1: \documentclass[aps,preprint,superbib]{revtex4}%
2: \usepackage{amsfonts}
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{graphicx}%
6: \setcounter{MaxMatrixCols}{30}
7: %TCIDATA{OutputFilter=latex2.dll}
8: %TCIDATA{Version=5.50.0.2890}
9: %TCIDATA{CSTFile=revtex4.cst}
10: %TCIDATA{Created=Thursday, September 30, 2004 09:27:56}
11: %TCIDATA{LastRevised=Wednesday, May 30, 2007 21:23:21}
12: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
13: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
14: %TCIDATA{BibliographyScheme=BibTeX}
15: %TCIDATA{<META NAME="DocumentShell" CONTENT="Articles\SW\REVTeX 4">}
16: %TCIDATA{Language=American English}
17: %BeginMSIPreambleData
18: \providecommand{\U}[1]{\protect\rule{.1in}{.1in}}
19: %EndMSIPreambleData
20: \newtheorem{theorem}{Theorem}
21: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
22: \newtheorem{algorithm}[theorem]{Algorithm}
23: \newtheorem{axiom}[theorem]{Axiom}
24: \newtheorem{claim}[theorem]{Claim}
25: \newtheorem{conclusion}[theorem]{Conclusion}
26: \newtheorem{condition}[theorem]{Condition}
27: \newtheorem{conjecture}[theorem]{Conjecture}
28: \newtheorem{corollary}[theorem]{Corollary}
29: \newtheorem{criterion}[theorem]{Criterion}
30: \newtheorem{definition}[theorem]{Definition}
31: \newtheorem{example}[theorem]{Example}
32: \newtheorem{exercise}[theorem]{Exercise}
33: \newtheorem{lemma}[theorem]{Lemma}
34: \newtheorem{notation}[theorem]{Notation}
35: \newtheorem{problem}[theorem]{Problem}
36: \newtheorem{proposition}[theorem]{Proposition}
37: \newtheorem{remark}[theorem]{Remark}
38: \newtheorem{solution}[theorem]{Solution}
39: \newtheorem{summary}[theorem]{Summary}
40: \newenvironment{proof}[1][Proof]{\noindent\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
41: \renewcommand{\baselinestretch}{1.5}
42: \begin{document}
43: \title{Observation-assisted optimal control of quantum dynamics}
44: \author{Feng Shuang, Alexander Pechen, Tak-San Ho and Herschel Rabitz}
45: \affiliation{Department of Chemistry, Princeton University, Princeton, NJ 08544}
46: \date{\today}
47:
48: \begin{abstract}
49: This paper explores the utility of instantaneous and continuous observations
50: in the optimal control of quantum dynamics. Simulations of the processes are
51: performed on several multilevel quantum systems with the goal of population
52: transfer. Optimal control fields are shown to be capable of cooperating or
53: fighting with observations to achieve a good yield, and the nature of the
54: observations may be optimized to more effectively control the quantum
55: dynamics. Quantum observations also can break dynamical symmetries to increase
56: the controllability of a quantum system. The quantum Zeno and anti-Zeno
57: effects induced by observations are the key operating principles in these
58: processes. The results indicate that quantum observations can be effective
59: tools in the control of quantum dynamics.
60:
61: \end{abstract}
62: \maketitle
63:
64:
65: \section{Introduction}
66:
67: The control of quantum processes is actively being pursued
68: theoretically\cite{Rice00,Rabitz0364,Shapiro03} and
69: experimentally\cite{Walmsley0343,Brixner011} with a variety of control
70: scenarios\cite{Rabitz884950,Kosloff89201,Shapiro864103,Bergmann905363,
71: Rice982885}. An increasing number of successful control experiments, including
72: in complex systems\cite{Gerber98919, Gerber9910381, Kunde00924, Bartels00164,
73: Brixner0157, Levis01709, Herek02533, Daniel03536}, employ closed-loop optimal
74: control\cite{Judson921500}. The latter experiments commonly aim to enhance the
75: yield of a particular desired final state, where a measurement of the quantum
76: system is only performed after the evolution is over. Utilizing quantum
77: observations \textit{during} the control process may offer an opportunity to
78: enhance performance\cite{Rice049984, Sugawara05204115}. Recent
79: studies\cite{Shuang049270,Shuang06154105} also have shown that controlled
80: quantum dynamics can operate in the presence of significant field noise and
81: decoherence, and even cooperate with them under suitable circumstances. This
82: paper will demonstrate that analogous control cooperation can occur between
83: the actions of applied external fields and observations with both aiming to
84: manipulate the system's quantum dynamics.
85:
86: A characteristic feature of quantum mechanics is that the performance of a
87: measurement unavoidably affects the subsequent system dynamics. A well known
88: manifestation of this observation driven back action is the uncertainty
89: principle\cite{Mensky1993}. A direct influence of a measurement is revealed
90: through a change in the system state. In the von Neuman view of quantum
91: mechanics, an instantaneous measurement projects the state of the system onto
92: an eigenstate of the observable operator\cite{VonNeumann1955}. The measurement
93: process induces irreversible dynamics and results in a lack of system
94: coherence, corresponding to the off-diagonal matrix elements of the density
95: matrix decaying to zero or the phase of the wavefunction amplitudes being randomized.
96:
97: This paper is concerned with measurements carried out over a period of time.
98: One of the earliest approaches to continuous quantum measurements was
99: suggested by Feynman in terms of path integrals\cite{Feynman48367}. When
100: measurements are performed the Feynman propagator is modified by restricting
101: the paths to cross (or not to cross) certain space-time regions. An
102: approximate technique was developed by Mensky\cite{Mensky94159} who
103: incorporated Gaussian cut-offs in the phase space path integrals and showed
104: its equivalence to the phenomenological master equation approach for open
105: quantum system dynamics using models of system-environment
106: coupling\cite{Walls1994}.
107:
108: Prevention of a quantum system's time evolution by means of repetitive,
109: frequent observations or continuous observations of the system's state is
110: called the quantum Zeno effect (QZE). The QZE was proposed by Misra and
111: Sudarshan\cite{Misra77756} and was experimentally
112: demonstrated\cite{Itano902295} in a repeatedly measured two-level system
113: undergoing Rabi oscillations. A time-dependent observable projection operator
114: inducing up to $100\%$ transfer from one state to another
115: state\cite{Roy004019} is called the quantum anti-Zeno effect (QAZE). The
116: impacts of QZE and QAZE operations are the key processes explored in this
117: paper to help control quantum dynamics.
118:
119: This paper explores the scope of what might be gained in terms of better
120: control performance from utilizing suitable observations. The practical means
121: of executing observations in this fashion will be the subject of future works.
122: The remainder of the paper is broken down the following way. Section II
123: reviews the main concepts of performing instantaneous and continuous
124: measurements, which are utilized in this paper. Section III presents the model
125: system, and Section IV presents simulations of the closed-loop management of
126: quantum dynamics assisted by measurements. A brief summary of the findings is
127: given in Section V.
128:
129: \section{Quantum Observations}
130:
131: \subsection{Instantaneous observations}
132:
133: An ideal instantaneous measurement occurs at one point of time, or a sequence
134: of such observations can follow each other at different
135: times\cite{VonNeumann1955}. An instantaneous measurement may be characterized
136: by a set of projectors $\left\{ P_{i}\right\} $ satisfying conditions of
137: completeness and orthogonality%
138: \begin{equation}
139: \sum_{k}P_{k}=1,\ \ P_{i}P_{j}=0\ \text{for }i\neq j\text{.}%
140: \end{equation}
141: The instantaneous measurement converts the state $\rho$ into the state
142: $\rho^{\prime}$,%
143: \begin{equation}
144: \rho^{\prime}=\sum_{k}P_{k}\rho P_{k}\text{.} \label{MRho}%
145: \end{equation}
146: We may also observe a physical quantity represented by the operator $A$,
147: \begin{equation}
148: A=\sum_{i}a_{i}\left\vert a_{i}\right\rangle \left\langle a_{i}\right\vert
149: \text{,}%
150: \end{equation}
151: where $a_{i}$ and $\left\vert a_{i}\right\rangle $ are the i-th eigenvalue and
152: eigenstate, respectively, of the observable operator $A$, and the density
153: matrix maybe expressed in the form%
154: \begin{equation}
155: \rho=\sum_{k,j}\rho_{kj}\left\vert a_{k}\right\rangle \left\langle
156: a_{j}\right\vert \text{.}%
157: \end{equation}
158: When a measurement of $A$ is performed, the reduction
159: \begin{equation}
160: \rho_{kj}\rightarrow0,\text{ for }a_{k}\neq a_{j}%
161: \end{equation}
162: occurs, thereby destroying the coherence between nondegenerate states of
163: operator $A$. If $A$ has no degenerate eigenstates, then $\rho$ will contain
164: only diagonal elements after an instantaneous quantum measurement%
165: \begin{equation}
166: \rho\rightarrow\sum_{k}\rho_{kk}\left\vert a_{k}\right\rangle \left\langle
167: a_{k}\right\vert \text{.} \label{InsM}%
168: \end{equation}
169: If a projection operator $P$ is observed, it's easy to deduce from Eq.
170: (\ref{MRho}) that after the observation process, the density matrix is
171: transformed to $\rho^{\prime}$ given by%
172:
173: \begin{subequations}
174: \begin{align}
175: \rho^{\prime} & =P\rho P+\left( 1-P\right) \rho\left( 1-P\right) \\
176: & =\rho-\left[ P,\left[ P,\rho\right] \right] \text{.} \label{OInst}%
177: \end{align}
178: The operation $\left[ P,\left[ P,\rho\right] \right] $ may be viewed as
179: the "kick" resulting from the instantaneous observation of the projection
180: operator $P$.
181:
182: \subsection{Continuous observations}
183:
184: The employment of restricted path integrals and master equations (ME) form two
185: equivalent techniques in the theory of continuous quantum
186: measurements\cite{Mensky94159}. For simplicity, we adopt the ME formalism.
187: With a continuous measurement of a single observable $A$ the ME takes the form
188: \cite{Walls1994}:%
189:
190: \end{subequations}
191: \begin{equation}
192: \dot{\rho}=-i\left[ H,\rho\right] -\frac{1}{2}\kappa\left[ A,\left[
193: A,\rho\right] \right] \text{.} \label{ME}%
194: \end{equation}
195: Here, $H$ is the Hamiltonian of the measured system, and $\kappa$ indicates
196: the "strength" of the observation. Equation (\ref{ME}) is similar to the
197: equation describing a system interacting with the environment. The first term
198: in Eq. (\ref{ME}) describes the propagation of the free system, while the
199: second term provides the decay of the off-diagonal matrix elements, such that%
200: \begin{equation}
201: \frac{\partial}{\partial t}\left\langle a_{i}|\rho|a_{j}\right\rangle
202: =-i\left\langle a_{i}\right\vert \left[ H,\rho\right] \left\vert
203: a_{j}\right\rangle -\frac{1}{2}\kappa\left( a_{i}-a_{j}\right)
204: ^{2}\left\langle a_{i}|\rho|a_{j}\right\rangle \text{.}%
205: \end{equation}
206:
207:
208: \section{The Model System}
209:
210: The effect of measurements on controlled quantum dynamics is explored here in
211: the context of population transfer in several multilevel systems characterized
212: by the Hamiltonian,%
213: \begin{subequations}
214: \begin{align}
215: H & =H_{0}-\mu E(t)\text{,}\label{Ht}\\
216: H_{0} & =\sum_{v}\varepsilon_{\upsilon}\left\vert \upsilon\right\rangle
217: \left\langle \upsilon\right\vert \text{,} \label{H0}%
218: \end{align}
219: where $\left\vert \upsilon\right\rangle $ is an eigenstate of $H_{0}$ and
220: $\varepsilon_{\upsilon}$ is the associated field-free eigenenergy, and $\mu$
221: is the dipole operator. The control field $E(t)$ is taken to have the
222: following form which may be implemented in the laboratory\cite{Rabitz05013419,
223: Dantus_OptExpress061030},
224: \end{subequations}
225: \begin{subequations}
226: \label{E0}%
227: \begin{align}
228: E(t) & =s(t)\sum_{l}^{M}A_{l}\cos\left( \omega_{l}t+\theta_{l}\right)
229: \text{,}\label{GE}\\
230: s(t) & =\exp\left[ -\left( t-T_{f}/2\right) ^{2}/2\sigma^{2}\right]
231: \text{,}%
232: \end{align}
233: where $\left\{ \omega_{l}\right\} $ are the $M$ allowed resonant transition
234: frequencies of the system and $s(t)$ is the pulse envelope function. The
235: controls are the amplitudes $\left\{ A_{l}\right\} $ and phases $\left\{
236: \theta_{l}\right\} $.
237:
238: Closed-loop control simulations will be performed to model a laboratory
239: circumstance with the cost function:%
240:
241: \end{subequations}
242: \begin{subequations}
243: \label{Obj}%
244: \begin{align}
245: J\left[ E(t)\right] & =\left\vert O\left[ E(t)\right] -O_{T}\right\vert
246: ^{2}+\alpha F\text{,}\label{J0}\\
247: F & =\sum_{l}\left( A_{l}\right) ^{2}\text{,} \label{F0}%
248: \end{align}
249: where $O_{T}$ is the target value (expressed as a percent yield) and
250: \end{subequations}
251: \begin{equation}
252: O\left[ E(t)\right] =\text{Tr}[\rho(T_{f})\hat{O}] \label{O}%
253: \end{equation}
254: is the outcome produced by the field $E(t)$ at time $T_{f}$, and $F$ is the
255: fluence of the control field whose contribution is weighted by the constant,
256: $\alpha>0$. In the present work, $\hat{O}=\left\vert \Psi_{f}\right\rangle
257: \left\langle \Psi_{f}\right\vert $ is a projection operator for transferring
258: population into the target state $\left\vert \Psi_{f}\right\rangle $. The goal
259: of this study is to explore the role that observations can play in aiding the
260: control process and possibly reducing the fluence of $E\left( t\right) $ to
261: more effectively achieve the desired final state.
262:
263: \section{Observations Serving as Controls}
264:
265: In this section, we numerically investigate four simple model systems in Fig.
266: \ref{Fig_Model} to explore the use of observations in the control of quantum
267: dynamics. In model 1, the control field is optimized and shown to be capable
268: of fighting against the effect of instantaneous observations of different
269: operators when they act as disturbances. The optimal control fields are also
270: capable of cooperating with the observation of the dipole to attain a better
271: value for the objective, even when the desired target yield is large. In model
272: 2, the control field is fixed but the instantaneous observed operators are
273: optimized. It is shown how the presence of even a non-optimal control field
274: can help the observation processes meet the target yield. Quantum observations
275: are used to break the dynamical symmetry in model 3, and the optimized
276: continuous observations are shown to assist in making the control process more
277: effective. In model 4, continuous observation is used to avoid population loss
278: into an undesired state. In the first two models, the QAZE is used to induce
279: population transfer, while the QZE is the operating process in models 3 and 4
280: used to prohibit population transfer. In all the illustrations a genetic
281: algorithm\cite{Goldberg97} is employed to optimize the control fields and observations.
282:
283: \subsection{Model 1}
284:
285: This model uses the five-level system in Fig. \ref{Fig_Model}(a) with
286: eigenstates $\left\vert i\right\rangle $, $i=0,\cdots,4$ of the field free
287: Hamiltonian $H_{0}$, having only nearest neighbor transitions with frequencies
288: $\omega_{01}=1.511$, $\omega_{12}=1.181$, $\omega_{23}=0.761$, and
289: $\omega_{34}=0.553$ in rad fs$^{-1}$, and associated transition dipole moments
290: $\mu_{01}=0.5855$, $\mu_{12}=0.7079$, $\mu_{23}=0.8352$ and $\mu_{34}=0.9281$
291: in $1.0\times10^{-30}$ C m. The target time is $T_{f}=200$ fs, the pulse width
292: in Eq. (\ref{E0}) is $\sigma=30$ fs, and the weight coefficient in Eq.
293: (\ref{J0}) is $\alpha=0.05$. The control objective is to transfer population
294: from the initially prepared ground state $\left\vert 0\right\rangle $ to the
295: highest excited state $\left\vert 4\right\rangle $, such that $\hat
296: {O}=\left\vert 4\right\rangle \left\langle 4\right\vert $ in Eq. (\ref{O}). As
297: a reference control case, we first determine the optimal control field without
298: any observations. Figure \ref{Fig_Field} depicts the amplitude and power
299: spectrum of control field. A population transfer of $98.44\%$ is achieved in
300: the target state by the optimal control field which has the fluence $0.063$.
301: The fields in all of the illustrations in this paper have general structure
302: similar to that in Fig. \ref{Fig_Field} due to the imposed form in Eq.
303: (\ref{E0}), and these other fields will not be explicitly shown.
304:
305: Assuming that for some auxiliary purpose we need to detect a physical quantity
306: $A$ at the middle of dynamical evolution at%
307: \begin{equation}
308: T_{m}=\frac{T_{f}}{2}\text{,} \label{Tm}%
309: \end{equation}
310: Table I shows how the optimally determined control fields (i.e., each
311: observation has a distinct optimal field of the form in Eq. (\ref{E0})) fight
312: against the observation of the dipole $\mu$, the energy $H_{0}$ and the
313: population of each level%
314: \begin{equation}
315: P_{k}=\left\vert k\right\rangle \left\langle k\right\vert \label{PopM}%
316: \end{equation}
317: with $k=0,\cdots,4$. The second column of Table I indicates that the control
318: field can fight very effectively with the disturbance caused by the individual
319: quantum observations. Note that the results for population observations (the
320: third column of Table I with $P_{k}$, $k=0,\cdots,4$) are all near zero, which
321: reveals the mechanism employed by each control field to fight against its
322: associated observation: the control field $E_{k}\left( t\right) $ associated
323: with the observation operator $P_{k}$ drives the system to a state
324: $\rho\left( T_{m}\right) =\left\vert \psi\right\rangle \left\langle
325: \psi\right\vert $ that is nearly orthogonal to the observed state $\left\vert
326: k\right\rangle $,
327: \begin{equation}
328: \left\langle \psi|k\right\rangle \approx0\text{,}%
329: \end{equation}
330: such that
331: \begin{equation}
332: \left[ P_{k},\left[ P_{k},\rho\left( T_{m}\right) \right] \right]
333: \approx0\text{.}%
334: \end{equation}
335: This behavior assures that the observation of $P_{k}$ has little effect on the
336: system state, or equivalently the "kick" from the observation disappears from
337: Eq. (\ref{OInst}). After checking the results of observing the energy and
338: dipole, we find a similar mechanism: their observed values at $T_{m}$ are all
339: nearly equal to an eigenvalue of the observed operators, which means that the
340: control field drives the system to an eigenstate of the observed operators at
341: $T_{m}$, again so that the observation has little effect on the system state.
342: It is evident in this case that the deleterious impact of any instantaneous
343: observation can be corrected because a suitable control field can drive model
344: 1 to any state. The fourth column in Table I uses the optimal fields
345: determined in the presence of the observation, but the dynamics are carried
346: out in the end without the observation being present. The very similar yields
347: in the second and fourth columns are consistent with the mechanism indicated
348: above. The last column in Table I shows that fighting against the disturbance
349: created by the observation increases the control field fluence, whose values
350: depend on the particular observation operator. These results collectively
351: indicate that in the present model when seeking a high target yield the most
352: efficient strategy for the control field is to fight the impact of the
353: observation, which is acting as a disturbance disruptive to the control goal.
354:
355: The observation of the dipole can have the dual competitive role of destroying
356: the coherence of the system, while also inducing population transfer. A
357: calculation shows that performance of an observation of the dipole $\mu$
358: without the control field being present can induce $22.19\%$ population
359: transfer from the initial state to the target state. Table II describes how
360: the optimal control fields work with an observation of the dipole $\mu$ to
361: reach different posed target yields. The second column shows that the target
362: yield can be reached in all the cases, with some lose in achieved fidelity at
363: the highest demanded yield of $O_{T}=100\%$. In order to reveal the
364: contributions of the observations upon the optimally controlled dynamics, the
365: third column of Table II shows the yield from the field alone without the
366: observation being made, yet with the field determined in the presence of the
367: observation. Comparison of the second and third column in Table II shows that
368: a remarkable degree of cooperation is found when the expected target yield
369: lies in the range greater than $22.19\%$ up to $\sim50\%$, and the effect is
370: even evident at the $70\%$ target yield. For example, at the target yield of
371: $O_{T}=40\%$, the observation and optimal field alone, respectively, produce
372: yields of $22.19\%$ and $2.69\%$. But, the same field operating in the
373: presence of the observation produces a yield of $39.82\%$. This behavior
374: indicates that the field is cooperating with the observation to more
375: effectively achieve the posed goal. Above a target yield of $\sim80\%,$ the
376: field works to fight against the observation acting as a disturbance. The
377: fourth column of Table II shows that the fluence generally follows this
378: behavior. Below a target yield of $\sim70\%$ and higher than $22.19\%$, the
379: reduced fluence with the observation being present shows the enhanced control
380: efficiency. Above that value the observation increasingly acts as a
381: disturbance, which calls for an enhanced field fluence to fight against it.
382:
383: \subsection{Model 2}
384:
385: Model 2 has the same Hamiltonian and dipole elements as model 1, but we
386: concentrate on studying the effects of a sequence of instantaneous
387: observations treated now as controls for the population transfer. Again, the
388: objective is to transfer population from level $0$ to level $4$ at the target
389: time $T_{f}=200$ fs. We assume that any projection operator may be observed in
390: a suitably performed experiment. A sequence of $N$ instantaneous projection
391: observations, specified by the operators
392: \begin{subequations}
393: \label{ProOp}%
394: \begin{align}
395: P_{k} & =\left\vert \psi_{k}\right\rangle \left\langle \psi_{k}\right\vert
396: \text{, }k=1,\cdots,N\text{,}\\
397: \left\vert \psi_{k}\right\rangle & =\sum_{j=0}^{4}a_{jk}\left\vert
398: j\right\rangle \text{, \ }\sum_{j=0}^{4}\left\vert a_{jk}\right\vert ^{2}=1
399: \end{align}
400: are performed at equally spaced time intervals,
401: \end{subequations}
402: \begin{equation}
403: t_{k}=\frac{k}{N+1}T_{f}\text{, }k=1,\cdots,N\text{,}%
404: \end{equation}
405: respectively. The variables subjected to optimization are the complex
406: coefficients $\left\{ a_{jk}\right\} $ in the projection operators of Eq.
407: (\ref{ProOp}). A control field, of the form Eq. (\ref{E0}), is utilized in
408: some of the simulations, but the amplitudes and phases are picked \textit{a
409: priori} without any attempt at optimization. At first, the control field is
410: turned off and the objective functional,%
411: \begin{equation}
412: J\left[ \mathbf{P}_{N}\right] =\left\vert O\left[ \mathbf{P}_{N}\right]
413: -100\%\right\vert ^{2}\text{,}\label{JOb}%
414: \end{equation}
415: is optimized with respect to the coefficients $\left\{ a_{jk}\right\} $ in
416: the $N$ observed operators
417: \begin{equation}
418: \mathbf{P}_{N}=\left( P_{1},\cdots,P_{N}\right) \text{.}%
419: \end{equation}
420: In Eq.(\ref{JOb}) $O\left[ \mathbf{P}_{N}\right] $ is the population yield
421: attained from the observations without the control field. The second column of
422: Table III shows the largest attainable population transfer with different
423: numbers of optimized observations when the control field is off. It has been
424: proved that the QAZE induced by suitable time-dependent measurements can fully
425: transfer population to a target state in the frequent measurement
426: limit\cite{Roy004019}, $\lim_{N\rightarrow\infty}O\left[ \mathbf{P}%
427: _{N}\right] =100\%$. We now introduce a weak control field of the form in Eq.
428: (\ref{E0}) with all of the amplitudes being $0.07$ and phases set at $0.0$.
429: The target time is $T_{f}=200$ fs, and the pulse width in Eq. (\ref{E0}) is
430: $\sigma=30$ fs. This fixed non-optimal control field can only drive $12.93\%$
431: of the population to target state when acting alone (i.e., without
432: observation). The objective is now a functional of both the control field and
433: measured operators,
434: \begin{equation}
435: J\left[ E\left( t\right) ,\mathbf{P}_{N}\right] =\left\vert O\left[
436: E(t),\mathbf{P}_{N}\right] -100\%\right\vert ^{2}\text{,}%
437: \end{equation}
438: but still only the observation operators $\mathbf{P}_{N}$ are optimized. The
439: third column in Table III shows the attained population transfer induced by
440: both the control field and the optimized observations acting together. \ The
441: contribution from the observations acting alone is listed in the fourth
442: column. A high degree of cooperation between the control field and observation
443: is found. For example, for $N=5$ the observations carried out alone produce a
444: yield of $20.46\%$ and the yield from the non-optimal control field alone is
445: $12.93\%$, but the yield from both acting together is $79.22\%$, much larger
446: than their simple summation. Table III indicates that when $N<9$, the presence
447: of the control field is helpful for achieving a higher yield. Further
448: numerical simulations show that, when $N\geq9$, the presence of the control
449: field becomes less helpful, which reflects the strength of observations acting
450: alone as controls. This behavior may be confirmed by an analytical
451: assessment\cite{Pechen06052102, Shuang2006_1} of $O\left[ P_{N}\right] $,
452: which proves that, when $N\geq9$, the maximum population transfer induced by
453: $N$ observations is larger than $80\%$.
454:
455: \subsection{Model 3}
456:
457: Model 3 in Fig. \ref{Fig_Model}(b) is a high symmetry three-level system with
458: the Hamiltonian $H_{0}$ and dipole $\mu$ given by
459: \begin{equation}
460: H_{0}=\left(
461: \begin{array}
462: [c]{ccc}%
463: 1 & 0 & 0\\
464: 0 & 2 & 0\\
465: 0 & 0 & 3
466: \end{array}
467: \right) \text{, }\mu=\left(
468: \begin{array}
469: [c]{ccc}%
470: 0 & 1 & 0\\
471: 1 & 0 & 1\\
472: 0 & 1 & 0
473: \end{array}
474: \right) \text{. }%
475: \end{equation}
476: The system is initially prepared in its ground state $\left\vert
477: 0\right\rangle $, and the objective is to transfer the population to state
478: $\left\vert 1\right\rangle $ at target time $T_{f}=200$ fs. If only a
479: dipole-coupled external field is employed, the high symmetry in $H_{0}$ and
480: $\mu$ implies that the system is not fully controllable, and by inspection at
481: most $50\%$ of the population maybe be transferred to state $\left\vert
482: 1\right\rangle $. This assessment can be made rigorous in the following
483: analysis. It has been proved\cite{Rabitz011} that there is a hidden dynamical
484: symmetry in this system,
485: \begin{equation}
486: \left\vert C_{0}\left( t\right) C_{2}\left( t\right) -\frac{C_{1}%
487: ^{2}\left( t\right) }{2}\right\vert =\left\vert C_{0}\left( 0\right)
488: C_{2}\left( 0\right) -\frac{C_{1}^{2}\left( 0\right) }{2}\right\vert
489: =0\text{,} \label{Ck}%
490: \end{equation}
491: where $C_{k}\left( t\right) \,,$ $k=1,2,3$ are complex coefficients of the
492: system wavefunction%
493: \begin{equation}
494: \psi\left( t\right) =\sum_{k=0}^{2}C_{k}\left( t\right) \left\vert
495: k\right\rangle \text{.}%
496: \end{equation}
497: Rewriting Eq. (\ref{Ck}) in terms of density matrix elements gives%
498: \begin{equation}
499: \rho_{00}\left( t\right) \rho_{22}\left( t\right) =\frac{\rho_{11}%
500: ^{2}\left( t\right) }{4}\text{.} \label{RhoS}%
501: \end{equation}
502: The following inequality based on Eq. (\ref{RhoS}) shows that no more than
503: $50\%$ of the population can be driven from its ground state $\left\vert
504: 0\right\rangle $ to the state $\left\vert 1\right\rangle $%
505: \begin{equation}
506: \rho_{11}\left( t\right) =2\sqrt{\rho_{00}\left( t\right) \rho_{22}\left(
507: t\right) }\leq\rho_{00}\left( t\right) +\rho_{22}\left( t\right)
508: =1-\rho_{11}\left( t\right) \text{.}%
509: \end{equation}
510: To explore if observations can break the $50\%$ yield limit, first a simple
511: instantaneous observation and then a time-dependent continuous observation is
512: applied. The control field is a simple resonant rectangular
513: pulse\cite{Richardson_JLT010501},%
514: \begin{equation}
515: \left\{
516: \begin{array}
517: [c]{c}%
518: E\left( t\right) =A\cos t\text{, }0\leq t\leq T_{f}\text{,}\\
519: E\left( t\right) =0\text{, \ \ \ \ \ \ \ otherwise, \ }%
520: \end{array}
521: \right. \label{EA}%
522: \end{equation}
523: where only the amplitude $A$ is adjusted for optimization.
524:
525: First, an instantaneous observation is performed at the middle of the control
526: $T_{m}=T_{f}/2$. Table IV shows various control yields when different
527: instantaneous observations are carried out, where $P_{k}$ is the population
528: measurement operator in Eq. (\ref{PopM}). The simulation shows that an
529: instantaneous population observation of state $\left\vert 0\right\rangle $ or
530: $\left\vert 2\right\rangle $ can increase the population transfer to the
531: target state $\left\vert 1\right\rangle $, but at the expense of requiring
532: stronger control fields. In contrast, an observation of the target state
533: population is not helpful. This behavior can be explained by the broken
534: dynamical symmetry induced by the observation of state $\left\vert
535: 0\right\rangle $ or $\left\vert 2\right\rangle $, but this outcome will not be
536: the case from observation of state $\left\vert 1\right\rangle $. An analytical
537: treatment\cite{Shuang2006_1} shows that the maximum attainable population
538: transfer to the level $\left\vert 1\right\rangle $ by a coherent field
539: assisted from measuring $P_{0}$ or $P_{2}$ is $68.7\%$, which is closely
540: approximated by the value of $\simeq67\%$ in Table IV.
541:
542: Now consider carrying out time-dependent continuous observations together with
543: a control field $E\left( t\right) $ having the form in Eq. (\ref{EA}), where
544: the density matrix satisfies%
545: \begin{equation}
546: \dot{\rho}=-i\left[ H_{0}-\mu E\left( t\right) ,\rho\right] -\frac{1}%
547: {2}\kappa\left( t\right) \left[ P,\left[ P,\rho\right] \right] \text{.}%
548: \end{equation}
549: Here the observation strength $\kappa\left( t\right) $ is allowed to be
550: time-dependent, and a simple form of $\kappa\left( t\right) $ is adopted as
551: it proved to be sufficient in the control of model 3:
552: \begin{equation}
553: \kappa\left( t\right) =\left\{
554: \begin{array}
555: [c]{l}%
556: \gamma\text{,}\ \text{ }T_{1}<t<T_{2}\text{,}\\
557: 0\text{,}\ \text{ otherwise. \ \ \ \ \ \ \ }%
558: \end{array}
559: \right.
560: \end{equation}
561: In this case the objective functional $J$ is optimized with respect to not
562: only the control field parameter $A$ in Eq. (\ref{EA}), but also the
563: observation strength $\gamma$ and time interval $T_{1}$, $T_{2}$,
564: \begin{equation}
565: J\left[ A,\gamma,T_{1},T_{2}\right] =\left\vert O\left[ A,\gamma
566: ,T_{1},T_{2}\right] -100\%\right\vert ^{2}+\alpha A^{2}\text{.} \label{JMF}%
567: \end{equation}
568: The coefficient $\alpha$ in Eq. (\ref{JMF}) is $0.01$. In the simulation, the
569: observation strength $\gamma$ was optimized over the range from $0.0$ to
570: $5.0$. Table V shows that with the help of the optimized continuous
571: observations of the population in state $\left\vert 0\right\rangle $ or
572: $\left\vert 2\right\rangle $, the control field can induce almost $100\%$
573: population transfer between the initial state $\left\vert 0\right\rangle $ and
574: target state $\left\vert 1\right\rangle $. As expected, observation of the
575: state $\left\vert 1\right\rangle $ is not helpful. Figure \ref{Fig_Symm}(a)
576: shows the state populations when the optimized continuous observation is on
577: state $\left\vert 0\right\rangle $ and Fig. \ref{Fig_Symm}(b) shows the state
578: populations when the optimized continuous observation is on state $\left\vert
579: 2\right\rangle $. The results in Fig. \ref{Fig_Symm} indicate that the
580: observation of $P_{0}$ or $P_{2}$ eliminate population from the state being
581: observed, and the three-level system becomes an effective two-level system in
582: the time interval $T_{1}<t<T_{2}$. This behavior is consistent with the
583: observation acting under the QZE. In both cases $\gamma$ adopts its maximum
584: value of $5.0$ under optimization to evidently take full advantage of the QZE.
585: The simulations with this simple model show that observations can
586: fundamentally alter the effective dynamical structure of a quantum system.
587: This role of observations will be confirmed again in a forthcoming analytical
588: treatment\cite{Shuang2006_1}. Naturally, for more complex systems, additional
589: specially tailored time-dependent observations may be required for optimal
590: impact on the controlled dynamics.
591:
592: \subsection{Model 4}
593:
594: The structure of model 4 is given in Fig. \ref{Fig_Model}(c), and the
595: objective is to transfer population from level $0$ to level $3$. There are two
596: degenerate transitions, $\omega_{11^{\prime}}=\omega_{23}=0.8$, and the other
597: transition frequencies are $\omega_{01}=3.3$, $\omega_{12}=2.6$. The non-zero
598: dipole elements are: $\mu_{01}=0.13$, $\mu_{12}=0.15$, $\mu_{23}=0.23$ and
599: $\mu_{11^{\prime}}=0.21$. The control field has the form of Eq. (\ref{E0})
600: with the resonant three amplitudes and phases subjected to optimization. The
601: target time is $T_{f}=200$ fs, the pulse width in Eq. (\ref{E0}) is
602: $\sigma=30$ fs and the weight coefficient in Eq. (\ref{J0}) is $\alpha=0.01$.
603: The simulation in the first row of Table VI shows that under these conditions,
604: with no observation, the control field can only drive $71.96\%$ of population
605: to the target state, mainly because some population is locked in the undesired
606: state $\left\vert 1^{\prime}\right\rangle $. If a constant continuous
607: observation of the population of state $\left\vert 1^{\prime}\right\rangle $
608: is carried out, the dynamics of model 4 is described by following equation:%
609: \begin{subequations}
610: \begin{align}
611: \dot{\rho} & =-i\left[ H_{0}-\mu E\left( t\right) ,\rho\right] -\frac
612: {1}{2}\kappa\left[ P_{1^{\prime}},\left[ P_{1^{\prime}},\rho\right]
613: \right] \text{,}\label{M4Dyna}\\
614: P_{1^{\prime}} & =\left\vert 1^{\prime}\right\rangle \left\langle 1^{\prime
615: }\right\vert \text{.}%
616: \end{align}
617: Table VI shows that increasing $\kappa$ results in a reduction of the
618: population in state $\left\vert 1^{\prime}\right\rangle $. The phenomena can
619: be explained by the QZE: the strong continuous measurement of state
620: $\left\vert 1^{\prime}\right\rangle $ prohibits population transfer between
621: state $\left\vert 1\right\rangle $ and $\left\vert 1^{\prime}\right\rangle $
622: and avoids population loss to the undesired state $\left\vert 1^{\prime
623: }\right\rangle $, thereby increasing the population in the target state. In
624: all the cases in Table VI the fluence of the control field remains
625: approximately the same at F $\simeq$ 0.57, despite the fact that some of the
626: amplitudes $A_{l}$ in Eq. (\ref{GE}) changed to some degree as $\kappa$
627: varied. Population loss to undesired states is commonly encountered in the
628: practical control of quantum dynamics. This model shows a mechanism to avoid
629: the loss.
630:
631: \section{\bigskip Conclusions}
632:
633: \bigskip This paper discusses observations serving as indirect controls in the
634: manipulation of quantum dynamics. In this context, the field entering the
635: Hamiltonian can be viewed as a direct control. Instantaneous and continuous
636: observations were both considered along with control fields to manipulate
637: population transfer. The simulations show that suitable observations can be
638: very helpful in the manipulation of quantum dynamics. In favorable cases the
639: optimal control field can cooperate with observations to achieve the target
640: more effectively, even when the objective yield is large. In turn, optimal
641: observations can work with an existing or constrained control field to
642: transfer more population from an initial state to a target state. Observations
643: can break dynamical symmetry to increase controllability as well as prohibit
644: transfer of amplitude to undesired states. The QZE and QAZE are the key
645: operational processes associated with the observations to assist the control
646: field to more effectively achieve the target objective. The performance of
647: optimal observations hopefully will become routine with advancing technology,
648: as observations can be powerful tools in the control of quantum dynamics.
649: \end{subequations}
650: \begin{acknowledgments}
651: The authors acknowledge support from the NSF, DARPA and ARO-MURI.
652: \end{acknowledgments}
653:
654: \bibliographystyle{apsrev}
655: \begin{thebibliography}{38}
656: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
657: \expandafter\ifx\csname bibnamefont\endcsname\relax
658: \def\bibnamefont#1{#1}\fi
659: \expandafter\ifx\csname bibfnamefont\endcsname\relax
660: \def\bibfnamefont#1{#1}\fi
661: \expandafter\ifx\csname citenamefont\endcsname\relax
662: \def\citenamefont#1{#1}\fi
663: \expandafter\ifx\csname url\endcsname\relax
664: \def\url#1{\texttt{#1}}\fi
665: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
666: \providecommand{\bibinfo}[2]{#2}
667: \providecommand{\eprint}[2][]{\url{#2}}
668:
669: \bibitem[{\citenamefont{Rice and Zhao}(2000)}]{Rice00}
670: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Rice}} \bibnamefont{and}
671: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zhao}},
672: \emph{\bibinfo{title}{Optical Control of Molecular Dynamics}}
673: (\bibinfo{publisher}{Wiley, New York}, \bibinfo{year}{2000}).
674:
675: \bibitem[{\citenamefont{Rabitz}(2003)}]{Rabitz0364}
676: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
677: \bibinfo{journal}{Theor. Chem. Acc.} \textbf{\bibinfo{volume}{109}},
678: \bibinfo{pages}{64} (\bibinfo{year}{2003}).
679:
680: \bibitem[{\citenamefont{Shapiro and Brumer}(2003)}]{Shapiro03}
681: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Shapiro}} \bibnamefont{and}
682: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Brumer}},
683: \emph{\bibinfo{title}{Principles of the Quantum Control of Molecular
684: Processes}} (\bibinfo{publisher}{John Wiley}, \bibinfo{address}{New York},
685: \bibinfo{year}{2003}).
686:
687: \bibitem[{\citenamefont{Walmsley and Rabitz}(2003)}]{Walmsley0343}
688: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Walmsley}} \bibnamefont{and}
689: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
690: \bibinfo{journal}{Phys. Today} \textbf{\bibinfo{volume}{56}},
691: \bibinfo{pages}{43} (\bibinfo{year}{2003}).
692:
693: \bibitem[{\citenamefont{Brixner
694: et~al.}(2001{\natexlab{a}})\citenamefont{Brixner, Damrauer, and
695: Gerber}}]{Brixner011}
696: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Brixner}},
697: \bibinfo{author}{\bibfnamefont{N.~H.} \bibnamefont{Damrauer}},
698: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gerber}},
699: in \emph{\bibinfo{booktitle}{Advances in Atomic, Molecular, and Optical
700: Physics}}, edited by
701: \bibinfo{editor}{\bibfnamefont{B.}~\bibnamefont{Bederson}} \bibnamefont{and}
702: \bibinfo{editor}{\bibfnamefont{H.}~\bibnamefont{Walther}}
703: (\bibinfo{publisher}{Academic}, \bibinfo{address}{San Diego, CA},
704: \bibinfo{year}{2001}{\natexlab{a}}), vol.~\bibinfo{volume}{46}, pp.
705: \bibinfo{pages}{1--54}.
706:
707: \bibitem[{\citenamefont{Peirce et~al.}(1988)\citenamefont{Peirce, Dahleh, and
708: Rabitz}}]{Rabitz884950}
709: \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Peirce}},
710: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Dahleh}},
711: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
712: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{37}},
713: \bibinfo{pages}{4950} (\bibinfo{year}{1988}).
714:
715: \bibitem[{\citenamefont{Kosloff et~al.}(1989)\citenamefont{Kosloff, Rice,
716: Gaspard, Tersigni, and Tannor}}]{Kosloff89201}
717: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Kosloff}},
718: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Rice}},
719: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Gaspard}},
720: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Tersigni}}, \bibnamefont{and}
721: \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Tannor}},
722: \bibinfo{journal}{Chem. Phys.} \textbf{\bibinfo{volume}{139}},
723: \bibinfo{pages}{201} (\bibinfo{year}{1989}).
724:
725: \bibitem[{\citenamefont{Shapiro and Brumer}(1986)}]{Shapiro864103}
726: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Shapiro}} \bibnamefont{and}
727: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Brumer}},
728: \bibinfo{journal}{J. Chem. Phys} \textbf{\bibinfo{volume}{84}},
729: \bibinfo{pages}{4103} (\bibinfo{year}{1986}).
730:
731: \bibitem[{\citenamefont{Gaubatz et~al.}(1990)\citenamefont{Gaubatz, Rudecki,
732: Schiemann, and Bergmann}}]{Bergmann905363}
733: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Gaubatz}},
734: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Rudecki}},
735: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Schiemann}},
736: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Bergmann}},
737: \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{92}},
738: \bibinfo{pages}{5363} (\bibinfo{year}{1990}).
739:
740: \bibitem[{\citenamefont{Kobrak and Rice}(1998)}]{Rice982885}
741: \bibinfo{author}{\bibfnamefont{M.~N.} \bibnamefont{Kobrak}} \bibnamefont{and}
742: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Rice}},
743: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{57}},
744: \bibinfo{pages}{2885} (\bibinfo{year}{1998}).
745:
746: \bibitem[{\citenamefont{Assion et~al.}(1998)\citenamefont{Assion, Baumert,
747: Bergt, Brixner, Kiefer, Seyfried, Strehle, and Gerber}}]{Gerber98919}
748: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Assion}},
749: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Baumert}},
750: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bergt}},
751: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Brixner}},
752: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Kiefer}},
753: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Seyfried}},
754: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Strehle}}, \bibnamefont{and}
755: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gerber}},
756: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{282}},
757: \bibinfo{pages}{919} (\bibinfo{year}{1998}).
758:
759: \bibitem[{\citenamefont{Bergt et~al.}(1999)\citenamefont{Bergt, Brixner,
760: Kiefer, Strehle, and Gerber}}]{Gerber9910381}
761: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bergt}},
762: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Brixner}},
763: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Kiefer}},
764: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Strehle}}, \bibnamefont{and}
765: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gerber}},
766: \bibinfo{journal}{J. Phys. Chem. A} \textbf{\bibinfo{volume}{103}},
767: \bibinfo{pages}{10381} (\bibinfo{year}{1999}).
768:
769: \bibitem[{\citenamefont{Kunde et~al.}(2000)\citenamefont{Kunde, Baumann, Arlt,
770: Morier-Genoud, Siegner, and Keller}}]{Kunde00924}
771: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kunde}},
772: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Baumann}},
773: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Arlt}},
774: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Morier-Genoud}},
775: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Siegner}}, \bibnamefont{and}
776: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Keller}},
777: \bibinfo{journal}{Appl. Phys. Lett.} \textbf{\bibinfo{volume}{77}},
778: \bibinfo{pages}{924} (\bibinfo{year}{2000}).
779:
780: \bibitem[{\citenamefont{Bartels et~al.}(2000)\citenamefont{Bartels, Backus,
781: Zeek, Misoguti, Vdovin, Christov, Murnane, and Kapteyn}}]{Bartels00164}
782: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bartels}},
783: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Backus}},
784: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zeek}},
785: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Misoguti}},
786: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Vdovin}},
787: \bibinfo{author}{\bibfnamefont{I.~P.} \bibnamefont{Christov}},
788: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Murnane}},
789: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~C.}
790: \bibnamefont{Kapteyn}}, \bibinfo{journal}{Nature}
791: \textbf{\bibinfo{volume}{406}}, \bibinfo{pages}{164} (\bibinfo{year}{2000}).
792:
793: \bibitem[{\citenamefont{Brixner
794: et~al.}(2001{\natexlab{b}})\citenamefont{Brixner, Damrauer, Niklaus, and
795: Gerber}}]{Brixner0157}
796: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Brixner}},
797: \bibinfo{author}{\bibfnamefont{N.~H.} \bibnamefont{Damrauer}},
798: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Niklaus}}, \bibnamefont{and}
799: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gerber}},
800: \bibinfo{journal}{Nature (London)} \textbf{\bibinfo{volume}{414}},
801: \bibinfo{pages}{57} (\bibinfo{year}{2001}{\natexlab{b}}).
802:
803: \bibitem[{\citenamefont{Levis et~al.}(2001)\citenamefont{Levis, Menkir, and
804: Rabitz}}]{Levis01709}
805: \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Levis}},
806: \bibinfo{author}{\bibfnamefont{G.~M.} \bibnamefont{Menkir}},
807: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
808: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{292}},
809: \bibinfo{pages}{709} (\bibinfo{year}{2001}).
810:
811: \bibitem[{\citenamefont{Herek et~al.}(2002)\citenamefont{Herek, Wohlleben,
812: Cogdell, Zeidler, and Motzkus}}]{Herek02533}
813: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Herek}},
814: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Wohlleben}},
815: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cogdell}},
816: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Zeidler}}, \bibnamefont{and}
817: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Motzkus}},
818: \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{417}},
819: \bibinfo{pages}{533} (\bibinfo{year}{2002}).
820:
821: \bibitem[{\citenamefont{Daniel et~al.}(2003)\citenamefont{Daniel, Full,
822: Gonz\'{a}lez, Lupulescu, Manz, Merli, \v{S}. Vajda, and
823: W\"{o}ste}}]{Daniel03536}
824: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Daniel}},
825: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Full}},
826: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Gonz\'{a}lez}},
827: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Lupulescu}},
828: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Manz}},
829: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Merli}},
830: \bibinfo{author}{\bibnamefont{\v{S}. Vajda}}, \bibnamefont{and}
831: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{W\"{o}ste}},
832: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{299}},
833: \bibinfo{pages}{536} (\bibinfo{year}{2003}).
834:
835: \bibitem[{\citenamefont{Judson and Rabitz}(1992)}]{Judson921500}
836: \bibinfo{author}{\bibfnamefont{R.~S.} \bibnamefont{Judson}} \bibnamefont{and}
837: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
838: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{68}},
839: \bibinfo{pages}{1500} (\bibinfo{year}{1992}).
840:
841: \bibitem[{\citenamefont{Gong and Rice}(2004)}]{Rice049984}
842: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Gong}} \bibnamefont{and}
843: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Rice}},
844: \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{120}},
845: \bibinfo{pages}{9984} (\bibinfo{year}{2004}).
846:
847: \bibitem[{\citenamefont{Sugawara}(2005)}]{Sugawara05204115}
848: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sugawara}},
849: \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{123}},
850: \bibinfo{pages}{204115} (\bibinfo{year}{2005}).
851:
852: \bibitem[{\citenamefont{Shuang and Rabitz}(2004)}]{Shuang049270}
853: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Shuang}} \bibnamefont{and}
854: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
855: \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{121}},
856: \bibinfo{pages}{9270} (\bibinfo{year}{2004}).
857:
858: \bibitem[{\citenamefont{Shuang and Rabitz}(2006)}]{Shuang06154105}
859: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Shuang}} \bibnamefont{and}
860: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
861: \bibinfo{journal}{J. Chem. Phys.} \textbf{\bibinfo{volume}{124}},
862: \bibinfo{pages}{154105} (\bibinfo{year}{2006}).
863:
864: \bibitem[{\citenamefont{Mensky}(1993)}]{Mensky1993}
865: \bibinfo{author}{\bibfnamefont{M.~B.} \bibnamefont{Mensky}},
866: \emph{\bibinfo{title}{Continuous Quantum Measurements and Path Integrals}}
867: (\bibinfo{publisher}{IOP Publishing}, \bibinfo{address}{Bristol and
868: Philadelphia}, \bibinfo{year}{1993}).
869:
870: \bibitem[{\citenamefont{Neumann}(1955)}]{VonNeumann1955}
871: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Neumann}},
872: \emph{\bibinfo{title}{Mathematical foundations of quantum mechanics}}
873: (\bibinfo{publisher}{Princeton University Press},
874: \bibinfo{address}{Princeton}, \bibinfo{year}{1955}).
875:
876: \bibitem[{\citenamefont{Feynman}(1948)}]{Feynman48367}
877: \bibinfo{author}{\bibfnamefont{R.~P.} \bibnamefont{Feynman}},
878: \bibinfo{journal}{Rev. Mod. Phys.} \textbf{\bibinfo{volume}{20}},
879: \bibinfo{pages}{367} (\bibinfo{year}{1948}).
880:
881: \bibitem[{\citenamefont{Mensky}(1994)}]{Mensky94159}
882: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Mensky}},
883: \bibinfo{journal}{Phys. Lett. A} \textbf{\bibinfo{volume}{196}},
884: \bibinfo{pages}{159} (\bibinfo{year}{1994}).
885:
886: \bibitem[{\citenamefont{Walls and Milburn}(1994)}]{Walls1994}
887: \bibinfo{author}{\bibfnamefont{D.~F.} \bibnamefont{Walls}} \bibnamefont{and}
888: \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Milburn}},
889: \emph{\bibinfo{title}{Quantum Optics}} (\bibinfo{publisher}{Springer},
890: \bibinfo{address}{Berlin}, \bibinfo{year}{1994}).
891:
892: \bibitem[{\citenamefont{Misra and Sudarshan}(1977)}]{Misra77756}
893: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Misra}} \bibnamefont{and}
894: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Sudarshan}},
895: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{18}},
896: \bibinfo{pages}{756} (\bibinfo{year}{1977}).
897:
898: \bibitem[{\citenamefont{Itano et~al.}(1990)\citenamefont{Itano, Heinzen,
899: Bollinger, and Wineland}}]{Itano902295}
900: \bibinfo{author}{\bibfnamefont{W.~M.} \bibnamefont{Itano}},
901: \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Heinzen}},
902: \bibinfo{author}{\bibfnamefont{J.~J.} \bibnamefont{Bollinger}},
903: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.~J.}
904: \bibnamefont{Wineland}}, \bibinfo{journal}{Phys. Rev. A}
905: \textbf{\bibinfo{volume}{41}}, \bibinfo{pages}{2295} (\bibinfo{year}{1990}).
906:
907: \bibitem[{\citenamefont{Balachandran and Roy}(2000)}]{Roy004019}
908: \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Balachandran}}
909: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Roy}},
910: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{84}},
911: \bibinfo{pages}{4019} (\bibinfo{year}{2000}).
912:
913: \bibitem[{\citenamefont{Lindinger et~al.}(2005)\citenamefont{Lindinger, Weber,
914: Lupulescu, Vetter, Plewicki, Merli, Woste, Bartelt, and
915: Rabitz}}]{Rabitz05013419}
916: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Lindinger}},
917: \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Weber}},
918: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Lupulescu}},
919: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Vetter}},
920: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Plewicki}},
921: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Merli}},
922: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Woste}},
923: \bibinfo{author}{\bibfnamefont{A.~F.} \bibnamefont{Bartelt}},
924: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
925: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{71}},
926: \bibinfo{pages}{013419} (\bibinfo{year}{2005}).
927:
928: \bibitem[{\citenamefont{Xu et~al.}(2006)\citenamefont{Xu, Coello, Lozovoy,
929: Harris, and Dantus}}]{Dantus_OptExpress061030}
930: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Xu}},
931: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Coello}},
932: \bibinfo{author}{\bibfnamefont{V.~V.} \bibnamefont{Lozovoy}},
933: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Harris}},
934: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Dantus}},
935: \bibinfo{journal}{Opt. Express} \textbf{\bibinfo{volume}{14}},
936: \bibinfo{pages}{10939} (\bibinfo{year}{2006}).
937:
938: \bibitem[{\citenamefont{Goldberg}(1989)}]{Goldberg97}
939: \bibinfo{author}{\bibfnamefont{D.~E.} \bibnamefont{Goldberg}},
940: \emph{\bibinfo{title}{Genetic Algorithms in Search, Optimization, and Machine
941: Learning}} (\bibinfo{publisher}{Addison-Wesley}, \bibinfo{address}{Reading,
942: MA}, \bibinfo{year}{1989}).
943:
944: \bibitem[{\citenamefont{Pechen et~al.}(2006)\citenamefont{Pechen, Il'in,
945: Shuang, and Rabitz}}]{Pechen06052102}
946: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Pechen}},
947: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Il'in}},
948: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Shuang}}, \bibnamefont{and}
949: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
950: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{74}},
951: \bibinfo{pages}{052102} (\bibinfo{year}{2006}).
952:
953: \bibitem[{\citenamefont{Shuang et~al.}()\citenamefont{Shuang, Zhou, Pechen, and
954: Rabitz}}]{Shuang2006_1}
955: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Shuang}},
956: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zhou}},
957: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Pechen}}, \bibnamefont{and}
958: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}}, \bibinfo{note}{to
959: be published}.
960:
961: \bibitem[{\citenamefont{Turinici and Rabitz}(2001)}]{Rabitz011}
962: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Turinici}} \bibnamefont{and}
963: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rabitz}},
964: \bibinfo{journal}{Chem. Phys.} \textbf{\bibinfo{volume}{267}},
965: \bibinfo{pages}{1} (\bibinfo{year}{2001}).
966:
967: \bibitem[{\citenamefont{Petropoulos et~al.}(2001)\citenamefont{Petropoulos,
968: Ibsen, Ellis, and Richardson}}]{Richardson_JLT010501}
969: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Petropoulos}},
970: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ibsen}},
971: \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Ellis}}, \bibnamefont{and}
972: \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Richardson}},
973: \bibinfo{journal}{J. Lightwave Technol.} \textbf{\bibinfo{volume}{19}},
974: \bibinfo{pages}{746} (\bibinfo{year}{2001}).
975:
976: \end{thebibliography}
977:
978: \pagebreak
979:
980: \bigskip TABLE I. Optimal control fields fighting against the disturbance of
981: various observations for model 1 with the goal of a high target yield
982: $O_{T}=100\%$.%
983:
984: \begin{tabular}
985: [c]{|c|c|c|c|c|}\hline
986: $A\ ^{a}$ & $O\left[ E\left( t\right) ,A\right] \ ^{b}$(\%) & Tr$\left[
987: \rho\left( T_{m}\right) A\right] \ ^{c}$ & $O\left[ E\left( t\right)
988: ,0\right] \ ^{d}$(\%) & $F\ ^{e}$\\\hline
989: -- & 98.44 & -- & 98.44 & 0.063\\\hline
990: $\mu$ & 92.42 & 0.66 & 94.03 & 0.37\\\hline
991: $H_{0}$ & 85.45 & 3.94 & 85.17 & 1.29\\\hline
992: $P_{0}$ & 97.14 & \textit{0.0037 } & 95.77 & 0.49\\\hline
993: $P_{1}$ & 96.19 & \textit{0.021 } & 93.71 & 0.56\\\hline
994: $P_{2}$ & 93.26 & \textit{0.055 } & 92.98 & 0.77\\\hline
995: $P_{3}$ & 97.64 & \textit{0.0010 } & 97.27 & 0.78\\\hline
996: $P_{4}$ & 96.59 & \textit{0.0032 } & 95.68 & 0.72\\\hline
997: \end{tabular}
998:
999:
1000: $^{a}\ $The operator observed at time $T_{m}=T_{f}/2$. Here $\mu$ is the
1001: dipole; $H_{0}$ is the field-free
1002:
1003: \ \ Hamiltonian; $P_{k}$ is a population projection operator for state
1004: $\left\vert k\right\rangle $, $k=0,\cdots,4$.
1005:
1006: $^{b}\ $Yield from the optimal control field and an instantaneous observation
1007: at time $T_{m}=T_{f}/2$.
1008:
1009: $^{c}\ $Observed value of operator $A$.
1010:
1011: $^{d}\ $Yield arising from the control field without actually performing the
1012: observation, but
1013:
1014: \ \ the control field is determined in the presence of the observation of
1015: operator $A$.
1016:
1017: $^{e}\ $Fluence of the control field.\pagebreak
1018:
1019: TABLE II. Optimal control fields interacting with an observation of the dipole
1020: $\mu$ for model 1 with different objective yields.%
1021:
1022: \begin{tabular}
1023: [c]{|c|c|c|c|c|}\hline
1024: $O_{T}$ $^{a}$(\%) & $O\left[ E\left( t\right) ,\mu\right] ^{b}$(\%) &
1025: $O\left[ E\left( t\right) ,0\right] ^{c}$(\%) & $F^{d}$ & $F_{0}$ $^{e}%
1026: $\\\hline
1027: 10 & 10.00 & 2.03$\times$10$^{-7}$ & 0.0020 & 0.017\\\hline
1028: 20 & 20.03 & 2.74$\times$10$^{-9}$ & 0.00026 & 0.023\\\hline
1029: 30 & 29.86 & 0.0052 & 0.0034 & 0.027\\\hline
1030: 40 & 39.82 & 2.69 & 0.017 & 0.031\\\hline
1031: 50 & 49.73 & 13.87 & 0.027 & 0.034\\\hline
1032: 60 & 59.73 & 40.20 & 0.036 & 0.038\\\hline
1033: 70 & 69.74 & 48.12 & 0.041 & 0.042\\\hline
1034: 80 & 79.18 & 81.80 & 0.31 & 0.046\\\hline
1035: 90 & 88.86 & 89.36 & 0.34 & 0.052\\\hline
1036: 100 & 92.42 & 94.03 & 0.37 & 0.063\\\hline
1037: \end{tabular}
1038:
1039:
1040: $^{a}$ Objective yield in Eq. (\ref{J0}).
1041:
1042: $^{b}$ Yield from an optimal control field and an observation of the dipole
1043: $\mu$ at time $T_{m}=T_{f}/2$.
1044:
1045: $^{c}$ Yield arising from the control field without an observation of the
1046: dipole, but with the
1047:
1048: \ \ control field determined in the presence of an observation of the dipole.
1049:
1050: $^{d}$ Fluence of the control field optimized with the observation present.
1051:
1052: $^{e}$ Fluence of the control field optimized without the observation present.
1053:
1054: \pagebreak
1055:
1056: \bigskip TABLE III. Optimal control of model 2 with a sequence of
1057: instantaneous observations%
1058:
1059: \begin{tabular}
1060: [c]{|c|c|c|c|}\hline
1061: $N$ $^{a}$ & $O\left[ \mathbf{P}\right] ^{b}$(\%) & $O\left[ E\left(
1062: t\right) ,\mathbf{P}\right] ^{c}$(\%) & $O\left[ 0,\mathbf{P}\right] ^{d}%
1063: $(\%)\\\hline
1064: 0 & 0 & 12.93$^{e}$ & 0.00\\\hline
1065: 1 & 50.00 & 56.46 & 11.82\\\hline
1066: 3 & 62.50 & 72.60 & 16.90\\\hline
1067: 5 & 71.04 & 79.22 & 20.46\\\hline
1068: 7 & 73.72 & 80.61 & 18.22\\\hline
1069: 9 & 80.11 & 80.45 & 19.65\\\hline
1070: \end{tabular}
1071:
1072:
1073: $^{a}$ Number of observations $N$ performed at times $T_{k}=\frac{k}{N+1}%
1074: T_{f}$, $k=1,\cdots N$.
1075:
1076: $^{b}$ Yield from the optimal observations without a control field.
1077:
1078: $^{c}$ Yield from the optimal observations in the presence of a non-optimal
1079: control field.
1080:
1081: $^{d}$ Yield from the optimal observations without a control field, but with
1082: the optimal
1083:
1084: \ \ observations determined in the presence of non-optimal control field.
1085:
1086: $^{e}$ The fluence of the non-optimal control field is $F=0.0196$.
1087:
1088: \bigskip\pagebreak
1089:
1090: \bigskip\bigskip TABLE IV. Optimal control of model 3 with various
1091: instantaneous observations at time $T_{m}=T_{f}/2$.%
1092:
1093: \begin{tabular}
1094: [c]{|c|c|c|c|c|}\hline
1095: $P\ ^{a}$ & $O\left[ E\left( t\right) ,P\right] \ ^{b}$(\%) & Tr$\left[
1096: \rho\left( T_{m}\right) P\right] \ ^{c}$ & $O\left[ E\left( t\right)
1097: ,0\right] \ ^{d}$(\%) & $F\ ^{e}$\\\hline
1098: -- & 49.99 & -- & 49.99 & 0.0031\\\hline
1099: $P_{0}$ & 66.90 & 0.068 & 46.04 & 0.76\\\hline
1100: $P_{1}$ & 49.99 & 0.50 & 50.00 & 0.96\\\hline
1101: $P_{2}$ & 66.66 & 0.066 & 46.37 & 0.49\\\hline
1102: \end{tabular}
1103:
1104:
1105: $^{a}$\ The operator observed at time $T_{m}=T_{f}/2$. Here $P_{k}$ is a
1106: population projection operator
1107:
1108: \ \ for state $\left\vert k\right\rangle $, $k=0$, $1$, $2$.
1109:
1110: $^{b,c,d,e}$ Refer to Table I.
1111:
1112: \pagebreak
1113:
1114: TABLE V. Control of model 3 with an optimized continuous observation.%
1115:
1116: \begin{tabular}
1117: [c]{|c|c|c|c|c|c|}\hline
1118: $P\ ^{a}$ & $O\left[ A,\gamma,T_{1},T_{2}\right] \ ^{b}$(\%) & $F\ ^{c}$ &
1119: $\gamma$ & $T_{1}$ & $T_{2}$\\\hline
1120: -- & 49.99 & 0.0031 & -- & -- & --\\\hline
1121: $P_{0}$ & 98.92 & 0.021 & 5.00 & 119.13 & 199.96\\\hline
1122: $P_{1}$ & 49.99 & 0.0086 & 0.0025 & 6.24 & 6.33\\\hline
1123: $P_{2}$ & 99.55 & 0.0037 & 5.00 & 2.55 & 199.24\\\hline
1124: \end{tabular}
1125:
1126:
1127: $^{a}$ The population operator $P_{k}$ is observed between time $T_{1}$ and
1128: $T_{2}$ with the strength $\gamma$.
1129:
1130: \ \ Here $P_{k}$ indicates observation of the population in state $\left\vert
1131: k\right\rangle $, $k=0$, $1$, $2$.
1132:
1133: $^{b}$ Yield from the optimal control field and a continuous observation
1134: between times $T_{1}$ and $T_{2}$ with the strength $\gamma$.
1135:
1136: $^{c}$ Fluence of the control field.
1137:
1138: \bigskip\bigskip\pagebreak
1139:
1140: \bigskip
1141:
1142: TABLE VI. Optimal control of Model 4 with different continuous quantum observations
1143:
1144: \bigskip%
1145: \begin{tabular}
1146: [c]{|c|c|c|}\hline
1147: $\kappa^{\ a}$ & $O\left[ E(t),P_{1^{\prime}}\right] $(\%)$^{b}$ & $\left.
1148: P_{1^{\prime}}\right. ^{c}$\\\hline
1149: 0.00 & 71.96 & 14.22\\\hline
1150: 0.01 & 75.53 & 13.52\\\hline
1151: 0.03 & 80.77 & 12.03\\\hline
1152: 0.05 & 84.32 & 10.50\\\hline
1153: 0.09 & 88.61 & 8.27\\\hline
1154: 0.15 & 91.81 & 6.43\\\hline
1155: 0.20 & 93.28 & 5.33\\\hline
1156: 0.30 & 94.78 & 4.25\\\hline
1157: \end{tabular}
1158:
1159:
1160: $^{a}$ Observation strength of state $\left\vert 1^{\prime}\right\rangle $;
1161: refer to Eq. (\ref{M4Dyna}).
1162:
1163: $^{b}$ Population yield in the target state $\left\vert 3\right\rangle $ from
1164: the optimal control field and continuous
1165:
1166: \ \ observations of the population in state $\left\vert 1^{\prime
1167: }\right\rangle $.
1168:
1169: $^{c}$ Population in the undesired state $\left\vert 1^{\prime}\right\rangle $.
1170:
1171: \bigskip\pagebreak
1172:
1173: Fig\ref{Fig_Model}. \bigskip Three multilevel systems used to investigate the
1174: impact of observations in the optimally controlled quantum dynamics
1175: simulations in Sec. IV. (a) The five-level ladder configuration used for
1176: models 1 and 2. (b) Model 3 with degenerate transition frequencies
1177: $\omega_{01}=\omega_{02}$. (c) Model 4, where the two transition frequencies
1178: $\omega_{11^{\prime}}=\omega_{23}$ are degenerate.
1179:
1180: Fig\ref{Fig_Field}. The optimal control field and its power spectrum for model
1181: 1 without an observation being present. The field is found using the cost
1182: function in Eq. (\ref{J0}) with a high expected yield of $O_{T}=100\%$. The
1183: spectral features are at the system transition frequencies.
1184:
1185: Fig\ref{Fig_Symm}. The population evolution of model 3 driven by an optimal
1186: control field with the help of optimized continuous observations performed
1187: between time $T_{1}$ and $T_{2}$. $P_{k}$ denotes the population in level $k$,
1188: $k=1,2,3$. The observation is on state $\left\vert 0\right\rangle $ in plot
1189: (a) and on state $\left\vert 2\right\rangle $ in plot (b).
1190:
1191: \bigskip\bigskip\pagebreak
1192:
1193: \bigskip%
1194: %TCIMACRO{\FRAME{fhFU}{6.0222in}{4.5164in}{0pt}{\Qcb{F Shuang et al}%
1195: %}{\Qlb{Fig_Model}}{fig1.eps}{\special{ language "Scientific Word";
1196: %type "GRAPHIC"; maintain-aspect-ratio TRUE; display "USEDEF";
1197: %valid_file "F"; width 6.0222in; height 4.5164in; depth 0pt;
1198: %original-width 4.1632in; original-height 3.0329in; cropleft "0";
1199: %croptop "1.0279"; cropright "1"; cropbottom "0";
1200: %filename 'Fig1.eps';file-properties "XNPEU";}}}%
1201: %BeginExpansion
1202: \begin{figure}
1203: [h]
1204: \begin{center}
1205: \includegraphics[
1206: trim=0.000000in 0.000000in 0.000000in -0.084618in,
1207: height=4.5164in,
1208: width=6.0222in
1209: ]%
1210: {Fig1.eps}%
1211: \caption{F Shuang et al}%
1212: \label{Fig_Model}%
1213: \end{center}
1214: \end{figure}
1215: %EndExpansion
1216: \pagebreak%
1217: %TCIMACRO{\FRAME{fhFU}{6.0231in}{3.0115in}{0pt}{\Qcb{F Shuang et al}%
1218: %}{\Qlb{Fig_Field}}{fig2.eps}{\special{ language "Scientific Word";
1219: %type "GRAPHIC"; maintain-aspect-ratio TRUE; display "USEDEF";
1220: %valid_file "F"; width 6.0231in; height 3.0115in; depth 0pt;
1221: %original-width 4.3863in; original-height 2.0851in; cropleft "0";
1222: %croptop "1.0472"; cropright "1"; cropbottom "0";
1223: %filename 'Fig2.eps';file-properties "XNPEU";}}}%
1224: %BeginExpansion
1225: \begin{figure}
1226: [h]
1227: \begin{center}
1228: \includegraphics[
1229: trim=0.000000in 0.000000in 0.000000in -0.098417in,
1230: height=3.0115in,
1231: width=6.0231in
1232: ]%
1233: {Fig2.eps}%
1234: \caption{F Shuang et al}%
1235: \label{Fig_Field}%
1236: \end{center}
1237: \end{figure}
1238: %EndExpansion
1239: \pagebreak\pagebreak%
1240: %TCIMACRO{\FRAME{fhFU}{6.0231in}{4.5164in}{0pt}{\Qcb{F Shuang et al}%
1241: %}{\Qlb{Fig_Symm}}{fig4.eps}{\special{ language "Scientific Word";
1242: %type "GRAPHIC"; maintain-aspect-ratio TRUE; display "USEDEF";
1243: %valid_file "F"; width 6.0231in; height 4.5164in; depth 0pt;
1244: %original-width 3.9825in; original-height 2.1153in; cropleft "0";
1245: %croptop "1.4098"; cropright "1"; cropbottom "0";
1246: %filename 'Fig4.eps';file-properties "XNPEU";}}}%
1247: %BeginExpansion
1248: \begin{figure}
1249: [h]
1250: \begin{center}
1251: \includegraphics[
1252: trim=0.000000in 0.000000in 0.000000in -0.866850in,
1253: height=4.5164in,
1254: width=6.0231in
1255: ]%
1256: {Fig4.eps}%
1257: \caption{F Shuang et al}%
1258: \label{Fig_Symm}%
1259: \end{center}
1260: \end{figure}
1261: %EndExpansion
1262:
1263:
1264:
1265: \end{document}