: identity $x_1^2+x_2^2 = 1$ without requiring the evaluation of a
2529: square root.
2530:
2531: \noindent\textsl{\bfseries Execution Times.}
2532: To evaluate how quickly the control algorithm can be iterated, we
2533: implemented the control algorithm on a modern, general-purpose, serial
2534: microprocessor (1.25 GHz Alpha EV68). A single iteration of the
2535: estimator according to Eqs.~(\ref{gest}) and a curve fit, including
2536: all steps necessary to compute the control output, can be completed in
2537: about 260 ns. Switching to the faster method of Eqs.~(\ref{fastgest})
2538: with the more stable updates (\ref{fastgestmod}) can be completed
2539: in 140 ns with a truncation at
2540: fourth order in $\twoktDeltaXe$, and 120 ns with a second-order
2541: truncation (including a normalization
2542: step). The cooling performance is essentially the same with all
2543: these algorithms, in that they all reach the final temperature;
2544: the exception is an algorithm that uses Eqs.~(\ref{fastgest})
2545: without the modifications of Eqs.~(\ref{fastgestmod}), which has
2546: substantially worse cooling performance. The tradeoff in
2547: the truncated-expansion algorithms
2548: is that lower-order truncations are faster but also
2549: more unstable and thus require more resets of the estimator.
2550:
2551: % (about 56\% of trajectories
2552: %needed to be reset in the fourth-order case, and about 75\% in the
2553: %second-order case, compared to about 25\% using the original equations).
2554:
2555: Based on this performance,
2556: and assuming Moore's law, general-purpose hardware will not
2557: be able to iterate this algorithm in 3 ns for about a decade.
2558: However, even presently available, specialized technologies
2559: such as multicore processors and field-programmable gate arrays
2560: (FPGAs) can likely be used to significantly decrease the execution time.
2561: For example, because the curve fit can be moved onto a separate
2562: logical processor, the estimator can be updated more quickly
2563: (84 ns for the fourth-order version, 70 ns for the second-order
2564: version); the extra 60 ns required to generate the feedback
2565: value via the curve fit then represents an additional delay, but
2566: does not limit the rate at which the estimator can be iterated.
2567: Further parallelization should bring these times down
2568: substantially.
2569:
2570: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2571: % Delay sweep figure link inserted here.
2572: % This link uses the graphicx package.
2573: \begin{figure}[tb]
2574: \begin{center}
2575: \includegraphics[scale=0.41]{fig15.eps}
2576: \end{center}
2577: \vspace{-5mm}
2578: \caption
2579: {(Color online) Final energies, measured over the interval from $t=90$
2580: to $100$, as a function of the feedback delay $\taud$.
2581: (Time is measured in units of the 5.8~$\mu$s
2582: harmonic-oscillator period.)
2583: Cooling is according to the improved algorithm based
2584: on the Gaussian estimator, with extrapolation implemented
2585: in the curve fit; error bars reflect standard
2586: errors from averages over 128 trajectories.
2587: Other parameters are as in the canonical set.
2588: \label{fig:delaysweep}}
2589: \end{figure}
2590: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2591:
2592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2593: % Update sweep figure link inserted here.
2594: % This link uses the graphicx package.
2595: \begin{figure}[tb]
2596: \begin{center}
2597: \includegraphics[scale=0.41]{fig16.eps}
2598: \end{center}
2599: \vspace{-5mm}
2600: \caption
2601: {(Color online) Final energies, measured over the interval from $t=90$
2602: to $100$, as a function of the Gaussian estimator
2603: time step $\DtG$.
2604: (Time is measured in units of the 5.8~$\mu$s
2605: harmonic-oscillator period.)
2606: Shown are results obtained by
2607: reducing the number of samples to keep the time interval
2608: for the quadratic curve constant (circles) as well as
2609: results obtained via the multiple-estimator staggering
2610: method to maintain a 300-point fit (triangles).
2611: Error bars reflect standard
2612: errors from averages over 128 trajectories.
2613: Other parameters are as in the canonical set.
2614: \label{fig:updatesweep}}
2615: \end{figure}
2616: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2617:
2618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2619: % Realistic conditions sweep figure link inserted here.
2620: % This link uses the graphicx package.
2621: \begin{figure}[tb]
2622: \begin{center}
2623: \includegraphics[scale=0.41]{fig17.eps}
2624: \end{center}
2625: \vspace{-5mm}
2626: \caption
2627: {(Color online) Ensemble energy evolution illustrating the combined effects of
2628: detection efficiency $\eta$, feedback delay $\taud$,
2629: and estimator time step $\DtG$.
2630: (Time is measured in units of the 5.8~$\mu$s
2631: harmonic-oscillator period.)
2632: These parameter
2633: values for the curves (a)-(d)
2634: are given in the text; others are as in the
2635: canonical set. Extrapolation is implemented in the
2636: curve fit, and multiple-estimator staggering is used
2637: to maintain a 300-point fit.
2638: Each curve represents an average over 128 trajectories.
2639: \label{fig:realistic}}
2640: \end{figure}
2641: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2642:
2643: \noindent\textsl{\bfseries Impact on Cooling Performance.}
2644: Given these limitations in computing power it is necessary to
2645: evaluate how quickly the algorithm needs to be iterated in order
2646: to achieve good cooling performance. The cooling performance
2647: in the presence of feedback delay $\taud=d\DtG$ is shown in
2648: Fig.~\ref{fig:delaysweep}. The cooling is very robust to the
2649: delay, since we can compensate by extrapolating the curve fit.
2650: The cooling behavior is essentially unaffected out to a delay
2651: of about 460 ns, which is well above the 60 ns figure that we
2652: mentioned for generating the feedback value, and thus additional
2653: signal propagation and processing delays should not be too problematic.
2654:
2655: Limitations on the time $\DtG$ required to iterate the Gaussian
2656: estimator can also seriously impact the cooling performance.
2657: Not only does a large $\DtG$ step reduce the quality of the
2658: estimated solution, but in fitting the quadratic curve over a
2659: finite interval, there are fewer points involved in the fit,
2660: reducing the effectiveness of the fit's noise reduction.
2661: These effects on the cooling behavior are illustrated in
2662: Fig.~\ref{fig:updatesweep}, which shows that the late-time ensemble
2663: energies grow approximately linearly with $\DtG$.
2664: Here, the 70 ns figure mentioned above for $\DtG$ yields a
2665: final energy of around 25, compared to around 9 for the smallest
2666: value of $\DtG=0.0005$ (3 ns) shown, a substantial decrease in
2667: the cooling effectiveness.
2668: One possible way to address this problem is to use multiple
2669: Gaussian estimators that are staggered in time. For example,
2670: defining $\Delta t_\mathrm{min}=0.0005$, and given that the
2671: computational hardware is limited to $\DtG = N\Delta t_\mathrm{min}$
2672: for some integer $N$, we can arrange to have $N$ independent
2673: implementations of the Gaussian estimator executing in parallel.
2674: They all receive the same measurement record as input, but they
2675: are staggered in time such that one produces output at each
2676: small time step $\Delta t_\mathrm{min}=0.0005$. In this way,
2677: each of the individual solutions is of lower quality than
2678: a fine-time-step solution, but the curve fit can be fit to
2679: all of them (restoring a full 300 point curve fit),
2680: thus helping to stabilize the cooling algorithm. The implementation
2681: of this method is also displayed in Fig.~\ref{fig:updatesweep},
2682: which shows that this method restores much of the
2683: quality of the cooling that was lost in using the simpler
2684: algorithm. The rightmost point in this figure corresponds
2685: to operating 55 parallel estimators.
2686:
2687: So far, we have studied the performance impact on cooling of
2688: each of the three equipment-limitation effects (detection
2689: efficiency, feedback delay, and estimator time step) separately.
2690: Some representative examples of their combined effect are shown
2691: in Fig.~\ref{fig:realistic}. The four curves, in order of
2692: increasing departure from the ideal case, correspond to the
2693: following parameters:
2694: (a) $\eta=1$, $\taud=0$, $\DtG=\Delta t_\mathrm{min}=0.0005$ (i.e., 3 ns in physical time); %basic_coolingr
2695: (b) $\eta=0.8$, $\taud=\DtG=10\Delta t_\mathrm{min}=0.005$ (i.e., 29 ns); %realistic3r
2696: (c) $\eta=0.7$, $\taud=\DtG=25\Delta t_\mathrm{min}=0.0125$ (i.e., 72 ns); %realistic1rr.
2697: (d) $\eta=0.5$, $\taud=\DtG=40\Delta t_\mathrm{min}=0.02$ (i.e., 116 ns); %realistic2r.
2698: (e) $\eta=0.3$, $\taud=\DtG=50\Delta t_\mathrm{min}=0.025$ (i.e., 145 ns). %realistic4r.
2699: The sums of the populations in the lowest two bands for these cases,
2700: computed between $t=90$ and $100$, are
2701: $(94\pm 5)\%$, %basic_cooling.param
2702: $(84\pm 5)\%$, %realistic3.param
2703: $(45\pm 4)\%$, %realistic1rr.param
2704: $(47\pm 4)\%$, and %realistic2r.param
2705: $(13\pm 2)\%$, %realistic4r.param
2706: respectively. The cooling performance, even for the case that should
2707: be within reach of current technology (d), is not unreasonable in
2708: comparison to the ideal case (a), although cooling quickly becomes
2709: worse as these parameter values are relaxed (e). Even with these
2710: effects, this method is certainly suitable for long-time storage of
2711: atoms in the optical potential and can still reach the ground
2712: state with serviceable efficiency.
2713:
2714:
2715:
2716:
2717: \begin{thebibliography}{74}
2718: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2719: \expandafter\ifx\csname bibnamefont\endcsname\relax
2720: \def\bibnamefont#1{#1}\fi
2721: \expandafter\ifx\csname bibfnamefont\endcsname\relax
2722: \def\bibfnamefont#1{#1}\fi
2723: \expandafter\ifx\csname citenamefont\endcsname\relax
2724: \def\citenamefont#1{#1}\fi
2725: \expandafter\ifx\csname url\endcsname\relax
2726: \def\url#1{\texttt{#1}}\fi
2727: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
2728: \providecommand{\bibinfo}[2]{#2}
2729: \providecommand{\eprint}[2][]{\url{#2}}
2730:
2731: \bibitem[{\citenamefont{Wheeler and Zurek}(1983)}]{Wheeler83}
2732: \bibinfo{editor}{\bibfnamefont{J.~A.} \bibnamefont{Wheeler}} \bibnamefont{and}
2733: \bibinfo{editor}{\bibfnamefont{W.~H.} \bibnamefont{Zurek}}, eds.,
2734: \emph{\bibinfo{title}{Quantum Theory and Measurement}}
2735: (\bibinfo{publisher}{Princeton University Press},
2736: \bibinfo{address}{Princeton}, \bibinfo{year}{1983}).
2737:
2738: \bibitem[{\citenamefont{Milburn et~al.}(1994)\citenamefont{Milburn, Jacobs, and
2739: Walls}}]{Milburn94}
2740: \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Milburn}},
2741: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}}, \bibnamefont{and}
2742: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}},
2743: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{50}},
2744: \bibinfo{pages}{5256} (\bibinfo{year}{1994}).
2745:
2746: \bibitem[{\citenamefont{Mabuchi et~al.}(1999)\citenamefont{Mabuchi, Ye, and
2747: Kimble}}]{Mabuchi99}
2748: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2749: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ye}}, \bibnamefont{and}
2750: \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Kimble}},
2751: \bibinfo{journal}{Appl.\ Phys.\ B} \textbf{\bibinfo{volume}{68}},
2752: \bibinfo{pages}{1095} (\bibinfo{year}{1999}).
2753:
2754: \bibitem[{\citenamefont{M{\"u}nstermann
2755: et~al.}(1999)\citenamefont{M{\"u}nstermann, Fischer, Maunz, Pinkse, and
2756: Rempe}}]{Muenstermann99}
2757: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{M{\"u}nstermann}},
2758: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Fischer}},
2759: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Maunz}},
2760: \bibinfo{author}{\bibfnamefont{P.~W.~H.} \bibnamefont{Pinkse}},
2761: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rempe}},
2762: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{82}},
2763: \bibinfo{pages}{3791} (\bibinfo{year}{1999}).
2764:
2765: \bibitem[{\citenamefont{Hood et~al.}(2000)\citenamefont{Hood, Lynn, Doherty,
2766: Parkins, and Kimble}}]{Hood00}
2767: \bibinfo{author}{\bibfnamefont{C.~J.} \bibnamefont{Hood}},
2768: \bibinfo{author}{\bibfnamefont{T.~W.} \bibnamefont{Lynn}},
2769: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}},
2770: \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{Parkins}},
2771: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~J.}
2772: \bibnamefont{Kimble}}, \bibinfo{journal}{Science}
2773: \textbf{\bibinfo{volume}{287}}, \bibinfo{pages}{1447} (\bibinfo{year}{2000}).
2774:
2775: \bibitem[{\citenamefont{Pinkse et~al.}(2000)\citenamefont{Pinkse, Fischer,
2776: Maunz, and Rempe}}]{Pinkse00}
2777: \bibinfo{author}{\bibfnamefont{P.~W.~H.} \bibnamefont{Pinkse}},
2778: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Fischer}},
2779: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Maunz}}, \bibnamefont{and}
2780: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rempe}},
2781: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{404}},
2782: \bibinfo{pages}{365} (\bibinfo{year}{2000}).
2783:
2784: \bibitem[{\citenamefont{Belavkin}(1983)}]{Belavkin83}
2785: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}},
2786: \bibinfo{journal}{Automatica and Remote Control}
2787: \textbf{\bibinfo{volume}{44}}, \bibinfo{pages}{178} (\bibinfo{year}{1983}).
2788:
2789: \bibitem[{\citenamefont{Belavkin}(1987)}]{Belavkin87}
2790: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}}, in
2791: \emph{\bibinfo{booktitle}{Information Complexity and Control in Quantum
2792: Physics}}, edited by
2793: \bibinfo{editor}{\bibfnamefont{A.}~\bibnamefont{Blaqui{\`e}re}},
2794: \bibinfo{editor}{\bibfnamefont{S.}~\bibnamefont{Diner}}, \bibnamefont{and}
2795: \bibinfo{editor}{\bibfnamefont{G.}~\bibnamefont{Lochak}}
2796: (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{New York},
2797: \bibinfo{year}{1987}).
2798:
2799: \bibitem[{\citenamefont{Sauer et~al.}(2004)\citenamefont{Sauer, Fortier, Chang,
2800: Hamley, and Chapman}}]{Sauer04}
2801: \bibinfo{author}{\bibfnamefont{J.~A.} \bibnamefont{Sauer}},
2802: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Fortier}},
2803: \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Chang}},
2804: \bibinfo{author}{\bibfnamefont{C.~D.} \bibnamefont{Hamley}},
2805: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~S.}
2806: \bibnamefont{Chapman}}, \bibinfo{journal}{{Phys.\ Rev.\ A}}
2807: \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{051804(R)}
2808: (\bibinfo{year}{2004}).
2809:
2810: \bibitem[{\citenamefont{Raizen et~al.}(1998)\citenamefont{Raizen, Koga,
2811: Sundaram, Kishimoto, Takuma, and Tajima}}]{Raizen98}
2812: \bibinfo{author}{\bibfnamefont{M.~G.} \bibnamefont{Raizen}},
2813: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Koga}},
2814: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Sundaram}},
2815: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kishimoto}},
2816: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takuma}}, \bibnamefont{and}
2817: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Tajima}},
2818: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{58}},
2819: \bibinfo{pages}{4757} (\bibinfo{year}{1998}).
2820:
2821: \bibitem[{\citenamefont{Belavkin}(1999)}]{Belavkin99}
2822: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}},
2823: \bibinfo{journal}{{Rep.\ Math.\ Phys.}} \textbf{\bibinfo{volume}{43}},
2824: \bibinfo{pages}{405} (\bibinfo{year}{1999}).
2825:
2826: \bibitem[{\citenamefont{Doherty and Jacobs}(1999)}]{Doherty99}
2827: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}} \bibnamefont{and}
2828: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}},
2829: \bibinfo{journal}{Phys.\ Rev.\ A} \textbf{\bibinfo{volume}{60}},
2830: \bibinfo{pages}{2700} (\bibinfo{year}{1999}).
2831:
2832: \bibitem[{\citenamefont{Doherty et~al.}(2000)\citenamefont{Doherty, Habib,
2833: Jacobs, Mabuchi, and Tan}}]{Doherty00}
2834: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}},
2835: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Habib}},
2836: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}},
2837: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}}, \bibnamefont{and}
2838: \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Tan}},
2839: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{62}},
2840: \bibinfo{pages}{012105} (\bibinfo{year}{2000}).
2841:
2842: \bibitem[{\citenamefont{Armen et~al.}(2002)\citenamefont{Armen, Au, Stockton,
2843: Doherty, and Mabuchi}}]{Armen02}
2844: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Armen}},
2845: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Au}},
2846: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2847: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}},
2848: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2849: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{89}},
2850: \bibinfo{pages}{133602} (\bibinfo{year}{2002}).
2851:
2852: \bibitem[{\citenamefont{Fischer et~al.}(2002)\citenamefont{Fischer, Maunz,
2853: Pinkse, Puppe, and Rempe}}]{Fischer02}
2854: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Fischer}},
2855: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Maunz}},
2856: \bibinfo{author}{\bibfnamefont{P.~W.~H.} \bibnamefont{Pinkse}},
2857: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Puppe}}, \bibnamefont{and}
2858: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rempe}},
2859: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{88}},
2860: \bibinfo{pages}{163002} (\bibinfo{year}{2002}).
2861:
2862: \bibitem[{\citenamefont{D.~Vitali and Tombesi}(2003)}]{Vitali03}
2863: \bibinfo{author}{\bibfnamefont{L.~R.} \bibnamefont{D.~Vitali},
2864: \bibfnamefont{S.~Mancini}} \bibnamefont{and}
2865: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tombesi}},
2866: \bibinfo{journal}{{J.\ Opt.\ Soc.\ Am.\ B}} \textbf{\bibinfo{volume}{20}},
2867: \bibinfo{pages}{1054} (\bibinfo{year}{2003}).
2868:
2869: \bibitem[{\citenamefont{Hopkins et~al.}(2003)\citenamefont{Hopkins, Jacobs,
2870: Habib, and Schwab}}]{Hopkins03}
2871: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Hopkins}},
2872: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}},
2873: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Habib}}, \bibnamefont{and}
2874: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Schwab}},
2875: \bibinfo{journal}{{Phys.\ Rev.\ B}} \textbf{\bibinfo{volume}{68}},
2876: \bibinfo{pages}{235328} (\bibinfo{year}{2003}).
2877:
2878: \bibitem[{\citenamefont{Geremia et~al.}(2003)\citenamefont{Geremia, Stockton,
2879: and Mabuchi}}]{Geremia03}
2880: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Geremia}},
2881: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2882: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2883: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0309034}
2884: (\bibinfo{year}{2003}).
2885:
2886: \bibitem[{\citenamefont{Geremia
2887: et~al.}(2004{\natexlab{a}})\citenamefont{Geremia, Stockton, and
2888: Mabuchi}}]{Geremia04}
2889: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Geremia}},
2890: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2891: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2892: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{304}},
2893: \bibinfo{pages}{270} (\bibinfo{year}{2004}{\natexlab{a}}).
2894:
2895: \bibitem[{\citenamefont{va{n H}andel et~al.}(2005)\citenamefont{va{n H}andel,
2896: Stockton, and Mabuchi}}]{vanHandel04}
2897: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{va{n H}andel}},
2898: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2899: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2900: \bibinfo{journal}{IEEE Trans. Automat. Control}
2901: \textbf{\bibinfo{volume}{50}}, \bibinfo{pages}{768} (\bibinfo{year}{2005}).
2902:
2903: \bibitem[{\citenamefont{Geremia
2904: et~al.}(2004{\natexlab{b}})\citenamefont{Geremia, Stockton, and
2905: Mabuchi}}]{Geremia04b}
2906: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Geremia}},
2907: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2908: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2909: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0401107}
2910: (\bibinfo{year}{2004}{\natexlab{b}}).
2911:
2912: \bibitem[{\citenamefont{Geremia et~al.}(2005)\citenamefont{Geremia, Stockton,
2913: and Mabuchi}}]{Geremia05}
2914: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Geremia}},
2915: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Stockton}},
2916: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2917: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0501033}
2918: (\bibinfo{year}{2005}).
2919:
2920: \bibitem[{\citenamefont{Smith et~al.}(2004)\citenamefont{Smith, Chaudhury,
2921: Silberfarb, Deutsch, and Jessen}}]{Smith04}
2922: \bibinfo{author}{\bibfnamefont{G.~A.} \bibnamefont{Smith}},
2923: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Chaudhury}},
2924: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Silberfarb}},
2925: \bibinfo{author}{\bibfnamefont{I.~H.} \bibnamefont{Deutsch}},
2926: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~S.}
2927: \bibnamefont{Jessen}}, \bibinfo{journal}{{Phys.\ Rev.\ Lett.}}
2928: \textbf{\bibinfo{volume}{93}}, \bibinfo{pages}{163602}
2929: (\bibinfo{year}{2004}).
2930:
2931: \bibitem[{\citenamefont{Smith et~al.}(2002)\citenamefont{Smith, Reiner, Orozco,
2932: Kuhr, and Wiseman}}]{Smith02}
2933: \bibinfo{author}{\bibfnamefont{W.~P.} \bibnamefont{Smith}},
2934: \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Reiner}},
2935: \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Orozco}},
2936: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kuhr}}, \bibnamefont{and}
2937: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}},
2938: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{89}},
2939: \bibinfo{pages}{133601} (\bibinfo{year}{2002}).
2940:
2941: \bibitem[{\citenamefont{James}(2004)}]{James04}
2942: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{James}},
2943: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{69}},
2944: \bibinfo{pages}{032108} (\bibinfo{year}{2004}).
2945:
2946: \bibitem[{\citenamefont{Reiner et~al.}(2004)\citenamefont{Reiner, Smith,
2947: Orozco, Wiseman, and Gambetta}}]{Reiner04}
2948: \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Reiner}},
2949: \bibinfo{author}{\bibfnamefont{W.~P.} \bibnamefont{Smith}},
2950: \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Orozco}},
2951: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}},
2952: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Gambetta}},
2953: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{70}},
2954: \bibinfo{pages}{023819} (\bibinfo{year}{2004}).
2955:
2956: \bibitem[{\citenamefont{Hanssen et~al.}(2002)\citenamefont{Hanssen, Milner,
2957: Meyrath, and Raizen}}]{Hanssen02}
2958: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hanssen}},
2959: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Milner}},
2960: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Meyrath}}, \bibnamefont{and}
2961: \bibinfo{author}{\bibfnamefont{M.~G.} \bibnamefont{Raizen}},
2962: \emph{\bibinfo{title}{Real time control of atomic motion using feedback}},
2963: \bibinfo{howpublished}{to be published in the conferrence proceedings of
2964: ICQO8} (\bibinfo{year}{2002}).
2965:
2966: \bibitem[{\citenamefont{Morrow et~al.}(2002)\citenamefont{Morrow, Dutta, and
2967: Raithel}}]{Morrow02}
2968: \bibinfo{author}{\bibfnamefont{N.~V.} \bibnamefont{Morrow}},
2969: \bibinfo{author}{\bibfnamefont{S.~K.} \bibnamefont{Dutta}}, \bibnamefont{and}
2970: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Raithel}},
2971: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{88}},
2972: \bibinfo{pages}{093003} (\bibinfo{year}{2002}).
2973:
2974: \bibitem[{\citenamefont{Vuleti{\'c} et~al.}(2004)\citenamefont{Vuleti{\'c},
2975: Black, and Thompson}}]{Vuletic04}
2976: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Vuleti{\'c}}},
2977: \bibinfo{author}{\bibfnamefont{A.~T.} \bibnamefont{Black}}, \bibnamefont{and}
2978: \bibinfo{author}{\bibfnamefont{J.~K.} \bibnamefont{Thompson}},
2979: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0410168}
2980: (\bibinfo{year}{2004}).
2981:
2982: \bibitem[{\citenamefont{Steck et~al.}(2004)\citenamefont{Steck, Jacobs,
2983: Mabuchi, Bhattacharya, and Habib}}]{Steck04}
2984: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Steck}},
2985: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}},
2986: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mabuchi}},
2987: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Bhattacharya}},
2988: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Habib}},
2989: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{92}},
2990: \bibinfo{pages}{223004} (\bibinfo{year}{2004}).
2991:
2992: \bibitem[{\citenamefont{Steixner et~al.}(2005)\citenamefont{Steixner, Rabl, and
2993: Zoller}}]{Rabl05}
2994: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Steixner}},
2995: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Rabl}}, \bibnamefont{and}
2996: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zoller}},
2997: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0506187}
2998: (\bibinfo{year}{2005}).
2999:
3000: \bibitem[{\citenamefont{Rabl et~al.}(2005)\citenamefont{Rabl, Steixner, and
3001: Zoller}}]{Steixner05}
3002: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Rabl}},
3003: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Steixner}}, \bibnamefont{and}
3004: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zoller}},
3005: \bibinfo{howpublished}{arXiv.org preprint quant-ph/0506185}
3006: (\bibinfo{year}{2005}).
3007:
3008: \bibitem[{\citenamefont{Wiseman and Doherty}(2005)}]{Wiseman05}
3009: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}} \bibnamefont{and}
3010: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}},
3011: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{94}},
3012: \bibinfo{pages}{070405} (\bibinfo{year}{2005}).
3013:
3014: \bibitem[{\citenamefont{Horak et~al.}(1997)\citenamefont{Horak, Hechenblaikner,
3015: Gheri, Stecher, and Ritsch}}]{Horak97}
3016: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Horak}},
3017: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hechenblaikner}},
3018: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Gheri}},
3019: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Stecher}}, \bibnamefont{and}
3020: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3021: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{79}},
3022: \bibinfo{pages}{4974} (\bibinfo{year}{1997}).
3023:
3024: \bibitem[{\citenamefont{Alge et~al.}(1997)\citenamefont{Alge, Ellinger,
3025: Stecher, Gheri, and Ritsch}}]{Alge97}
3026: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Alge}},
3027: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Ellinger}},
3028: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Stecher}},
3029: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Gheri}}, \bibnamefont{and}
3030: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3031: \bibinfo{journal}{{Europhys.\ Lett.}} \textbf{\bibinfo{volume}{39}},
3032: \bibinfo{pages}{491} (\bibinfo{year}{1997}).
3033:
3034: \bibitem[{\citenamefont{Gangl and Ritsch}(2000{\natexlab{a}})}]{Gangl00}
3035: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Gangl}} \bibnamefont{and}
3036: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3037: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{61}},
3038: \bibinfo{pages}{011402(R)} (\bibinfo{year}{2000}{\natexlab{a}}).
3039:
3040: \bibitem[{\citenamefont{Gangl and Ritsch}(2000{\natexlab{b}})}]{Gangl00b}
3041: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Gangl}} \bibnamefont{and}
3042: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3043: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{61}},
3044: \bibinfo{pages}{043405} (\bibinfo{year}{2000}{\natexlab{b}}).
3045:
3046: \bibitem[{\citenamefont{Vuleti{\'c} and Chu}(2000)}]{Vuletic00}
3047: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Vuleti{\'c}}} \bibnamefont{and}
3048: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Chu}},
3049: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{84}},
3050: \bibinfo{pages}{3787} (\bibinfo{year}{2000}).
3051:
3052: \bibitem[{\citenamefont{Domokos et~al.}(2001)\citenamefont{Domokos, Horak, and
3053: Ritsch}}]{Domokos01}
3054: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Domokos}},
3055: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Horak}}, \bibnamefont{and}
3056: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3057: \bibinfo{journal}{{J.\ Phys.\ B}} \textbf{\bibinfo{volume}{34}},
3058: \bibinfo{pages}{187} (\bibinfo{year}{2001}).
3059:
3060: \bibitem[{\citenamefont{Gangl and Ritsch}(2001)}]{Gangl01}
3061: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Gangl}} \bibnamefont{and}
3062: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3063: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{64}},
3064: \bibinfo{pages}{063414} (\bibinfo{year}{2001}).
3065:
3066: \bibitem[{\citenamefont{Horak and Ritsch}(2001)}]{Horak01}
3067: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Horak}} \bibnamefont{and}
3068: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ritsch}},
3069: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{64}},
3070: \bibinfo{pages}{033422} (\bibinfo{year}{2001}).
3071:
3072: \bibitem[{\citenamefont{van Enk et~al.}(2001)\citenamefont{van Enk, McKeever,
3073: Kimble, and Ye}}]{vanEnk01}
3074: \bibinfo{author}{\bibfnamefont{S.~J.} \bibnamefont{van Enk}},
3075: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{McKeever}},
3076: \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Kimble}},
3077: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ye}},
3078: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{64}},
3079: \bibinfo{pages}{013407} (\bibinfo{year}{2001}).
3080:
3081: \bibitem[{\citenamefont{Maunz et~al.}(2004)\citenamefont{Maunz, Puppe,
3082: Schuster, Syassen, Pinkse, and Rempe}}]{Maunz04}
3083: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Maunz}},
3084: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Puppe}},
3085: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Schuster}},
3086: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Syassen}},
3087: \bibinfo{author}{\bibfnamefont{P.~W.~H.} \bibnamefont{Pinkse}},
3088: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Rempe}},
3089: \bibinfo{journal}{{Nature}} \textbf{\bibinfo{volume}{428}},
3090: \bibinfo{pages}{50} (\bibinfo{year}{2004}).
3091:
3092: \bibitem[{\citenamefont{Belavkin}(1988)}]{Belavkin88}
3093: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}}, in
3094: \emph{\bibinfo{booktitle}{Modelling and Control of Systems}}, edited by
3095: \bibinfo{editor}{\bibfnamefont{A.}~\bibnamefont{Blaqui{\`e}re}}
3096: (\bibinfo{publisher}{Springer}, \bibinfo{address}{New York},
3097: \bibinfo{year}{1988}).
3098:
3099: \bibitem[{\citenamefont{Belavkin}(1989)}]{Belavkin89}
3100: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}},
3101: \bibinfo{journal}{{Phys.\ Lett.\ A}} \textbf{\bibinfo{volume}{140}},
3102: \bibinfo{pages}{355} (\bibinfo{year}{1989}).
3103:
3104: \bibitem[{\citenamefont{Chru{\'s}ei{\'n}ski and
3105: Staszewski}(1992)}]{Chruseinski92}
3106: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chru{\'s}ei{\'n}ski}}
3107: \bibnamefont{and}
3108: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Staszewski}},
3109: \bibinfo{journal}{{Physica Scripta}} \textbf{\bibinfo{volume}{45}},
3110: \bibinfo{pages}{193} (\bibinfo{year}{1992}).
3111:
3112: \bibitem[{\citenamefont{Srinivas and Davies}(1981)}]{Srinivas81}
3113: \bibinfo{author}{\bibfnamefont{M.~D.} \bibnamefont{Srinivas}} \bibnamefont{and}
3114: \bibinfo{author}{\bibfnamefont{E.~B.} \bibnamefont{Davies}},
3115: \bibinfo{journal}{{Optica Acta}} \textbf{\bibinfo{volume}{28}},
3116: \bibinfo{pages}{981} (\bibinfo{year}{1981}).
3117:
3118: \bibitem[{\citenamefont{Barchielli et~al.}(1983)\citenamefont{Barchielli, Lanz,
3119: and Prosperi}}]{Barchielli83}
3120: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Barchielli}},
3121: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Lanz}}, \bibnamefont{and}
3122: \bibinfo{author}{\bibfnamefont{G.~M.} \bibnamefont{Prosperi}},
3123: \bibinfo{journal}{{Optica Acta}} \textbf{\bibinfo{volume}{13}},
3124: \bibinfo{pages}{779} (\bibinfo{year}{1983}).
3125:
3126: \bibitem[{\citenamefont{Barchielli and Lupieri}(1985)}]{Barchielli85}
3127: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Barchielli}} \bibnamefont{and}
3128: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Lupieri}},
3129: \bibinfo{journal}{{J. Math. Phys.}} \textbf{\bibinfo{volume}{26}},
3130: \bibinfo{pages}{2222} (\bibinfo{year}{1985}).
3131:
3132: \bibitem[{\citenamefont{Diosi}(1986{\natexlab{a}})}]{Diosi86}
3133: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Diosi}},
3134: \bibinfo{journal}{{Phys.\ Lett.\ A}} \textbf{\bibinfo{volume}{114}},
3135: \bibinfo{pages}{451} (\bibinfo{year}{1986}{\natexlab{a}}).
3136:
3137: \bibitem[{\citenamefont{Diosi}(1986{\natexlab{b}})}]{Diosi88}
3138: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Diosi}},
3139: \bibinfo{journal}{{Phys.\ Lett.\ A}} \textbf{\bibinfo{volume}{129}},
3140: \bibinfo{pages}{419} (\bibinfo{year}{1986}{\natexlab{b}}).
3141:
3142: \bibitem[{\citenamefont{Gisin}(1984)}]{Gisin84}
3143: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Gisin}},
3144: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{52}},
3145: \bibinfo{pages}{1657} (\bibinfo{year}{1984}).
3146:
3147: \bibitem[{\citenamefont{Caves and Milburn}(1987)}]{Caves87}
3148: \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{Caves}} \bibnamefont{and}
3149: \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Milburn}},
3150: \bibinfo{journal}{Phys.\ Rev.\ A} \textbf{\bibinfo{volume}{36}},
3151: \bibinfo{pages}{5543} (\bibinfo{year}{1987}).
3152:
3153: \bibitem[{\citenamefont{Barchielli}(1993)}]{Barchielli93}
3154: \bibinfo{author}{\bibnamefont{Barchielli}}, \bibinfo{journal}{{Int.\ J. Theor.\
3155: Phys.}} \textbf{\bibinfo{volume}{32}}, \bibinfo{pages}{2221}
3156: (\bibinfo{year}{1993}).
3157:
3158: \bibitem[{\citenamefont{Carmichael}(1993)}]{Carmichael93}
3159: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Carmichael}},
3160: \emph{\bibinfo{title}{An Open Systems Approach to Quantum Optics}}
3161: (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{Berlin},
3162: \bibinfo{year}{1993}).
3163:
3164: \bibitem[{\citenamefont{Wiseman and Milburn}(1993)}]{Wiseman93}
3165: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}} \bibnamefont{and}
3166: \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Milburn}},
3167: \bibinfo{journal}{Phys.\ Rev.\ A} \textbf{\bibinfo{volume}{47}},
3168: \bibinfo{pages}{642} (\bibinfo{year}{1993}).
3169:
3170: \bibitem[{\citenamefont{Jacobs}(1993)}]{Jacobs93}
3171: \bibinfo{author}{\bibfnamefont{O.~L.~R.} \bibnamefont{Jacobs}},
3172: \emph{\bibinfo{title}{Introduction to Control Theory}}
3173: (\bibinfo{publisher}{Oxford University Press}, \bibinfo{address}{Oxford},
3174: \bibinfo{year}{1993}).
3175:
3176: \bibitem[{\citenamefont{Maybeck}(1996)}]{Maybeck82}
3177: \bibinfo{author}{\bibfnamefont{P.~S.} \bibnamefont{Maybeck}},
3178: \emph{\bibinfo{title}{Stochastic Models, Estimation, and Control}}
3179: (\bibinfo{publisher}{Wiley}, \bibinfo{address}{Chichester},
3180: \bibinfo{year}{1996}).
3181:
3182: \bibitem[{\citenamefont{Whittle}(1982)}]{Whittle96}
3183: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Whittle}},
3184: \emph{\bibinfo{title}{Optimal Control: Basics and Beyond}}
3185: (\bibinfo{publisher}{Academic Press}, \bibinfo{address}{New York},
3186: \bibinfo{year}{1982}).
3187:
3188: \bibitem[{\citenamefont{Zhou et~al.}(1995)\citenamefont{Zhou, Doyle, and
3189: Glover}}]{Zhou95}
3190: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Zhou}},
3191: \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Doyle}}, \bibnamefont{and}
3192: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Glover}},
3193: \emph{\bibinfo{title}{Robust and Optimal Control}}
3194: (\bibinfo{publisher}{Prentice Hall}, \bibinfo{address}{Englewood Cliffs},
3195: \bibinfo{year}{1995}).
3196:
3197: \bibitem[{\citenamefont{Wiseman et~al.}(2002)\citenamefont{Wiseman, Mancini,
3198: and Wang}}]{Wiseman02}
3199: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}},
3200: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Mancini}}, \bibnamefont{and}
3201: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Wang}},
3202: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{66}},
3203: \bibinfo{pages}{013807} (\bibinfo{year}{2002}).
3204:
3205: \bibitem[{\citenamefont{Jacobs}(2003)}]{Jacobs03}
3206: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Jacobs}},
3207: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{67}},
3208: \bibinfo{pages}{030301(R)} (\bibinfo{year}{2003}).
3209:
3210: \bibitem[{\citenamefont{Holland et~al.}(1991)\citenamefont{Holland, Walls, and
3211: Zoller}}]{Holland91}
3212: \bibinfo{author}{\bibfnamefont{M.~J.} \bibnamefont{Holland}},
3213: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}}, \bibnamefont{and}
3214: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zoller}},
3215: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{67}},
3216: \bibinfo{pages}{1716} (\bibinfo{year}{1991}).
3217:
3218: \bibitem[{\citenamefont{Storey et~al.}(1992)\citenamefont{Storey, Collett, and
3219: Walls}}]{Storey92}
3220: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Storey}},
3221: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Collett}}, \bibnamefont{and}
3222: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Walls}},
3223: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{68}},
3224: \bibinfo{pages}{472} (\bibinfo{year}{1992}).
3225:
3226: \bibitem[{\citenamefont{Quadt et~al.}(1995)\citenamefont{Quadt, Collett, and
3227: Walls}}]{Quadt95}
3228: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Quadt}},
3229: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Collett}}, \bibnamefont{and}
3230: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}},
3231: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{74}},
3232: \bibinfo{pages}{351} (\bibinfo{year}{1995}).
3233:
3234: \bibitem[{\citenamefont{Doherty et~al.}(1998)\citenamefont{Doherty, Parkins,
3235: Tan, and Walls}}]{Doherty98}
3236: \bibinfo{author}{\bibfnamefont{A.~C.} \bibnamefont{Doherty}},
3237: \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{Parkins}},
3238: \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Tan}}, \bibnamefont{and}
3239: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}},
3240: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{57}},
3241: \bibinfo{pages}{4804} (\bibinfo{year}{1998}).
3242:
3243: \bibitem[{not()}]{noteSME}
3244: \bibinfo{note}{The terms in the stochastic master equation that describe
3245: homodyne detection are derived in~\cite{Wiseman93}. The term that describes
3246: spontaneous emission is given in, e.g.,~\cite{Jacobs96, Marte94}.}
3247:
3248: \bibitem[{\citenamefont{Belavkin}(1979)}]{Belavkin79}
3249: \bibinfo{author}{\bibfnamefont{V.~P.} \bibnamefont{Belavkin}},
3250: \bibinfo{journal}{{Preprint Instytut Fizyki}} \textbf{\bibinfo{volume}{411}},
3251: \bibinfo{pages}{3} (\bibinfo{year}{1979}).
3252:
3253: \bibitem[{\citenamefont{Wiseman}(1994)}]{Wiseman94b}
3254: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}},
3255: \bibinfo{type}{{Ph.D.}\ thesis}, \bibinfo{school}{University of Queensland}
3256: (\bibinfo{year}{1994}).
3257:
3258: \bibitem[{\citenamefont{Hood et~al.}(1998)\citenamefont{Hood, Chapman, Lynn,
3259: and Kimble}}]{Hood98}
3260: \bibinfo{author}{\bibfnamefont{C.~J.} \bibnamefont{Hood}},
3261: \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Chapman}},
3262: \bibinfo{author}{\bibfnamefont{T.~W.} \bibnamefont{Lynn}}, \bibnamefont{and}
3263: \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Kimble}},
3264: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{80}},
3265: \bibinfo{pages}{4157} (\bibinfo{year}{1998}).
3266:
3267: \bibitem[{\citenamefont{Kloeden and Platen}(1992)}]{Kloeden92}
3268: \bibinfo{author}{\bibfnamefont{P.~E.} \bibnamefont{Kloeden}} \bibnamefont{and}
3269: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Platen}},
3270: \emph{\bibinfo{title}{Numerical Solution of Stochastic Differential
3271: Equations}} (\bibinfo{publisher}{Springer}, \bibinfo{address}{Berlin},
3272: \bibinfo{year}{1992}).
3273:
3274: \bibitem[{\citenamefont{Zurek et~al.}(1993)\citenamefont{Zurek, Habib, and
3275: Paz}}]{Zurek93}
3276: \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Zurek}},
3277: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Habib}}, \bibnamefont{and}
3278: \bibinfo{author}{\bibfnamefont{J.~P.} \bibnamefont{Paz}},
3279: \bibinfo{journal}{{Phys.\ Rev.\ Lett.}} \textbf{\bibinfo{volume}{70}},
3280: \bibinfo{pages}{1187} (\bibinfo{year}{1993}).
3281:
3282: \bibitem[{\citenamefont{Jacobs et~al.}(1996)\citenamefont{Jacobs, Collett,
3283: Wiseman, Tan, and Walls}}]{Jacobs96}
3284: \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Jacobs}},
3285: \bibinfo{author}{\bibfnamefont{M.~J.} \bibnamefont{Collett}},
3286: \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Wiseman}},
3287: \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Tan}}, \bibnamefont{and}
3288: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}},
3289: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{54}},
3290: \bibinfo{pages}{2260} (\bibinfo{year}{1996}).
3291:
3292: \bibitem[{\citenamefont{Marte et~al.}(1994)\citenamefont{Marte, Dum,
3293: Ta\"{\i}eb, Zoller, Shahriar, and Prentiss}}]{Marte94}
3294: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Marte}},
3295: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Dum}},
3296: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Ta\"{\i}eb}},
3297: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zoller}},
3298: \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Shahriar}},
3299: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Prentiss}},
3300: \bibinfo{journal}{{Phys.\ Rev.\ A}} \textbf{\bibinfo{volume}{49}},
3301: \bibinfo{pages}{4826} (\bibinfo{year}{1994}).
3302:
3303: \end{thebibliography}
3304:
3305:
3306:
3307:
3308:
3309: %%% proposal:
3310: %% feedback:
3311: %% geremia 04 [31]
3312: %% geremia 05 [32] Alkali
3313: %% smith 04 [37] also nonlinear
3314: %% smith02 [38]
3315: %% Reiner04 [39]
3316: %% zoller [52 53]
3317: %% hanssen [98], morrow [100]
3318: %%
3319: %% cqed measurement muenstermann 99 [46]
3320: %%
3321:
3322: %\end{thebibliography}
3323:
3324: \end{document}
3325:
3326: