physics0609074/kpr.tex
1: 
2: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
4: 
5: \bibliographystyle{apsrev}
6: \usepackage{graphicx}
7: \usepackage{dcolumn}
8: \usepackage{bm}
9: 
10: \begin{document}
11: 
12: \title{Selective excitation of metastable atomic states by femto- and attosecond laser pulses}
13: 
14: \author{A.~D. Kondorskiy$^{1,2}$}
15: \author{L.~P. Presnyakov$^{1}$\footnote{Deceased.}}
16: \affiliation{$^1$ P. N. Lebedev Physical Institute, Leninsky pr., 53, Moscow, 119991, Russia \\
17: $^2$ Moscow Institute of Physics and Technology, Institutsky per., 9, Dolgoprudny, Moscow region, 141700, Russia}
18: 
19: \author{Yu. Ralchenko}
20: \affiliation{National Institute of Standards and Technology, Gaithersburg MD 20899-8422, USA}
21: 
22: 
23: \date{\today}
24:  
25: \begin{abstract}
26: 
27: The possibility of achieving highly selective excitation of low metastable states
28: of  hydrogen and helium atoms by using short laser pulses with reasonable
29: parameters  is demonstrated theoretically. Interactions of atoms with the laser
30: field are studied  by solving the close-coupling equations without
31: discretization. The parameters of  laser pulses are calculated using different
32: kinds of optimization procedures. For  the excitation durations of hundreds of
33: femtoseconds direct optimization of the  parameters of one and two laser pulses
34: with Gaussian envelopes is used to introduce  a number of simple schemes of
35: selective excitation. To treat the case of shorter  excitation durations, 
36: optimal control theory is used and the calculated optimal  fields are
37: approximated by sequences of pulses with reasonable shapes. A new way  to
38: achieve selective excitation of metastable atomic states by using sequences of 
39: attosecond pulses is introduced.
40: 
41: \end{abstract}
42: 
43: \pacs{32.80.Qk; 32.80.Rm}
44: 
45: \keywords{Metastable atomic states; Selective excitation; The Optimal Control Theory}
46: 
47: \maketitle
48: 
49: 
50: 
51: \section{\label{sec:intro} Introduction}
52: 
53: Nonlinear laser spectroscopy provides new possibilities to create and study
54: selectively excited states of quantum systems \cite{Letokhov1977,Hansch1977a}.
55: Development of the two-photon excitation technique \cite{Hansch1977b} made it
56: possible to obtain small concentrations of hydrogen in the 2s metastable state.
57: This was sufficient to carry out precise optical measurements of the hyperfine
58: structure of 2s state \cite{Kolachevsky2004} and other relativistic and
59: radiative effects \cite{Fischer2004}. Metastable atoms and atomic ions also play
60: an important role in excitation and charge transfer processes, even in
61: high-temperature laboratory and astrophysical plasmas \cite{JPS1985}. Small
62: concentrations of metastable atoms and atomic ions can significantly affect the
63: radiative spectra of plasmas because of large cross sections of electronic
64: capture to the excited states of highly-charged ions during collisions with
65: the metastable atoms \cite{JPS1985}. This conclusion is based on theoretical and
66: indirect spectroscopic observations. Unfortunately, direct measurements are
67: still very difficult due to the low efficiency of modern techniques for the
68: production of beams of metastable atoms \cite{Gilbody2003}. The above
69: examples demonstrate the importance of the development of effective methods for
70: selective excitation of metastable atomic states.
71: 
72: Recent progress in laser techniques as well as in theoretical understanding of
73: multiphoton processes in atoms makes it possible to introduce novel methods for
74: effective selective excitation of atomic states by using short laser pulses. A
75: number of approaches to control population transfer between atomic and molecular
76: states have been proposed so far. They can be roughly divided into two groups
77: according to the basic principles. The methods from the first group exploit some
78: known physical mechanisms \cite{Allen1987, Shore1990,Melinger1992,Band1997, Bergman1998,Unanyan2001,Gong2004,Sola1999,Teranishi1997,Nagaya2001,Zou2005}, 
79: e.g., it is possible to introduce some \textit{reliable} and \textit{solvable} 
80: model equations to estimate the parameters of the controlling laser field analytically,
81: while the methods from the second group are based on the idea of direct variation 
82: of the laser field to maximize the desired output \cite{Rice2000,Shi1988,
83: Kosloff1989,Ohtsuki1998,ZBR1998}.
84: 
85: The schemes from the first group ultimately rely on the properties of $\pi$-pulses 
86: or on adiabatic properties. The controlling laser field is found as an accurate or 
87: approximate solution of the coupled equations within rotating wave approximation. 
88: These schemes, in turn, can be classified by the number of laser pulses involved.
89: 
90: Population inversion achieved by using a single $\pi$-pulse represents the
91: simplest case \cite{Allen1987, Shore1990}. The frequency of the laser should be
92: in resonance with the transition energy and the control is performed by changing
93: the pulse area. This method can also be used to control multiphoton transitions.
94: In that case, the frequency of the laser pulse should be a fraction of the
95: transition energy so that the energy conservation rules are satisfied. The
96: advantages of this scheme are that it is simple and that the intensity of the
97: controlling laser pulse is relatively small. The main disadvantages are that the
98: efficiency of the scheme is sensitive to the pulse area and that the resonance
99: conditions should be satisfied accurately, especially in the case of multiphoton
100: excitation.
101: 
102: The schemes based on adiabatic rapid passage (ARP) rely on adiabatic properties
103: \cite{Allen1987, Melinger1992, Band1997}. In this case a single chirped laser
104: pulse is used to establish an adiabatic regime so that the complete population
105: transfer occurs as a result of system evolution along the initially populated
106: adiabatic state. The population transfer is not sensitive to the pulse area once
107: an adiabatic regime is established, so the method is quite robust. However, the
108: pulse area must be much larger than in the case of $\pi$-pulses. This becomes
109: critical if the competing ionization or dissociation processes are significant.
110: 
111: Stimulated Raman adiabatic passage (STIRAP) emerged as a very efficient method
112: to provide almost complete population transfer in a three-level system
113: \cite{Bergman1998}. In its simplest form, the scheme involves a two-photon Raman
114: process, in which an interaction with a first (pump) pulse links the initial
115: state with an intermediate state, which in turn interacts via a second (Stokes)
116: pulse with a final target state. An important advantage is that almost no
117: population is placed into the intermediate state, and thus the process is
118: insensitive to any possible decay from that state. In more complicated cases the
119: method is used to create a maximal coherent superposition of several states
120: \cite{Unanyan2001} and complete population transfer via several intermediate
121: states \cite{Gong2004}. The analog of the STIRAP which relies on adiabatic
122: properties is called chirped adiabatic passage by two-photon absorption (CAPTA)
123: \cite{Sola1999}.
124: 
125: Recently an effective scheme based on the idea of periodic chirping was
126: introduced  \cite{Teranishi1997,Nagaya2001}. Within this scheme, a sequence of
127: chirped laser  pulses or a laser with periodic chirping is used to establish
128: multiple crossings  between dressed initial and target levels. Effective laser
129: control is performed  by manipulating the parameters of these crossings and
130: adiabatic phase differences  between the two crossings directly. In its simplest
131: formulation the complete population  transfer can be achieved by using a single
132: quadratically chirped pulse. Because of  interference between the crossings, the
133: necessary pulse area of each pulse is much  smaller compared to the ARP; however
134: the method is still quite robust. These advantages  make the method useful in
135: the field of wave packet control \cite{Zou2005}.  However, the area of the
136: pulse is still larger than in the case of control by  non-chirped pulses.
137: 
138: The methods from the second group are based on maximization of the
139: desired output by a step-by-step adjustment of the controlling field. The main
140: advantage of these methods is that controlling field search is performed
141: explicitly, so that complicated systems can be controlled. These methods
142: can be arranged according to the measure of the controlling parameter space.
143: 
144: In the simplest case, the number of controlling laser pulses and their
145: envelopes are fixed. Each pulse is described by a few parameters: peak
146: intensity, central frequency, chirp rate, central time and full width at half
147: maximum (FWHM). Control is performed by direct optimization of these parameters.
148: By solving the Schr\"{o}dinger equation, the final population of the desired
149: state can be determined as a function of the parameters of the controlling
150: pulses. Thus, well established numerical procedures for maximization of this
151: function of many variables can be used \cite{Press2002}. It is easy to
152: demonstrate that all the schemes discussed above can be reproduced by direct
153: optimization as particular cases. The approach can be considered as an
154: optimization with respect to a finite set of parameters, i.e., in a 
155: multi-dimensional space. 
156: 
157: To design the controlling field without any restrictions on its shape, the
158: optimal control theory (OCT) was developed \cite{Rice2000,Shi1988,
159: Kosloff1989,Ohtsuki1998,ZBR1998}. It is based on the idea that the controlling
160: laser pulse should maximize a certain functional. The basic variational
161: procedure leads to a set of equations for the optimal laser field, which include two
162: Schr\"{o}dinger equations to describe the dynamics starting from the initial and
163: target state wave functions. The optimal laser field is given by the imaginary
164: part of the correlation function of these two wave functions. This system of
165: equations of optimal control must be solved iteratively in general. The reader
166: can find a comprehensive review of OCT in \cite{Rice2000}. The main
167: disadvantage of this method is that in many cases the generated optimal field is
168: hardly possible to realize in an experiment. The approach can be considered as
169: optimization with respect to a continious set of parameters, \textit{i.e.}, in
170: the functional space.
171: 
172: In the case of selective excitation of metastable atomic states by laser
173: pulses,  a multiphoton interaction is the basic mechanism of the process. Such
174: effects as  direct ionization \cite{Delone1993, DiMauro1995}, resonant
175: transitions via   intermediate discrete and continuum states \cite{DiMauro1995,
176: Freeman1987,  Kondorskiy2001, Kondorskiy2002a, Kondorskiy2003}, above threshold
177: ionization  \cite{DiMauro1995, Dionissopoulou1995a, Mercouris1996,
178: Dionissopoulou1995b},  and electron rescattering by the atomic core
179: \cite{Paulus1994, Paulus1998, Corcum1993,  Shafer1993} are essential. An
180: accurate treatment of all these effects requires the  exact solution of the
181: quantum equations to describe the dynamics of the system.  Unfortunately, it is
182: impossible to introduce some \textit{reliable} and  \textit{solvable} model
183: equations to estimate the parameters of the controlling  laser field
184: analytically. Thus, the methods of maximization of desired output  should be
185: used. 
186: 
187: In the present work selective excitations of the metastable 2s state of hydrogen
188: and the singlet metastable 1s2s $^{1}$S state of helium are studied using both
189: the direct optimization of laser parameters and the optimal control theory. To
190: integrate the time-dependent Schr\"{o}dinger equation, the well
191: established close-coupling approach based on the properties of time-dependent
192: integral equations \cite{Kondorskiy2001, Kondorskiy2002a, Kondorskiy2003,
193: Kondorskiy2002b} is used. This approach is the most suitable for this kind of
194: spectroscopic calculation. As electron-atom collisions rapidly destroy  atomic
195: metastable states, a search of the optimal laser parameters is performed under
196: the requirement that not only high final population of desired metastable state
197: should be achieved, but also ionization probability should be small
198: \cite{Kolachevsky2004, Fischer2004}. 
199: 
200: The present paper is organized as follows. In the next section the theoretical 
201: approaches are summarized, and relations between the results of the optimal
202: control  calculations and the recently known schemes are demonstrated. In Sec.
203: \ref{sec:femto}  direct optimization of the parameters of one- and two laser
204: pulses with Gaussian  envelope is used to introduce a number of simple schemes
205: for selective excitation  of H(2s) and He(1s2s $^{1}$S). In Sec. \ref{sec:atto}
206: the optimal control theory  is used to look for other possible controlling laser
207: fields. A new mechanism of  selective excitation by using attosecond pulses is
208: introduced. Sec. \ref{sec:concl}  contains a summary.
209: 
210: 
211: \section{\label{sec:theory} Theory}
212: 
213: \subsection{\label{subsec:cce} Close-coupling approach}
214: 
215: The time-dependent Schr\"{o}dinger equation for an atom in a laser field is written 
216: as
217: 
218: \begin{equation}
219: \left[ i \frac{\partial }{\partial t}-\widehat{H}_{0}+\mathbf{F}(t) \mathbf{d}
220: (\mathbf{r})\right] \left\vert \Psi (t)\right\rangle =0,
221: \label{eq:schr_eq}
222: \end{equation}
223: where $\widehat{H}_{0}$ is the Hamiltonian of the unperturbed atom, 
224: $\mathbf{d}(\mathbf{r})$ is the dipole moment and $\mathbf{F}(t)$ is a laser field. 
225: Atomic units are used throughout the paper, unless otherwise noted.
226: 
227: We employ the close-coupling (CC) method on the basis of the orthogonal and 
228: normalized unperturbed atomic wave functions of the discrete and continuum states 
229: and expand the total wavefunction as
230: \begin{eqnarray}
231: \left\vert \Psi \left( t\right) \right\rangle = \sum\limits_{\nu } a_{\nu}(t) 
232: \left\vert \psi _{\nu }\right\rangle 
233: e^{-iE_{\nu}t}\nonumber \\ 
234: + \sum\limits_{\mu }\int b_{\mu E}(t)\left\vert \psi _{\mu E}\right\rangle 
235: e^{-iEt}dE \label{eq:expan}
236: \end{eqnarray}
237: where $\nu $ and $\mu $ are the indices that represent the integer quantum numbers 
238: of discrete and continuum states respectively. The functions $\left\vert \psi _{\nu }
239: \right\rangle$ and $\left\vert \psi _{\mu E}\right\rangle$ are the corresponding 
240: stationary state wavefunctions, $E_{\nu }$ are the energies of the discrete states, 
241: $E$ stands for the energy of the electron continuum and $a_{\nu}( t)$ and $b_{\mu E}(t)$ 
242: are unknown coefficients. 
243: 
244: The Hermitian system of coupled equations for the coefficients of the discrete and 
245: continuum states follows from substituting the expansion Eq.~(\ref{eq:expan}) into 
246: the time-dependent Schr\"{o}dinger equation Eq.~(\ref{eq:schr_eq}) as
247: \begin{eqnarray}
248: i\frac{da_{\nu }\left( t\right) }{dt} = \mathbf{F}(t)\sum\limits_{\nu ^{\prime}}
249: \mathbf{U}_{\nu \nu ^{\prime }}e^{i\left( E_{\nu }-E_{\nu ^{\prime
250: }}\right) t}a_{\nu ^{\prime }}\left( t\right) \nonumber \\
251:  + \mathbf{F}(t)\sum_{\mu }\int \mathbf{U}_{\nu \mu }(E)e^{i\left( E_{\nu }-E\right) 
252:  t}b_{\mu E}(t)dE, \label{eq:cce1}
253: \end{eqnarray}
254: \begin{eqnarray}
255: i\frac{db_{\mu E}\left( t\right) }{dt} = \mathbf{F}(t)\sum\limits_{\nu } 
256: \mathbf{U}_{\mu \nu }(E)e^{i\left( E-E_{\nu }\right) t}a_{\nu }\left(
257: t\right) \nonumber \\
258: + \mathbf{F}(t)\sum_{\mu ^{\prime }}\int \mathbf{U}_{\mu \mu ^{\prime }}\left( 
259: E,E^{\prime }\right) e^{i\left( E-E^{\prime }\right)
260: t}b_{\mu E^{\prime }}(t)dE^{\prime }. \label{eq:cce2}
261: \end{eqnarray}
262: The matrix elements $\mathbf{U}_{\nu \nu ^{\prime }}$, $\mathbf{U}_{\nu \mu }(E)$ 
263: and $\mathbf{U}_{\mu \mu ^{\prime }}(E,E^{\prime })$ are integrals over 
264: $\mathbf{r}$-space taken with the atomic dipole moment operator.
265: 
266: The first and the second sums in Eq.~(\ref{eq:cce1}) describe the bound-bound
267: and free-bound transitions, respectively. These transitions result in a
268: significant redistribution of the population of discrete states. This in turn
269: strongly affects the ionization process \cite{Kondorskiy2001, Kondorskiy2002a,
270: Kondorskiy2003, Kondorskiy2002b}. The first sum in Eq.~(\ref{eq:cce2}) describes
271: ionization from all the discrete states and the integral term (free-free
272: transitions) describes a multiphoton inverse bremsstrahlung process (within
273: quantum mechanical considerations) or rescattering processes (within
274: quasiclassical considerations). These processes play an important role in
275: formation of the photoelectronic spectra at high energies \cite{Paulus1994,
276: Paulus1998, Corcum1993, Shafer1993}. However, since the free-free matrix
277: elements practically do not affect the discrete state amplitudes $a_{\nu
278: }\left( t\right) $, it is possible to neglect them for the evaluation of $a_{\nu
279: }\left( t\right) $. The transitions neglected in this approximation are the
280: third-order (bound-free-free-bound) ones. The role of the free-free transitions
281: was carefully investigated in Refs. \cite{Dionissopoulou1995a, Mercouris1996,
282: Dionissopoulou1995b, Kondorskiy2002a, Kondorskiy2003}, and this assumption was
283: confirmed to work well.
284: 
285: The close-coupling equations Eqs.~(\ref{eq:cce1}-\ref{eq:cce2}) with the
286: free-free transitions neglected can be solved by discretizing the continuum or
287: by employing the recently developed approach based on the properties of the
288: time-dependent integral equations without any discretization of the continuum
289: \cite{Kondorskiy2001, Kondorskiy2002a, Kondorskiy2003, Kondorskiy2002b}. As the
290: nonresonant interaction between the lower and highly excited discrete states is
291: small, it is possible to adjust the number of discrete states involved in
292: expansion Eq.~(\ref{eq:expan}) so that the results of the calculations do not
293: change. Although, strictly speaking, this procedure does not prove the
294: convergence of the close-coupling approach, it is still widely used in
295: collisional physics \cite{JPS1985, Lebedev1998}.
296: 
297: In the present study we consider excitation and ionization of a single atomic 
298: electron in the field of linearly polarized laser pulses. 
299: 
300: 
301: \subsection{\label{subsec:direct-optim} Methods of direct optimization of 
302: laser parameters}
303: 
304: In the simplest case the controlling field is assumed to be a sequence of 
305: laser pulses and is written as
306: \begin{eqnarray}
307: F(t)&=&\sum_{j=1}^{N}F_{j}\times f\left( \tau _{j},t-t_{j}\right) \nonumber \\
308: &\times& \sin \left[\omega _{j}\left( t-t_{j}\right) +\alpha _{j}
309: \left( t-t_{j}\right) ^{2}\right] \label{eq:gauss_pulses}
310: \end{eqnarray}
311: where $f\left( \tau ,\Delta t\right) $ is a fixed envelope function with FWHM
312: equal to $\tau $ and centered to achieve maximum at time 
313: difference $\Delta t=0$. In the present study we focus on the laser pulses 
314: Eq.~(\ref{eq:gauss_pulses}) with a Gaussian shape:
315: \begin{equation}
316: f\left( \tau ,\Delta t\right) =\exp \left[ -\frac{4\ln 2}{\tau ^{2}}
317: \Delta t^{2}\right]. \label{eq:gauss_shape}
318: \end{equation}
319: 
320: The system is controlled by changing peak amplitudes $F_{j}$, frequencies
321: $\omega _{j}$, chirp rates $\alpha _{j}$, central times $t_{j}$, and FWHM's
322: $\tau _{j}$ of the component pulses. By solving the Schr\"{o}dinger equation one
323: can determine the final population of a selected state as a function of these
324: parameters. To find the maximum of this function, the conjugate gradient search
325: method \cite{Press2002} is used here. The gradient of the final population of
326: the target state with respect to the laser parameters is calculated numerically
327: using the finite-difference approximation. The optimization starts from some
328: initial guess parameters. 
329: 
330: As different laser pulse parameters have different dimensions, the parameters 
331: should be put on a common ground by using dimensionless units, or the
332: optimizations with respect to different types of parameters should be performed
333: separately. In the first case the convergence of the optimization procedure 
334: could depend on the dimensionless units used. Indeed, the efficiency of the
335: selective  excitation by non-chirped pulse strongly depends on whether or not
336: the resonant conditions  are achieved, so that by using common atomic units one
337: should find the maximum  of the function that strongly depends on one group of
338: arguments (frequencies) and only weakly on the other arguments (intensities,
339: durations etc.). This  example demonstrates that proper dimensionless units
340: should be used to ensure  reasonable convergence. Moreover, the derived optimal
341: parameters could  depend on how that units are defined. 
342: 
343: In the present study the optimizations with respect to  different types of
344: parameters are performed separately so that each step of the  optimization
345: procedure should contain a set of optimizations with respect to the  parameters
346: for all pulses that have the same dimension. Generally, some laser  parameters
347: should be linked with each other to achieve a maximum of the desired  output. If
348: the chirping rate is zero, the optimal intensity and duration of each  pulse are
349: related to each other by some pulse area conservation rule. To avoid this 
350: uncertainty, we assume the durations of the pulses to be equal. After the
351: optimal  parameters of the controlling sequence of laser pulses are found, we
352: can adjust  intensities and durations to fit the specifics of the experimental
353: technique.
354: 
355: The order of optimizations performed at each step of the direct search affects
356: the efficiency, convergence and resulting controlling scheme. However, since the
357: laser field, Eq.~(\ref{eq:gauss_pulses}), has only five different types of
358: parameters, the number of possible nonequivalent orders of optimization is
359: limited. Table \ref{tab:table1} presents the correspondence between the
360: controlling schemes generated by the direct optimization method with different
361: orders of optimizations performed at each step and previously known schemes. 
362: Since the last schemes have been established without bound-free transitions 
363: taken into account the continuum states are not included in the 
364: test calculations of Table \ref{tab:table1}. The continuum 
365: states are, however, taken into account at all other calculations, reported in the 
366: present paper.
367: 
368: The durations of all the pulses are assumed to be equal and do not change during 
369: optimization. Initial parameters used in all the calculations are 1 $GW/cm^{2}$ 
370: for intensities of all pulses, frequencies are estimated from the data for the 
371: unperturbed atom, chirping rates are zero for all the pulses, centers of the 
372: pulses coincide in the cases 1-4 and are equally distant with the $2\cdot$FWHM 
373: shift in cases 5 and 6. Our calculations show that number of steps required to 
374: achieve convergence is about one and a half the number of laser pulses used 
375: in the scheme.
376: 
377: \begin{table}
378: \caption{\label{tab:table1} Controlling schemes generated by the direct optimization 
379: method with different orders of optimizations performed at each step.}
380: \begin{ruledtabular}
381: \begin{tabular}{cclc}
382: \begin{tabular}{c} Case \\ No.\end{tabular} & \begin{tabular}{c} Number \\ of pulses \end{tabular} & 
383: \begin{tabular}{c} Order of \\ optimizations \end{tabular} & \begin{tabular}{c} Known \\ particular cases \end{tabular} 
384: \\ \hline \hline
385: 1 & 1 & \begin{tabular}{l} 1. Intensities \\ 2. Frequencies \end{tabular} & $\pi$-pulses 
386: \\ \hline
387: 2 & 1 & \begin{tabular}{l} 1. Intensities \\ 2. Chirping rates \end{tabular} & ARP \\ 
388: \hline
389: 3 & 2 & \begin{tabular}{l} 1. Central times \\ 2. Intensities \\ 3. Frequencies \end{tabular} & 
390: \begin{tabular}{c} Direct excitation \\ or STIRAP \end{tabular} \\ 
391: \hline
392: 4 & 2 & \begin{tabular}{l} 1. Central times \\ 2. Intensities \\ 3. Chirping rates \end{tabular} & 
393: \begin{tabular}{c} Direct excitation \\ or CAPTA \end{tabular} \\ 
394: \hline
395: 5 & Several & \begin{tabular}{l} 1. Frequencies \\ 2. Central times \\ 3. Intensities \end{tabular} & 
396: \begin{tabular}{l} Periodic sweeping \\ of laser intensity \end{tabular}\\ 
397: \hline
398: 6 & Several & \begin{tabular}{l} 1. Chirping rates \\ 2. Central times \\ 3. Intensities \end{tabular} & 
399: \begin{tabular}{c} Periodic sweeping \\ of laser frequency \end{tabular}\\ 
400: \end{tabular}
401: \end{ruledtabular}
402: \end{table}
403: 
404: The procedure of direct optimization of laser parameters can also be used to
405: calculate  intrinsic parameters of the atomic system. For example, by fixing the
406: intensity of the  laser and calculating the optimal frequency of selective
407: excitation of different target  states, the dynamic Stark shifts of these states
408: can be easily estimated.
409: 
410: 
411: \subsection{\label{subsec:oct} Methods of the optimal control theory}
412: 
413: The dynamics of the atom in the laser field is described by the time-dependent 
414: Schr\"{o}dinger equation (\ref{eq:schr_eq}). The \textit{initial} state wave
415: function  $\left\vert \Phi _{i}(t)\right\rangle $\ is specified at time $t=0$.
416: The goal of control  is to design such an external field $\mathbf{F}(t)$\ that
417: the wave packet calculated  with Eq. (\ref{eq:schr_eq}) up to time $t=T$ is
418: close enough to the desired  \textit{target} state wave packet $\left\vert
419: \Phi_{t}(t)\right\rangle $. One of the  most natural and flexible approaches to
420: design such a field is the optimal control  theory \cite{Rice2000}.
421: It is based on the idea that the controlling  laser pulse should maximize a
422: certain functional. The procedure leads to a set of  equations for the optimal
423: laser field, which include two Schr\"{o}dinger equations to  describe the
424: dynamics starting from the initial and target state wave packets. The  optimal
425: laser field is given by the imaginary part of the correlation function of  these
426: two wave packets. This system of equations of optimal control must be solved 
427: iteratively in general starting from some initial guess field.
428: 
429: A number of algorithms to realize this idea have been developed 
430: \cite{Rice2000,Shi1988,Kosloff1989,Ohtsuki1998,ZBR1998}. One of the most effective
431: is the algorithm by Zhu, Botina and Rabits \cite{ZBR1998} (ZBR algorithm), which is 
432: developed to solve the following system of the optimal control equations:
433: \begin{equation}
434: \left[ i\frac{\partial }{\partial t}-\widehat{H}_{0}+\mathbf{F}(t)\mathbf{d}
435: (\mathbf{r})\right] \left\vert \phi (t)\right\rangle =0, \quad 
436: \left\vert \phi(0)\right\rangle =\left\vert \Phi _{i}\right\rangle , \label{eq:zbr-initial}
437: \end{equation}
438: \begin{equation}
439: \left[ i\frac{\partial }{\partial t}-\widehat{H}_{0}+\mathbf{F}(t)\mathbf{d}(\mathbf{r})\right] 
440: \left\vert \chi (t)\right\rangle =0, \quad 
441: \left\vert \chi(T)\right\rangle =\left\vert \Phi _{t}\right\rangle , \label{eq:zbr-target} 
442: \end{equation}
443: \begin{equation}
444: \mathbf{F}(t)=-\frac{1}{\gamma }\textrm{Im}\left[ \left\langle \phi (t)\right. \left\vert 
445: \chi (t)\right\rangle \left\langle \chi (t)\right\vert \mathbf{d}(\mathbf{r})\left\vert 
446: \phi (t)\right\rangle \right]. \label{eq:zbr-field}
447: \end{equation}
448: Here $\gamma$ is a positive parameter chosen to weight the significance of the laser 
449: energy \cite{ZBR1998}. To integrate Eq. (\ref{eq:zbr-initial}) forward in time and Eq. 
450: (\ref{eq:zbr-target}) backward in time, we employ the close-coupling method discussed 
451: in subsection \ref{subsec:cce}.
452: 
453: To launch the ZBR algorithm, an initial field should be specified. In molecular
454: dynamics control, which is the main area of the application for the ZBR
455: algorithm, a zero initial field is sufficient for most cases. However, since in
456: the present study  initial and target states are stationary, a zero initial
457: field cannot be used.  Instead we use a single laser pulse of Gaussian shape as
458: an initial field and  generate different optimal control fields by changing its
459: parameters. Our calculations show that three to five iterations of the ZBR
460: algorithm are sufficient to achieve the convergence.
461: 
462: 
463: \section{\label{sec:femto} Selective excitation of the metastable atomic states 
464: by femtosecond pulses}
465: 
466: In the present section selective excitations of H(2s) and He(1s2s $^{1}$S) by
467: one  and two femtosecond laser pulses of Gaussian shape are studied. The case of
468: chirp  pulses \cite{Allen1987, Melinger1992, Band1997} or periodic chirping 
469: \cite{Teranishi1997,Nagaya2001} is not considered as the pulse area required is
470: larger  than in the case of non-chirped pulses. This in turn significantly
471: increases the  ionization. However, in some cases when the ionization process is
472: suppressed, periodic chirping can be used to improve the
473: robustness of the scheme. These will be discussed in a future publication.
474: 
475: \subsection{\label{subsec:femto-single} Selective excitation of H(2s) 
476: and He(1s2s $^{1}$S) by a single laser pulse}
477: 
478: In the simplest case a two-photon excitation of H(2s) and He(1s2s $^{1}$S) can
479: be achieved  by using a single laser pulse. The process can be affected by
480: changing only three  parameters of the laser: frequency, duration, and
481: intensity. To study this case we calculate  the final populations of the target
482: metastable states and ionization probabilities as  functions of intensity and
483: duration of the laser pulse for two-photon excitation of hydrogen  (wavelength
484: is around 240 nm) and helium (wavelength is around 120 nm). The frequency of the 
485: laser was optimized for each pair of the arguments to compensate level shifts
486: due to the  dynamic Stark effect and to maximize the output.
487: 
488: For the case of selective excitation of H(2s) the maximal values of the target
489: state  population achieved are $18~\%$ to $20~\%$. Unfortunately, the
490: corresponding ionization  probabilities are large, $27~\%$ to $30~\%$. The
491: maximal differences between the target state population and ionization
492: probability are found for the following set of parameters:
493: 
494: \begin{subequations}
495: \label{eq:hyd_single}
496: \begin{equation}
497: \text{FWHM [fs]} = 2.25\times 10^{14} \text{\Large{/}} I\text{ [W/cm}^{2}\text{]},
498: \end{equation}
499: \begin{eqnarray}
500: \omega \text{ [eV]} & = & \text{5.1021 eV} \nonumber \\ & + & 2.34 \times 10^{-15} 
501: \times I\text{ [W/cm}^{2}\text{]}
502: \end{eqnarray}
503: \end{subequations}
504: where $I$ is the pulse intensity. 
505: The target state population and ionization probability are found to be $8.4~\%$
506:  and $4.2~\%$, respectively, for this set of FWHM and $\omega $.
507: Equations~(\ref{eq:hyd_single}) are obtained  by a fit performed for the range
508: of intensities from 1 TW/cm$^2$ to 20 TW/cm$^2$.
509: 
510: The efficiency of the process is low because of significance of the
511: one-photon ionization from the target metastable state. As the photon energy is
512: about 5.10 eV to 5.14 eV (depending on pulse intensity), this process populates
513: the continuum states with low energies of about 1.7 eV, so that the
514: corresponding bound-free transition matrix elements are large.
515: 
516: An opposing situation is observed in the case of selective excitation of He(1s2s $^{1}$S). 
517: The best values of $65~\%$ for the target state population and $26~\%$ for ionization 
518: probability are found for the following parameters:
519: \begin{subequations}
520: \label{eq:he_single}
521: \begin{equation}
522: \text{FWHM [fs]}=7.12\times 10^{15} \text{\Large{/}} I\text{ [W/cm}^{2}\text{]},
523: \end{equation}
524: \begin{eqnarray}
525: \omega \text{ [eV]} & = & \text{10.3075 eV} \nonumber \\ & 
526: + & 1.54 \times 10^{-16} \times I\text{ [W/cm}^{2}\text{]}.
527: \end{eqnarray}
528: \end{subequations}
529: Eqations~(\ref{eq:he_single}) are obtained by a fit performed for the range of 
530: intensities from 50 TW/cm$^2$ to 400 TW/cm$^2$.
531: 
532: The photon energy is about two times higher then in the case of hydrogen (10.3
533: eV to 10.4 eV, depending on pulse intensity). The process is effective.
534: One-photon ionization from the target metastable state populates the continuum
535: states with energies of about 6.3 eV and corresponding bound-free transition
536: matrix elements are small, so that ionization does not undermine the process.
537: 
538: \subsection{\label{subsec:femto-two} Selective excitation of H(2S) by two laser pulses}
539: 
540: \begin{figure}[t]
541: \centering
542: \includegraphics[width=0.9\linewidth]{fig1.eps}
543: \caption[]{Selective excitation of H(2S) by two laser pulses: (top) final populations of 
544: 2s metastable state and ionization probability, (middle) time shift between the ``blue'' 
545: and ``red'' pulses and (bottom) intensities of the pulses given as functions of 
546: frequency of the ``blue'' pulse.}
547: \label{fig:1}
548: \end{figure}
549: 
550: To improve the efficiency of selective excitation of the metastable state of
551: hydrogen,  we increase the number of controlling parameters by introducing two
552: controlling laser  pulses with different frequencies instead of a single pulse.
553: For simplicity, we call  the pulse with lower frequency as ``red'' pulse and the
554: pulse with higher frequency as  ``blue'' pulse. Frequencies of the blue and
555: the red pulses should be linked  to fit the energy difference between the
556: ground and the metastable states. As optimal intensity and duration of each pulse 
557: are approximately related to each other by a pulse envelope area conservation rule, 
558: in the present study we fix durations of the pulses to be equal.
559: 
560: Figure \ref{fig:1} presents the final population of the metastable state and
561: ionization  probability of hydrogen as functions of frequency of the blue
562: pulse for different  durations of the pulses with the frequency of the red pulse,
563: intensities and central  times of both pulses adjusted to maximize the
564: population of the metastable state.
565: 
566: If the frequency of the red pulse becomes lower than the threshold of the
567: one-photon ionization of the 2s metastable state, the efficiency of the
568: selective excitation grows dramatically. This is caused by an abrupt decrease of
569: ionization rate, which makes it possible to apply more intense laser pulses and
570: achieve higher selective excitation without increasing the ionization.
571: Unfortunately, for hydrogen the one-photon ionization threshold for the 2s state
572: is equal to one third of the energy difference between the 1s and 2s states, and
573: the simple case, when the blue pulse is produced from the single red
574: pulse by the second harmonic generation, is still not good enough. Acceptable
575: results, however, can be obtained for the frequencies of the blue pulse
576: slightly higher and frequencies of the red pulse slightly lower than the
577: threshold values. An interesting observation comes from the analysis of the time
578: shift between the two optimized controlling pulses. If the frequencies of the pulses
579: are far from being in resonance with some bound state,  their centers coincide.
580: However, if the frequency of the blue pulse is close to the frequency of 3p
581: $\rightarrow$ 1s transition, the centers of the blue and the red pulses
582: separate. In this case the direct two-photon excitation process transforms into
583: STIRAP \cite{Bergman1998} (Table \ref{tab:table1}, case 3).
584: 
585: Consider three examples in more detail:
586: 
587: \textbf{\textit{Example 1.}} \textit{Final populations: 2s - 59 \%, continuum -
588: 21 \%.  Parameters of the pulses. Freqencies: blue - 7.12786 eV, red -
589: 3.13629 eV;  Intensities: blue - $5.8 \times 10^{12}$ $W/cm^{2}$, red -
590: $1.2 \times 10^{13}$ $W/cm^{2}$;  FWHMs are 40 fs for both pulses; Pulse centers
591: coincide.} The frequency of the  red pulse  is below but close to the
592: threshold of the one-photon ionization of the H(2s) state. Although the 
593: frequencies are adjusted to avoid resonances with highly excited discrete
594: states, 7p and  8p states are  populated at the level of 9 \% and 3 \%
595: respectively. 
596: 
597: \textbf{\textit{Example 2.}} \textit{Final populations: 2s - 92 \%, continuum -
598: 7 \%.  Parameters of the pulses. Freqencies: blue - 9.09943 eV, red -
599: 1.09933 eV;  Intensities: blue - $4.8 \times 10^{12}$ $W/cm^{2}$, red -
600: $1.6 \times 10^{12}$  $W/cm^{2}$; FWHMs: blue - 40 fs, red - 100 fs;
601: Pulse centers coincide.}  In this case durations of the pulses are different and
602: efficiency of the process is  less sensitive to the shifts between the pulse
603: centers.
604: 
605: \textbf{\textit{Example 3.}} \textit{Final populations: 2s - 91 \%, continuum -
606: 8 \%.  Parameters of the pulses. Freqencies: blue - 12.0878 eV, red -
607: 1.88656 eV;  Intensities: blue - $3.7 \times 10^{12}$ $W/cm^{2}$, red -
608: $4.0 \times 10^{11}$  $W/cm^{2}$; FWHMs are 100 fs for both pulses; blue
609: pulse is delayed for 58 fs.}  The frequencies of the pulses are close to be in
610: resonance with 3p state. In this case  the red pulse comes before the
611: blue pulse. This represents the STIRAP process  \cite{Bergman1998}.
612: 
613: 
614: 
615: \section{\label{sec:atto} Selective excitation of the metastable atomic states 
616: by attosecond pulses}
617: 
618: \begin{figure}[t]
619: \centering
620: \includegraphics[width=0.9\linewidth]{fig2.eps}
621: \caption[]{The optimal control field generated for selective excitation of 
622: the 1s2s $^{1}$S state of helium (top), its Fourier transform (middle) and 
623: time variation of the metastable state population and ionization probability 
624: of the controlled atom (bottom). Top and middle: dashed line is the OCT result, 
625: solid line is the fitted sequence of attosecond pulses. Insert: initial guess 
626: field. Bottom: solid line is the population of the 1s2s $^{1}$S state, dashed 
627: line is the ionization probability.}
628: \label{fig:2}
629: \end{figure}
630: 
631: \begin{figure}[t]
632: \centering
633: \includegraphics[width=0.9\linewidth]{fig3.eps}
634: \caption[]{Final populations of the metastable states and ionization probabilities 
635: obtained with the optimal field generated from a single Gaussian laser pulse given 
636: as a functions of the parameters of initial guess pulse. Top line: Results obtained 
637: for selective excitation of H(2s) with the frequency of initial guess pulse is 10.2 
638: eV. Bottom line: Results obtained for selective excitation of H(1s2s $^{1}$S) with 
639: the frequency of initial guess pulse is 10.3 eV.}
640: \label{fig:3}
641: \end{figure}
642: 
643: In the previous sections the controlling laser field was assumed to be composed 
644: of a few femtosecond laser pulses with Gaussian envelopes. To study other 
645: possibilities to achieve selective excitation we use the optimal control theory, 
646: as described in Sec. \ref{subsec:oct}. As the controlling schemes introduced in 
647: the previous section for long durations can hardly be improved, we focus on the 
648: case of short excitation times.
649: 
650: A typical example is shown in Fig. \ref{fig:2}. It presents the optimal control 
651: field generated by the ZBR algorithm for selective excitation of 1s2s $^{1}$S
652: state  of helium (top) and the time variation of population of the target
653: metastable  state and ionization probability during the interaction (bottom).
654: The calculations  are performed starting from the Gaussian pulse of intensity 80
655: TW/cm$^{2}$,  frequency 10.3 eV and FWHM 2.5 fs (full duration is 5 fs), which is
656: shown in the  insert of Fig. \ref{fig:2}. The weight parameter is taken to be
657: $\gamma = 0.1$.  The final population of the target state is $77~\%$ with
658: ionization probability $7~\%$. The convergence is achieved in four iterations.
659: 
660: During optimization the laser field completely transforms into a sequence of 
661: attosecond pulses. The main perturbations finish after 2 fs, so the duration of 
662: active control is also changed. Contrary to the case of controlling femtosecond 
663: pulses, the state populations now change abruptly. Physically, the sequence of 
664: attosecond laser pulses affects the atomic electrons as a series of pushes. Each 
665: push displaces the electrons for a small distance so they can not move far from 
666: the atomic core, and this suppresses ionization.
667: 
668: To simplify the optimal control field obtained with the ZBR algorithm we approximate 
669: it by a sequence of Gaussians:
670: \begin{equation}
671: F(t)=\sum_{j=1}^{N}\frac{A_{j}}{\sqrt{\beta \pi }}\exp \left[ -\beta \left(t-t_{j}\right) 
672: \right]. \label{eq:atto_approx}
673: \end{equation}
674: If the duration of each pulse is short enough, the momentum supplied to the
675: electrons  by each attosecond pulse is given by the pulse area $A_{j}$ and does
676: not depend on  $\beta$. To calculate the parameters ${ A_{j} }$ and ${ t_{j} }$
677: we roughly approximate  the controlling field by sequence (\ref{eq:atto_approx})
678: and use the method described  in Sec. \ref{subsec:direct-optim} to optimize ${
679: A_{j} }$ and ${ t_{j} }$. The resulting  sequence of attosecond pulses is presented
680: in Fig. \ref{fig:2} (top) as a bold line. The  calculation is performed with the width
681: parameter $\beta = 0.04$ (FWHM = 20 as).  The final population of the target
682: state and the ionization probability for that field  are $81~\%$ and $14~\%$,
683: respectively.
684: 
685: Figure \ref{fig:3} presents final populations of the metastable states and ionization 
686: probabilities obtained with the optimal fields generated from femtosecond pulses of 
687: different intensities and durations. Each point of the plot represents an
688: effective controlling sequence of attosecond pulses. One can see that the 
689: results depend on the parameters of the initial guess pulse in a very complicated way. 
690: The parameters for some effective controlling sequences of attosecond pulses 
691: Eq. (\ref{eq:atto_approx}) calculated by optimal fitting of the parameters 
692: ${ A_{j} }$ and ${ t_{j} }$ are presented in the Table \ref{tab:table2}.
693: 
694: Recently, q number of schemes to generate trains of attosecond laser pulses 
695: have been proposed and realized experimentally (see review
696: \cite{Agostini2004}).  The basic idea is the production of a comb of equidistant
697: frequencies in the  spectral domain with controlled relative phases. As a result
698: trains of  x-ray laser pulses of attosecond duration are obtained. The shapes
699: and carrier  frequency can be modified by eliminating certain harmonics. Also,
700: waveforms  containing optical bursts approaching one cycle have been
701: successfully  synthesized by superposing five phase-controlled sidebands
702: \cite{Sokolov2001}.  Trains of non-modulated attosecond pulses could be also
703: generated by laser-plasma interaction in the relativistic regime (see
704: \cite{Pirozhkov2006} and references therein.)
705: 
706: To generate sequence of pulses close to ones shown in Fig. 2, the spectral
707: components  with the frequencies $7.33$ eV, $22.0 \approx 3 \times 7.3$ eV,
708: $29.3\approx 4 \times  7.3$ eV and $44.0 \approx 6 \times 7.3$ eV should be
709: syncronized. Four main harmonics taken from equidistant spectral modes should
710: be superposed. Thus, for principal frequency  of 7.33 eV one should take 1st, 3rd, 4th
711: and 6th harmonics; for principal frequency  of 3.66 eV 2nd, 6th, 8th and 12th
712: harmonics, respectively, and so on.
713: 
714: \begin{table*}
715: \caption{\label{tab:table2} Parameters of some effective controlling sequences 
716: of attosecond pulses Eq. (\ref{eq:atto_approx}) calculated by optimal fitting of 
717: the optimal control fields. Column 1: target atomic state. Colunms 2-4: parameters 
718: of the initial guess femtosecond pulse. Columns 5-6: final target state populations 
719: and ionization probabilities. Colunms 7-8: Parameters of the sequencies of the 
720: attosecond pulses.}
721: \begin{ruledtabular}
722: \begin{tabular}{cccccccc}
723: 
724: \begin{tabular}{c} Target \\ state \end{tabular} & \begin{tabular}{c} $\omega_{ini}$ \\ $[eV]$ \end{tabular} & \begin{tabular}{c} $\text{FWHM}_{ini}$ \\ $[fs]$ \end{tabular} & \begin{tabular}{c} $I_{ini}$ \\ $[\text{W/cm}^{2}]$ \end{tabular} & $P_{target}$ & $P_{ion}$ & \begin{tabular}{c} $t_{j}$ \\ $[fs]$ \end{tabular} & \begin{tabular}{c} $A_{j}$ \\ $[a.u.]$ \end{tabular} \\ \hline \hline
725: 
726: H(2s) & 10.2 & 5 & $1.0 \times 10^{13}$ & 0.67 & 0.31 &
727: \begin{tabular}{l} 0.05, 0.32, 0.76, 1.33, 1.51, \\ 1.85, 2.02, 2.13, 2.36, 2.41, \\ 2.73 \end{tabular} & 
728: \begin{tabular}{l} -0.50, -0.44, -0.40, -0.93, -0.44,  \\ -0.84, -0.83, -0.93, -0.67, 0.46, \\ 0.95 \end{tabular} \\ \hline
729: 
730: H(2s) & 10.2 & 3.5 & $1.5 \times 10^{13}$ & 0.65 & 0.29 &
731: \begin{tabular}{l} 0.05, 0.69, 1.15, 1.90, 2.27, \\ 2.78, 3.03, 3.23, 3.45, 3.50 \end{tabular} & 
732: \begin{tabular}{l} 0.29, 0.47, 0.30, -0.33, -0.31, \\ -0.44, 0.31, -0.46, 0.61, 0.57 \end{tabular} \\ \hline
733: 
734: H(2s) & 10.2 & 4.0 & $2.3 \times 10^{13}$ & 0.63 & 0.29 &
735: \begin{tabular}{l} 0.05, 0.40, 0.45, 0.77, 0.94, \\ 1.19, 1.34, 1.56, 1.86, 1.93, \\ 2.08, 2.30, 2.45 \end{tabular} & 
736: \begin{tabular}{l} 0.44, 0.41, 0.4, 0.57, 0.41, \\ 0.44, 0.33, 0.26, 0.61, 0.58, \\ 0.49, -0.34, -0.81 \end{tabular} \\ \hline
737: 
738: H(2s) & 10.2 & 2.0 & $7.5 \times 10^{13}$ & 0.62 & 0.25 &
739: \begin{tabular}{l} 0.05, 0.35, 0.68, 1.03, 1.22, \\ 1.30, 1.33, 2.0, 2.45  \end{tabular} & 
740: \begin{tabular}{l} -0.09, -0.48, -0.34, -0.29, -0.7, \\ -0.23, -0.31, -0.71, -1.43  \end{tabular} \\ \hline
741: 
742: H(2s) & 5.1 & 5.0 & $1.0 \times 10^{12}$ & 0.70 & 0.20 &
743: \begin{tabular}{l} 0.05, 0.25, 0.80, 1.17, 1.50, \\ 1.55, 1.64, 2.14, 2.25, 2.32, \\ 2.70, 2.76 \end{tabular} & 
744: \begin{tabular}{l} 0.24, -0.43, -0.29, -0.43, -0.44, \\ -0.44, -0.29, -0.52, -0.57, -0.59, \\ 0.47, 0.43 \end{tabular} \\ \hline
745: 
746: H(2s) & 5.1 & 3.5 & $1.0 \times 10^{12}$ & 0.67 & 0.21 &
747: \begin{tabular}{l} 0.05, 0.28, 0.55, 0.71, 1.31, \\ 0.96, 1.09 \end{tabular} & 
748: \begin{tabular}{l} -0.2, 0.13, 0.37, -1.4, -0.57, \\ 0.39, 0.61 \end{tabular} \\ \hline
749: 
750: H(2s) & 5.1 & 2.5 & $1.0 \times 10^{12}$ & 0.65 & 0.25 &
751: \begin{tabular}{l} 0.05, 0.46, 0.87, 1.14, 1.22, \\ 1.42, 1.58, 1.66 \end{tabular} & 
752: \begin{tabular}{l} 0.40, 0.33, 0.19, 0.57, 0.51, \\ 0.52, -0.21, -0.81 \end{tabular} \\ \hline
753: 
754: He(1s2s $^{1}$S) & 10.3 & 5.0 & $8.0 \times 10^{13}$ & 0.81 & 0.14 &
755: \begin{tabular}{l} 0.05, 0.16, 0.25, 0.38, 0.54, \\ 0.71, 0.86, 0.92, 1.06, 1.11, \\ 1.21, 1.45, 1.63, 1.65 \end{tabular} & 
756: \begin{tabular}{l} -0.71, -0.71, -0.73, 0.70, 0.71, \\ 0.71, -0.69, 0.74, -0.69, 0.73, \\ 0.73, -0.71, -0.77, 0.66 \end{tabular} \\ \hline
757: 
758: He(1s2s $^{1}$S) & 10.3 & 5.0 & $5.0 \times 10^{13}$ & 0.75 & 0.10 &
759: \begin{tabular}{l} 0.05, 0.11, 0.23, 0.57, 0.73, \\ 0.89, 0.95, 1.07, 1.16, 1.27, \\ 1.32, 1.44, 1.48, 1.64, 1.67 \end{tabular} & 
760: \begin{tabular}{l} -0.70, -0.66, -0.61, 0.51, -0.54, \\ -0.6, -0.69, 0.86, -0.71, 0.65, \\ -0.76, 0.66, -0.73, -0.71, 0.69 \end{tabular} \\ \hline
761: 
762: He(1s2s $^{1}$S) & 10.3 & 1.0 & $2.0 \times 10^{14}$ & 0.78 & 0.06 &
763: \begin{tabular}{l} 0.05, 0.17, 0.24, 0.36, 0.45, \\ 0.51, 0.63, 0.73, 0.76 \end{tabular} & 
764: \begin{tabular}{l} -0.56, -0.53, -0.89, -0.59, -0.94, \\ 0.81, -0.66, 0.70, -0.79 \end{tabular} \\ \hline
765: 
766: 
767: He(1s2s $^{1}$S) & 10.3 & 6 & $1.0 \times 10^{13}$ & 0.79 & 0.10 &
768: \begin{tabular}{l} 0.05, 0.25, 0.33, 0.41, 0.57, \\ 0.72, 0.78, 0.90, 0.98, 1.05, \\ 1.15 \end{tabular} & 
769: \begin{tabular}{l} 1.0, 0.88, 0.43, 0.36, 0.51, \\ 0.66, 0.71, 0.74, 0.80, -0.78, \\ 0.90 \end{tabular} \\ \hline
770: 
771: 
772: \end{tabular}
773: \end{ruledtabular}
774: \end{table*}
775: 
776: 
777: \section{\label{sec:concl} Conclusions}
778: 
779: In the present study the controlling field search is performed using a well
780: established close-coupling approach for solving the time-dependent
781: Schr\"{o}dinger equation, so that all essential multiphoton effects are treated
782: accurately. It is demonstrated that optimization procedures provide an effective
783: technique to design laser fields for selective excitation of metastable atomic
784: states. Not only do they reproduce the recently developed schemes of laser control
785: as particular cases but also they introduce new ones. 
786: 
787: An efficient selective excitation of H(2s) and He(1s2s $^{1}$S) can be achieved 
788: by using one and two femtosecond laser pulses. Frequencies of the pulses should 
789: be adjusted properly to suppress single-photon ionization from the target
790: metastable  state. Optimal populations of the atomic states and the
791: corresponding parameters  of the laser pulses are calculated as functions of
792: preferable frequencies and durations of the pulses. These make it possible to
793: choose the parameters of the  controlling laser pulses that fit the capabilities
794: of the present experimental techniques.
795: 
796: A new way to achieve selective excitation of metastable atomic states by using 
797: sequences of attosecond pulses is introduced. This is important because of recent 
798: progress in production of attosecond pulses. While in the present study the 
799: parameters of the pulses are calculated using the optimal control theory, the 
800: direct optimization of high harmonics can be used to control the process in the future.
801: 
802: \begin{acknowledgments}
803: 
804: We wish to thank Prof. I. I. Sobel'man, Prof. N. B. Delone, Prof. A. N.
805: Grum-Grzhimailo,  Prof. H. Nakamura, Dr. N. N. Kolachevsky and Dr. A. A. Narits
806: for useful discussions. This work is supported in part by the Russian Foundation 
807: for Basic Research (project 06-02-17089) (A. K.) and the Office of Fusion
808: Energy Sciences of the U.S. Department of Energy (Yu. R.).
809: 
810: \end{acknowledgments}
811: 
812: 
813: 
814: \begin{thebibliography}{0}
815: 
816: \bibitem{Letokhov1977} V. S. Letokhov and V. P. Chebotaev, {\it Nonlinear laser spectroscopy} (Springer-Verlag, Berlin-New, York 1977).
817: 
818: \bibitem{Hansch1977a} T. W. H\"{a}nsch, {\it Nonlinear Spectroscopy}, edited by N. Bloembergen (Amsterdam, New-Holland 1977).
819: 
820: \bibitem{Hansch1977b} T. W. H\"{a}nsch, {\it Laser Spectroscopy III, Springer Series in Optical Sciences 7} edited by J. L. Hall and S. L. Carlsten, (Springer, Berlin-New York 1977).
821: 
822: \bibitem{Kolachevsky2004} N. Kolachevsky, M.Fischer, S. G. Karshenboim and T. W. H\"{a}nsch, Phys. Rev. Lett. {\bf 92}, 033003, (2004).
823: 
824: \bibitem{Fischer2004} M. Fischer {\it et al}, Phys. Rev. Lett. {\bf 92}, 230802, (2004).
825: 
826: \bibitem{JPS1985} R. K. Janev, L. P. Presnyakov and V. P. Shevelko, {\it Physics of Highly Charged Ions} (Springer Verlag, Berlin 1985).
827: 
828: \bibitem{Gilbody2003} H. B. Gilbody, {\it private communication} (2003).
829: 
830: \bibitem{Allen1987} L. Allen and J. H. Eberly, {\it Optical Response and Two-Level Atoms} (Dover, New York, 1987).
831: 
832: \bibitem{Shore1990} B. W. Shore, {\it The Theory of Coherent Atomic Excitation} (Wiley, New York 1990).
833: 
834: \bibitem{Melinger1992} J. S. Melinger, S. R. Gandhi, A. Hariharan, J. X. Tull and W. S. Warren, Phys. Rev. Lett. {\bf 68}, 2000 (1992).
835: 
836: \bibitem{Band1997} Y. B. Band and P. S. Julienne, J. Chem. Phys. {\bf 97}, 9107 (1997).
837: 
838: \bibitem{Bergman1998} K. Bergman, H. Theuer and B. W. Shore, Rev. Mod. Phys. {\bf 70}, 1003 (1998).
839: 
840: \bibitem{Unanyan2001} R. G. Unanyan, B. W. Shore and K. Bergmann, Phys. Rev. A {\bf 63}, 043401 (2001).
841: 
842: \bibitem{Gong2004} J. Gong and S. A. Rice, Phys. Rev. A {\bf 69}, 063410 (2004).
843: 
844: \bibitem{Sola1999} I. R. Sol\'a, V. S. Malinovsky, B. Y. Chang, J. Santamaria and K. Bergmann, Phys. Rev. A. {\bf 59}, 4494 (1999).
845: 
846: \bibitem{Teranishi1997} Y. Teranishi and H. Nakamura, Phys. Rev. Lett. {\bf 81}, 2032 (1998).
847: 
848: \bibitem{Nagaya2001} K. Nagaya Y. Teranishi and H. Nakamura, {\it Advances in Multiphoton Progress and Spectroscopy}, edited by R. J. Gordon and Y. Fujimura (World Scientific, Singapore, 2001), Vol. 14.
849: 
850: \bibitem{Zou2005} S. Zou, A. Kondorskiy, G. Mil'nikov and H. Nakamura, J. Chem. Phys. {\bf 122}, 084112 (2005).
851: 
852: \bibitem{Rice2000} S. A. Rice and M. Zhao, {\it Optical Control of Molecular Dynamics} (John-Willey \& Sons, 2000).
853: 
854: \bibitem{Shi1988} S. Shi, A. Woody and H. Rabitz, J. Chem. Phys. {\bf 88}, 6870 (1988).
855: 
856: \bibitem{Kosloff1989} R. Kosloff, S. Rice, P. Gaspard, S. Tersigni and D. Tannor, Chem. Phys. {\bf 139}, 201 (1989).
857: 
858: \bibitem{Ohtsuki1998} Y. Ohtsuki, H. Kono and Y. Fujimura, J. Chem. Phys. {\bf 109}, 9318 (1998).
859: 
860: \bibitem{ZBR1998} W. Zhu, J. Botina and H. Rabitz, J. Chem. Phys. {\bf 108}, 1953 (1998).
861: 
862: \bibitem{Press2002} W. H. Press, S. A. Teukolsky, W. T. Vetterling and B. P. Flannery, {\it Numerical Recepies in Fortran}; {\it Numerical Recepies in C}; {\it  Numerical Recepies in C++} (Cambridge University Press, 2001-2002).
863: 
864: \bibitem{Delone1993} N. B. Delone, {\it Basis of Interaction of Laser Radiation with Matter} (Gif-Sur-Yvette Cedex-France, Editiones Frontieres 1993).
865: 
866: \bibitem{DiMauro1995} L. F. DiMauro and P. Agostini, Adv. At. Mol. Opt. Phys. {\bf 35}, 79 (1995).
867: 
868: \bibitem{Freeman1987} R.R. Freeman {\it et al}, Phys. Rev. Lett., {\bf 59}, 1092 (1987).
869: 
870: \bibitem{Kondorskiy2001} A. D. Kondorskiy and L. P. Presnyakov, J. Phys. B: At. Mol. Opt. Phys. {\bf 34}, L663 (2001).
871: 
872: \bibitem{Kondorskiy2002a} A. D. Kondorskiy and L. P. Presnyakov, Laser Phys. {\bf 12}, 449 (2002).
873: 
874: \bibitem{Kondorskiy2003} A. D. Kondorskiy and L. P. Presnyakov, Proc. SPIE {\bf 5228}, 394 (2003).
875: 
876: \bibitem{Kondorskiy2002b} A. Kondorskiy, H. Nakamura, Phys. Rev. A {\bf 66}, 053412 (2002).
877: 
878: \bibitem{Dionissopoulou1995a} S. Dionissopoulou, Th. Mercouris, A. Lyras, Y. Komninos and C. A. Nicolaides, Phys. Rev. A {\bf 51}, 3104 (1995).
879: 
880: \bibitem{Mercouris1996} T. Mercouris {\it et al}, J. Phys. B: At. Mol. Opt. Phys. {\bf 29}, L13 (1996).
881: 
882: \bibitem{Dionissopoulou1995b} S. Dionissopoulou, Th. Mercouris, A. Lyras and C. A. Nicolaides, Phys. Rev. A {\bf 55}, 4397 (1997).
883: 
884: \bibitem{Paulus1994} G. G. Paulus {\it et al}, J. Phys. B: At. Mol. Opt. Phys. {\bf 27}, L703, (1994).
885: 
886: \bibitem{Paulus1998} G. G. Paulus {\it et al}, Phys Rev. Lett. {\bf 80}, 484 (1998).
887: 
888: \bibitem{Corcum1993} P. Corkum, Phys. Rev. Lett., {\bf 71}, 1994 (1993).
889: 
890: \bibitem{Shafer1993} K. J. Schafer, Baorui Yang, L. F. DiMauro and K. C. Kulander, Phys. Rev. Lett. {\bf 70}, 1599 (1993).
891: 
892: \bibitem{Lebedev1998} V. S. Lebedev and I. L. Beigman, {\it Physics of Highly Excited Atoms and Ions}, (Springer, Berlin-Heidelberg, 1998).
893: 
894: \bibitem{Agostini2004} P. Agostini and L. F. DiMauro, Rep. Prog. Phys., {\bf 67}, 813 (2004).
895: 
896: \bibitem{Sokolov2001} A. V. Sokolov {\it et al}, Phys. Rev. A, {\bf 63}, 051801 (2001); Phys. Rev. Lett, {\bf 87}, 033402 (2001);
897: 
898: \bibitem{Pirozhkov2006} A. S. Pirozhkov {\it et al}, Physics of Plasmas, {\bf 13}, 013107 (2006).
899: 
900: 
901: 
902: 
903: 
904: 
905: \end{thebibliography}
906: 
907: \end{document}
908: