Theory of selective excitation in stimulated Raman scattering

Similar documents
OBSERVATION AND CONTROL OF MOLECULAR MOTION USING ULTRAFAST LASER PULSES

Dark pulses for resonant two-photon transitions

Resonant uv pump-probe spectroscopy of dipicolinic acid via impulsive excitation

Module 4 : Third order nonlinear optical processes. Lecture 28 : Inelastic Scattering Processes. Objectives

Absorption-Amplification Response with or Without Spontaneously Generated Coherence in a Coherent Four-Level Atomic Medium

9 Atomic Coherence in Three-Level Atoms

Multi-cycle THz pulse generation in poled lithium niobate crystals

Supporting information for the manuscript. Excited state structural evolution during charge-transfer reactions in Betaine-30

Phase control in the vibrational qubit

Pulse shaping for mode-selective ultrafast coherent Raman spectroscopy of highly scattering solids

Spectral phase optimization of femtosecond laser pulses for narrow-band, low-background nonlinear spectroscopy

SUPPLEMENTARY INFORMATION

American Institute of Physics 319

Coherent Nonlinear Spectroscopy: From Femtosecond Dynamics to Control

Control and Characterization of Intramolecular Dynamics with Chirped Femtosecond Three-Pulse Four-Wave Mixing

On the effects of absolute laser phase on the interaction of a pulsed laser with polar versus nonpolar molecules

QUANTUM THEORY OF LIGHT EECS 638/PHYS 542/AP609 FINAL EXAMINATION

Effects of aligning pulse duration on the degree and the slope of nitrogen field-free alignment

An optical frequency comb is generated by use. Internal State Cooling With a Femtosecond Optical Frequency Comb S. MALINOVSKAYA, V. PATEL, T.

PRINCIPLES OF NONLINEAR OPTICAL SPECTROSCOPY

Sfb 658 Colloquium 11 May Part II. Introduction to Two-Photon-Photoemission (2PPE) Spectroscopy. Martin Wolf

Interference effects on the probe absorption in a driven three-level atomic system. by a coherent pumping field

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

Charge and Energy Transfer Dynamits in Molecular Systems

Models for Time-Dependent Phenomena

Energy Level Sets for the Morse Potential

7 Three-level systems

Matthias Lütgens, Frank Friedriszik, and Stefan Lochbrunner* 1 Concentration dependent CARS and Raman spectra of acetic acid in carbon tetrachloride

Models for Time-Dependent Phenomena. I. Laser-matter interaction: atoms II. Laser-matter interaction: molecules III. Model systems and TDDFT

Spectroscopy in frequency and time domains

Supplementary Information

Initial Hydrogen-Bonding Dynamics of. Photoexcited Coumarin in Solution with. Femtosecond Stimulated Raman Spectroscopy

Theoretical Photochemistry WiSe 2016/17

Supplementary Information for

Four-Wave Mixing Techniques Applied to the Investigation of Non-Adiabatic Dynamics in Polyatomic Molecules

The Two Level Atom. E e. E g. { } + r. H A { e e # g g. cos"t{ e g + g e } " = q e r g

Probing and Driving Molecular Dynamics with Femtosecond Pulses

Femtosecond Quantum Control for Quantum Computing and Quantum Networks. Caroline Gollub

Applied Physics 150a: Homework #3

Modern Optical Spectroscopy

(002)(110) (004)(220) (222) (112) (211) (202) (200) * * 2θ (degree)

Vibrational polarization beats in femtosecond coherent anti-stokes Raman spectroscopy: A signature of dissociative pump dump pump wave packet dynamics

Two-stage Rydberg charge exchange in a strong magnetic field

PHOTO-DISSOCIATION OF CO 2 GAS BY USING TWO LASERS

Inhibition of Two-Photon Absorption in a Four-Level Atomic System with Closed-Loop Configuration

Interaction of atoms and few cycle pulses in frequency and time domain

High-Harmonic Generation II

Electromagnetically Induced Transparency (EIT) via Spin Coherences in Semiconductor

VIC Effect and Phase-Dependent Optical Properties of Five-Level K-Type Atoms Interacting with Coherent Laser Fields

Performance Limits of Delay Lines Based on "Slow" Light. Robert W. Boyd

Mutual transparency of coherent laser beams through a terahertz-field-driven quantum well

1 Mathematical description of ultrashort laser pulses

arxiv:quant-ph/ v1 29 Apr 2003

As a partial differential equation, the Helmholtz equation does not lend itself easily to analytical

Linear pulse propagation

Instantaneous nonvertical electronic transitions with shaped femtosecond laser pulses: Is it possible?

Andrei V. Pisliakov, Tomáš Manal, and Graham R. Fleming

Short-pulse photoassociation in rubidium below the D 1 line

Coherence and optical electron spin rotation in a quantum dot. Sophia Economou NRL. L. J. Sham, UCSD R-B Liu, CUHK Duncan Steel + students, U Michigan

The Phase of a Bose-Einstein Condensate by the Interference of Matter Waves. W. H. Kuan and T. F. Jiang

Correlated Emission Laser, Quenching Of Spontaneous Noise and Coupled Pendulum Analogy

B2.III Revision notes: quantum physics

Nonlinear Optics. Second Editio n. Robert W. Boyd

Strong Field Quantum Control. CAMOS Spring Meeting 2012 o

Molecular spectroscopy

Supplementary Figure 1 Schematics of an optical pulse in a nonlinear medium. A Gaussian optical pulse propagates along z-axis in a nonlinear medium

Design strategies for pulse sequences in multidimensional optical spectroscopies

Electron and nuclear dynamics of molecular clusters in ultraintense laser fields. II. Electron dynamics and outer ionization of the nanoplasma

Towards using molecular ions as qubits: Femtosecond control of molecular fragmentation with multiple knobs

Atom assisted cavity cooling of a micromechanical oscillator in the unresolved sideband regime

Investigations of ultrafast nuclear response induced by resonant and nonresonant laser pulses

Theoretical Photochemistry WiSe 2017/18

5.74 Introductory Quantum Mechanics II

Molecular orientation via a dynamically induced pulse-train: Wave packet dynamics of NaI in a static electric field

PHYSICAL REVIEW A 70, (2004)

Zero-point cooling and low heating of trapped 111 Cd + ions

Wavelength scaling of high-order harmonic yield from an optically prepared excited state atom

SUPPLEMENTARY INFORMATION

FEMTOSECOND MID-INFRARED SPECTROSCOPY OF HYDROGEN-BONDED LIQUIDS

Investigations of amplitude and phase excitation profiles in femtosecond coherence spectroscopy

van Quantum tot Molecuul

Model Answer (Paper code: AR-7112) M. Sc. (Physics) IV Semester Paper I: Laser Physics and Spectroscopy

Supplementary Figure 1 Level structure of a doubly charged QDM (a) PL bias map acquired under 90 nw non-resonant excitation at 860 nm.

Part II Course Content. Outline Lecture 9. Frequency Correlations & Lineshapes. Nonlinear Spectroscopic Methods

Survey on Laser Spectroscopic Techniques for Condensed Matter

Surface Plasmon Amplification by Stimulated Emission of Radiation. By: Jonathan Massey-Allard Graham Zell Justin Lau

Optimization of dynamic molecular alignment and orientation by a phase-shaped femtosecond pulse

FEEDBACK DESIGN OF CONTROL ALGORITHMS FOR DISSOCIATION OF DIATOMIC MOLECULES

doi: /PhysRevLett

Insights on Interfacial Structure, Dynamics and. Proton Transfer from Ultrafast Vibrational Sum. Frequency Generation Spectroscopy of the

pulses. Sec. III contains the simulated results of the interaction process and their analysis, followed by conclusions in Sec. IV.

AMO physics with LCLS

Fundamentals of Spectroscopy for Optical Remote Sensing. Course Outline 2009

Electronic-Resonance-Enhanced (ERE) Coherent Anti-Stokes Raman Scattering (CARS) Spectroscopy of Nitric Oxide. and

Femtochemistry. Mark D. Ellison Department of Chemistry Wittenberg University Springfield, OH

Nonadiabatic dynamics and coherent control of nonequilibrium superconductors

Two-photon nonlinearity in general cavity QED systems

High-Speed Quadratic Electrooptic Nonlinearity in dc-biased InP

Transit time broadening contribution to the linear evanescent susceptibility

Dave S. Walker and Geraldine L. Richmond*

Transcription:

Theory of selective excitation in stimulated Raman scattering S. A. Malinovskaya, P. H. Bucksbaum, and P. R. Berman Michigan Center for Theoretical Physics, FOCUS Center, and Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA Received 11 June 003; published 8 January 004 A semiclassical model is used to investigate the possibility of selectively exciting one of two closely spaced, uncoupled Raman transitions. The duration of the intense pump pulse that creates the Raman coherence is shorter than the vibrational period of a molecule impulsive regime of interaction. Pulse shapes are found that provide either enhancement or suppression of particular vibrational excitations. DOI: 10.1103/PhysRevA.69.013801 PACS number s : 4.65.Dr, 4.50.Hz, 33.0.Fb, 8.53.Kp I. INTRODUCTION Studies of molecular dynamics accompanying excitation, ionization, or dissociation give rise to questions concerning the control of specific motion in molecules. Control using ultrafast lasers and adaptive learning algorithms, e.g., genetic-type algorithms 1 3, allows one to optimize feedback and achieve predetermined goals. Series of successful experiments implementing this technique offer proof that it is possible to exercise quantum coherent control over a range of processes. Two such experiments employ stimulated Raman scattering in liquid methanol 3 and gas phase carbon dioxide 4. In the former case a liquid medium is excited with a 100 fs pulse, which is equivalent to nonimpulsive Raman excitation and, therefore, makes propagation effects important. In the latter case the selective Raman excitation of vibrational modes is achieved with a pulse of duration within a typical molecular vibrational period. In this case the seed for the Stokes transitions is within the pump pulse. In the experiment with methanol the problem of intramolecular couplings of vibrational modes is discussed using a comparative analysis with experiments on benzene and deuterated benzene molecules. The controlling optical field may be constructed using coherent quantum control theory, see, e.g., Refs. 5,6. An alternative approach involves a search for an analytical pulse function. In Refs. 7,8 an antisymmetric phase function is proposed, which inverts the sign of the electric field at a given energy, inducing constructive interference of the off-resonant components of the spectrum and therefore maximizing the transition amplitude. Selectivity of the two Raman levels of pyridine, both of which lie within the bandwidth of the excitation pulses is achieved in CARS signals by positioning the phase step at each of the two peaks 9. In Ref. 10 the use of adaptive techniques in FAST CARS is explored for an identification of related biological objects such as bacterial spores. The Raman excitation of a single vibrational mode having frequency normally requires a pulse with 1, where is the pulse duration. Optimal excitation may occur if the intensity is modulated at a frequency that corresponds to the transition frequency. For any two vibrational modes 1, within the pulse bandwidth, ( 1 ) 1, the spectral width is too large to resolve the lines. In the present paper we propose a method for achieving selective excitation using broadband pulses. Applicability of the method is discussed for the case of many transitions within the pulse frequency region. In the frequency domain we introduce an intensity envelope that vanishes for a modulation frequency of the pulse equal to the frequency of vibration that we would like to suppress. The Fourier-transformed field, when applied to a molecular system, provides negligible excitation of that particular frequency together with significant excitation of another frequency. This picture is valid for weak fields. In strong fields the effect is not obvious and numerical analysis reveals that, for some intensities and the designed pulse shape, it is still possible to optimize one transition versus another. The paper is organized as follows. In the second section the problem is formulated based on a model of two, twolevel systems. An analytical function for the intensity envelope is proposed and the equations of motion for the state amplitudes are obtained. The third section contains results of numerical calculations that determine the field parameters that lead to selective excitation. The paper ends with a short summary. II. BASIC FORMALISM In our model a molecule is described by two, two-level quantum systems, each representing a normal Raman-active vibrational mode. The molecular medium is represented by an ensemble of such systems with no relaxation processes taken into account. Each two-level system interacts with an intense off-resonant femtosecond pulse that initiates stimulated Raman scattering via an off-resonant interaction with a virtual state. The duration of this pump pulse is shorter than a typical vibrational period of a molecule. In this case the frequencies of both two-level systems are within the bandwidth of the pulse and the Stokes component of the field is supplied by the same pulse. The coherently excited ensemble is analyzed by a weak probe pulse, not considered in this work, applied after a time delay shorter than the coherence time. The goal of the present work is to determine a pulse shape that provides selectivity for the excitation of one of the two-level systems. A semiclassical model of laser-molecule interactions is used. The model is represented schematically in Fig. 1 where 1 is the frequency of the first two-level system and 43 that of the second system. Initially only lower levels 1 and 1050-947/004/69 1 /013801 5 /$.50 69 013801-1 004 The American Physical Society

MALINOVSKAYA, BUCKSBAUM, AND BERMAN 3 of both two-level systems are populated, and the populations of these levels are equal. The time evolution of two, two-level systems is described in terms of probability amplitudes, which are written in the interaction representation 4 ȧ j i j 4 j 1 * j e ( j j ) t a j, j jbe p0 t. 1 Here j is the frequency of a single level, such that, e.g., ( 1 ) 1, j is a Rabi frequency, jb is a dipole moment matrix element, E p0 (t) is the pulse envelope, and is the detuning of the frequency of the pulse from the frequency of the virtual state b. Note that the pulse envelope E p0 (t) is the same for all transitions. The Rabi frequencies may differ owing to different dipole moment matrix elements. The system of coupled differential equations 1 is derived from the time-dependent Schrödinger equation with Hamiltonian 11 : H 1 0 0 0 1 cos p t 0 0 0 cos p t 0 0 3 0 3 cos p t, 0 0 0 4 4 cos p t 1 *cos p t *cos p t 3 *cos p t 4 *cos p t E b where p is the laser field carrier frequency. By adiabatically eliminating state b within the rotating wave approximation, we arrive at Eqs. 1. In this work we discuss the case of uncoupled two-level systems such that the probability for the population flow from one system to another via the external field is zero. Then Eqs. 1 are represented by two independent systems of coupled differential equations with two variables. Coherent excitation of a molecular medium induces a vibrational coherence 1 or 34. When a probe field is applied on the 1-b transition, the coherence 1 serves as a source for the generation of a field on the -b transition. Thus it is of prime interest to calculate 1 and 34. The goal of this paper is to choose a pump pulse such that 1 is suppressed while 34 is enhanced. We propose an analytical function for the intensity envelope which is included in the dynamical equations 1 for the probability amplitudes. It is easiest to choose this function in the frequency domain. To suppress excitation at frequency 1 and enhance excitation at frequency 43, we choose Ĩ I 0 e ( 43 ) T 1 e ( 1 ) T 1, where T and T 1 are free parameters. When the modulation frequency of the pulse is equal to 43, the intensity approaches its maximum I 0 for a sufficiently large parameter T 1, Ĩ 43 I 0 1 e T 1, 43 1. For equal to 1 the intensity is zero, Ĩ( 1 ) 0. The intensity envelope as a function of frequency is drawn in Fig.. The frequencies of the two-level systems are 1 1 and 43 1.1 in frequency units of 1. The intensity of the field 3 4 FIG. 1. Schematic picture of a model system consisting of two, two-level systems having frequencies 1 and 43. Initially, the lower levels are populated evenly. The uncoupled transitions are driven by an off-resonant femtosecond pulse. FIG.. Intensity spectral profile as a function of frequency for T 1 10,5,3 dashed, dotted, and solid lines. In the inset the intensity envelope as a function of time is presented for T 1 10,3. All frequencies are in units of 1 and times in units of 1 1. 013801-

THEORY OF SELECTIVE EXCITATION IN... at 43 and frequency region near 1 over which Ĩ( ) is approximately zero depend on T 1. The larger the T 1, the greater is the selectivity for suppressing the 1 frequency. The inverse Fourier transform of the spectral density 3 is a complex function. To arrive at a physically acceptable temporal pulse function, we take the real part of the inverse Fourier transform, given by I t I 0 T 1 e t /4T cos 43 t 1 e T [1 (T / )] (t /4 ) cos 1 T t, where T T 1. The expression for the field 5 is inserted in Eqs. 1 for the calculation of the probability amplitudes. The solution of Eqs. 1 in the limit of a weak field using perturbation theory is a 4 i 4b 3b * I t e i 43 t dt 4 i 4b 3b * 4 Ĩ 43 i 4b 3b * 4 I 0 1 e T1, a i b 1b * I t e i 1 t dt e T ( 43 1 ) 4 e T e ( 1 ) 0. The Fourier transform of the function I(t) represented in Eq. 5 is not identical to Eq. 3, since we took the real part of the Fourier transform to arrive at Eq. 5. It now contains counter-rotating terms, which are small for the chosen pulse shape. Thus, by construction, we have totally suppressed the 1- transition in the weak-field limit. On the other hand, the excitation of the 3-4 transition is still weak owing to the perturbative nature of the solution. Polyatomic molecules often possess several or many Raman-active modes with frequencies close enough to be within the bandwidth of the pulse. In order to enhance a single vibration and suppress other vibrations the function for the pulse may be constructed as a product of several terms, 5 6 Ĩ I 0 e ( 43 ) T j 1 e ( j ) T 1. 7 In a weak-field regime a pulse of such a shape drives one transition with predetermined frequency and suppresses other known transitions in a fixed frequency range. In strong fields the picture is more complicated. With our choice of pulse shape, we can examine the manner in which the selectivity depends on field intensity. It might be possible to construct alternative analytical functions FIG. 3. Coherence and populations of upper levels of the 3-4 and 1- two-level systems as a function of T 1 for 1 1, 43 1.1, and I 0 /8. that provide selective excitation regardless of the field intensity, but this possibility is not addressed in this work. Direct optimization of the pulse shape is certainly one method that can be pursued. However, it is often difficult to interpret a specific pulse shape in terms of the physical parameters associated with the system. III. NUMERICAL RESULTS In this section we discuss the results of numerical calculations based on the solution of Eqs. 1. For the pulse intensity I(t) given by Eq. 5, this numerical solution, obtained using the Runge-Kutta method 1, is exact, provided the excited-state amplitude a b can be eliminated adiabatically. No rotating wave approximation is imposed for the Raman transitions between states 1 and or 3 and 4. Parameters for the system are taken from the experimental data on impulsive excitation of vibrational modes in the molecular gas CO 3. InCO the frequencies of two selectively excited Raman modes are 36.8 and 4 THz. The full width at half maximum of the applied intense pulse is taken equal to 18 THz. In our calculations the frequency 1 is set equal to unity; in these units the frequency 43 is equal to 1.1. From experimental data, we estimate that the parameter T is about equal to 3, in frequency units of 1 1. The intensity of the field is determined by the parameter I 0. The parameter T 1 is related to the width of the spectral dip in Ĩ( ) centered at frequency 1. Although a value of T 1 T would provide the best selectivity, the choice for the parameter T 1 is strongly restricted by the requirement that the duration of the applied pulse be within a typical molecular vibrational period. It turns out that even for such values of T 1, it is possible to selectively excite one transition. We calculate the population distribution and the coherence of the excited and suppressed two-level systems as a function of T 1 as shown in Fig. 3. Bold solid and dotted lines depict the absolute value of the coherence 34 of the excited system and 1 of the suppressed system. Populations of the upper levels of both systems are shown by bold dashed and 013801-3

MALINOVSKAYA, BUCKSBAUM, AND BERMAN FIG. 4. Intensity dependence of the coherences 1 and 34 and upper-state populations and 44. Maximum coherence of the 3-4 system and negligibly small coherence of the 1- system are observed for I 0 1.4 and.08 in the intensity region shown with T 3 and T 1 3. FIG. 5. Time evolution of 1, 34,, and 44 for T 3, T 1 3, and I 0 1.4. The pulse intensity I(t), in arbitrary units, is also shown in the figure. dot-dashed lines. The intensity of the field I 0 is /8, which corresponds to a weak field in our calculations but not to the perturbative regime. For the value T 1 13 the population of levels of the 3-4 system is 0.5, and the coherence is optimal for the given intensity of the field and dipole moments ( i 1). The duration of the laser pulse corresponding to this value of the parameter T 1 is about 00 fs which does not satisfy the necessary requirements on pulse duration. According to Fig. 3 for smaller values of the parameter T 1, corresponding to shorter pulses, the coherence of the 3-4 system is significantly reduced with a simultaneous increase of the coherence of the 1- system. Optimal values for the coherence of both the 1- and 3-4 systems were found for T 1 T through a search over different intensities of the field. In Fig. 4 the coherence is plotted as a function of the intensity of the field for parameters T 3 and T 1 3. Coherence 34 of the 3-4 system is represented by a bold solid line and coherence 1 of the 1- system by a bold dashed line; thin lines show populations of the upper levels of both two-level systems. For the intense fields, coherence of the excited and suppressed systems possess somewhat chaotic structure. Several values of the intensity, e.g., I 0 and I 0 1.75 give rather low coherence of the 3-4 system but maximum coherence of the 1- system. A desired solution for maximum coherence of the excited 3-4 system is achieved for the intensity coefficient I 0.08. This is the result of redistribution of population within that two-level system; half of population is transferred to the upper level providing maximum coherence. The corresponding coherence of the 1- system at this intensity is nearly zero, where most population remains in the lower level. For the intensity I 0 1.4 the picture is similar, however, the low coherence of the 1- system is due to nearly complete population transfer to the upper level. The goal of control of the coherence of two uncoupled two-level systems is achieved with a pulse shape possessing a broad spectral dip at the suppressed frequency and a suitably chosen intensity of the field. This technique allows one to use pulses of duration T to selectively excite transitions having frequency separations 1/T. Our results should not be taken to imply that one can spectroscopically determine frequencies to better than the inverse temporal width of the pulse. On the other hand, if the frequencies are known from previous measurements, it is possible to suppress one transition and enhance the other by the method outlined above. Had we taken a frequency profile centered at 43 with T 3, the curves for the coherence and populations of the excited and suppressed systems would differ qualitatively from those shown in Fig. 4. The desired selectivity could not be achieved. The time dependence of the coherence, populations, and the field is shown in Fig. 5 for T 3,T 1 3, and I 0 1.4. The pulse duration is about 50 fs thin solid line. It induces oscillations in the population distribution thin lines which lead to oscillations of the coherence of the two-level systems bold lines. At long times the coherence and populations of levels achieve stationary values. We have carried out some preliminary calculations, including the coupling between vibrational modes via an external field used as a control mechanism. If the direct excitation of a particular vibrational mode is weak owing to a weak oscillator strength, the excitation may be enhanced through the coupling to the other Raman-active vibrational modes. Details will be published in Ref. 13. IV. SUMMARY We have presented a semiclassical description of stimulated Raman scattering involving the selective excitation of one of two closely spaced vibrational modes. The dynamics is described in terms of probability amplitudes and depends on the shape of a femtosecond laser pulse where duration is shorter than a typical molecular vibrational period. We propose an analytical function for the shape of the intensity envelope of the pulse that allows for the selective excitation 013801-4

THEORY OF SELECTIVE EXCITATION IN... of a predetermined vibrational motion with simultaneous suppression of an unfavorable one. The pulse leads to maximum coherence of a desired vibrational transition and consequently to maximum gain into Raman side bands when a probe pulse is applied. This pulse function may be used as an initial guess for the control of bond excitation in chemistry as well as a fit function within an adaptive learning algorithm. ACKNOWLEDGMENTS The authors acknowledge financial support from the National Science Foundation Grant Nos. PHY-9987916 and PHY-044841 through the Center for Frontiers in Optical Coherent and Ultrafast Science FOCUS and the U. S. Office of Army Research under Grant No. DAAD19-00-1-041. 1 R.S. Judson and H. Rabitz, Phys. Rev. Lett. 68, 1500 199. T.C. Weinacht, J.L. White, and P.H. Bucksbaum, J. Phys. Chem. A 103, 10166 1999. 3 B.J. Pearson, J.L. White, T.C. Weinacht, and P.H. Bucksbaum, Phys. Rev. A 63, 06341 001. 4 T.C. Weinacht, R. Bartels, S. Backus, P.H. Backsbaum, B. Pearson, J.M. Geremia, H. Rabitz, H.C. Kapteyn, and M.M. Murnane, Chem. Phys. Lett. 344, 333 001. 5 S. Shi, A. Woody, and H. Rabitz, J. Chem. Phys. 88, 6870 1988. 6 S. Shi and H. Rabitz, Chem. Phys. 139, 185 1989. 7 N. Dudovich, D. Oron, and Y. Silberberg, Phys. Rev. Lett. 88, 13004 00. 8 D. Oron, N. Dudovich, D. Yelin, and Y. Silberberg, Phys. Rev. Lett. 88, 063004 00. 9 D. Oron, N. Dudovich, D. Yelin, and Y. Silberberg, Phys. Rev. A 65, 043408 00. 10 M.O. Scully, G.W. Kattawar, R.P. Lucht, T. Opatrny, H. Pilloff, A. Rebane, A.V. Socolov, and M.S. Zubairy, Proc. Natl. Acad. Sci. U.S.A. 99, 10994 00. 11 M.O. Scully and M.S. Zubairy, Quantum Optics Cambridge University Press, Cambridge, 1997. 1 W.H. Press, S.A. Teukolsky, W.T. Vetterling, and B.P. Flannery, Numerical Recipes in Fortran 77 Cambridge University Press, Cambridge, 001. 13 S. Malinovskaya, P. Bucksbaum, and P. Berman unpublished. 013801-5