6
Ultrafast dynamics in atomic clusters: Analysis and control Vlasta Bonac ˇic ´-Koutecky ´* , Roland Mitric ´*, Ute Werner*, Ludger Wo ¨ ste , and R. Stephen Berry § *Institut fu ¨ r Chemie, Humboldt-Universita ¨ t zu Berlin, Brook-Taylor-Strasse 2, D-12489 Berlin, Germany; Institut fu ¨ r Experimentalphysik, Freie Universita ¨t Berlin, Arnimallee 14, D-14195 Berlin, Germany; and § Department of Chemistry, University of Chicago, Chicago, IL 60637 Edited by A. Welford Castleman, Jr., Pennsylvania State University, University Park, PA, and approved April 3, 2006 (received for review January 6, 2006) We present a study of dynamics and ultrafast observables in the frame of pump–probe negative-to-neutral-to-positive ion (NeNePo) spectroscopy illustrated by the examples of bimetallic trimers Ag2 Au Ag 2 AuAg 2 Au and silver oxides Ag 3 O 2 Ag 3 O 2 Ag3 O 2 in the context of cluster reactivity. First principle multistate adiabatic dynamics allows us to determine time scales of different ultrafast processes and conditions under which these processes can be experimentally observed. Furthermore, we present a strategy for optimal pump– dump control in complex systems based on the ab initio Wigner distribution approach and apply it to tailor laser fields for selective control of the isomerization process in Na3 F 2. The shapes of pulses can be assigned to underlying processes, and therefore control can be used as a tool for analysis. ultrafast ab initio dynamics NeNePo spectroscopy optimal control I nvestigation of femtosecond dynamical processes in elemental clusters and their control by tailored laser fields are of fundamental importance for learning how the interplay of size, structures, and laser fields can be used to manipulate optical properties and reactivity of these species (1). This research area involving combination of laser-selective femtochemistry (2–5) with the functionality of nanostructures opens new perspectives for basic research and numerous technical applications. In particular, exploration of clusters in the size regime in which each atom counts is attractive, because in this regime structures and the numbers of atoms directly determine size-selective properties (6 –12). Another important aspect is that the study of ultrafast dynamics in clusters with finite densities of states allows for separation of time scales of nuclear motion (1). Therefore, the identification of different ultrafast processes such as geo- metric relaxation, internal vibrational relaxation (IVR), differ- ent photoionization pathways, fragmentation, etc. becomes at- tainable (13–19). Moreover, optimization of the laser fields permits one to manipulate these processes by favoring or sup- pressing some of the chosen channels. In both contexts, the role of theory is essential from conceptual as well as from predictive point of view. Theory not only determines time scales of different processes and predicts ultrafast observables, but also finds conditions under which they can be experimentally realized (13). Moreover, the analysis of shaped laser pulses and the comparison with experimentally optimized laser fields allows us to identify the underlying processes and therefore to use optimal control (20– 23) as the tool for analysis (1, 23). In this contribution we address both aspects by showing what we can learn (i) from the pump–probe signals and (ii) from the optimized pulse shapes. We also wish to show that for accurate simulations of pump–probe signals and for developing new control strategies for complex systems, the semiclassical limit of the Liouville formulation of quantum mechanics offers a suit- able theoretical approach (1, 14, 15, 23). This is based on the Wigner–Moyal representation of the vibronic density matrix combined directly with adiabatic and nonadiabatic molecular dynamics (MD) ‘‘on the fly’’ (without precalculation of the energy surfaces) developed in one of our laboratories (14, 15, 23). This approach describes approximately quantum phenom- ena such as optical transitions by means of the averaged ensem- ble of classical trajectories. Moreover, the systematic introduc- tion of quantum effects is possible although computationally demanding and still in development (24). Our ab initio Wigner distribution approach is an appropriate choice to study ultrafast processes in elemental clusters with heavy atoms, for which in the first approximation the classical description of nuclear motion is acceptable and all degrees of freedom have to be considered because usually these clusters do not contain a ‘‘chromophore type’’ subunit and do not obey regular growth patterns. At the same time, we wish to show the scope of our pump– probe negative-to-neutral-to-positive ion (NeNePo) spectros- copy, introduced by some of the authors (25), which is capable of resolving structural properties, geometry relaxation, IVR, and isomerization processes (13). This will be illustrated on the example of bimetallic trimers and trimer-oxygen complexes. We wish also to present a strategy for control of ultrafast processes applicable to complex systems. Many control experi- ments are based on evolutionary algorithms in a feedback loop proposed by Judson and Rabitz (26). By using an iterative process, the method enables one to find an optimal pulse from which, under the given conditions, the target system can be reached. The aim of these experiments was to achieve maximal yields for a selected objective. However, the major goal of our theoretical approach is to attain information about the photo- induced process itself, which we wish to address here. In this context, a strategy for optimal control will be shown and applied to control the isomerization process in Na 3 F 2 cluster. Dynamics and Ultrafast Observables in the Framework of NeNePo Spectroscopy The real-time investigation of intra- and intercluster and mo- lecular electronic and nuclear dynamics by femtosecond spec- troscopy during the geometric transformation along the reaction coordinate is based on two steps: first on the preparation of the transition state of the chemical reaction by the optical excitation of a stable species in a nonequilibrium nuclear configuration in the pump step, and second on probing its time evolution by laser-induced techniques such as fluorescence, resonant mul- tiphoton ionization, or photoelectron spectroscopy (2, 4). A nonequilibrium or transition state can also be produced by vertical photodetachment of stable negative ions (27, 28). Ver- tical one-photon detachment techniques were advanced by in- troducing the NeNePo pump–probe experiments (25). They allowed for probing of structural relaxation and isomerization processes in neutral clusters as a function of the cluster size and Conflict of interest statement: No conflicts declared. This paper was submitted directly (Track II) to the PNAS office. Abbreviations: IVR, internal vibrational relaxation; MD, molecular dynamics; NeNePo, negative-to-neutral-to-positive ion; OCT, optimal control theory; ZEKE, zero kinetic energy electrons. To whom correspondence should be addressed. E-mail: [email protected]. © 2006 by The National Academy of Sciences of the USA 10594 –10599 PNAS July 11, 2006 vol. 103 no. 28 www.pnas.orgcgidoi10.1073pnas.0600151103 Downloaded by guest on February 1, 2021

Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

Ultrafast dynamics in atomic clusters:Analysis and controlVlasta Bonacic-Koutecky*†, Roland Mitric*, Ute Werner*, Ludger Woste‡, and R. Stephen Berry§

*Institut fur Chemie, Humboldt-Universitat zu Berlin, Brook-Taylor-Strasse 2, D-12489 Berlin, Germany; ‡Institut fur Experimentalphysik, Freie UniversitatBerlin, Arnimallee 14, D-14195 Berlin, Germany; and §Department of Chemistry, University of Chicago, Chicago, IL 60637

Edited by A. Welford Castleman, Jr., Pennsylvania State University, University Park, PA, and approved April 3, 2006 (received for review January 6, 2006)

We present a study of dynamics and ultrafast observables inthe frame of pump–probe negative-to-neutral-to-positive ion(NeNePo) spectroscopy illustrated by the examples of bimetallictrimers Ag2Au��Ag2Au�Ag2Au� and silver oxides Ag3O2

��Ag3O2�Ag3O2

� in the context of cluster reactivity. First principle multistateadiabatic dynamics allows us to determine time scales of differentultrafast processes and conditions under which these processes canbe experimentally observed. Furthermore, we present a strategyfor optimal pump–dump control in complex systems based on theab initio Wigner distribution approach and apply it to tailor laserfields for selective control of the isomerization process in Na3F2.

The shapes of pulses can be assigned to underlying processes, andtherefore control can be used as a tool for analysis.

ultrafast ab initio dynamics � NeNePo spectroscopy � optimal control

Investigation of femtosecond dynamical processes in elementalclusters and their control by tailored laser fields are of

fundamental importance for learning how the interplay of size,structures, and laser fields can be used to manipulate opticalproperties and reactivity of these species (1). This research areainvolving combination of laser-selective femtochemistry (2–5)with the functionality of nanostructures opens new perspectivesfor basic research and numerous technical applications. Inparticular, exploration of clusters in the size regime in whicheach atom counts is attractive, because in this regime structuresand the numbers of atoms directly determine size-selectiveproperties (6–12). Another important aspect is that the study ofultrafast dynamics in clusters with finite densities of states allowsfor separation of time scales of nuclear motion (1). Therefore,the identification of different ultrafast processes such as geo-metric relaxation, internal vibrational relaxation (IVR), differ-ent photoionization pathways, fragmentation, etc. becomes at-tainable (13–19). Moreover, optimization of the laser fieldspermits one to manipulate these processes by favoring or sup-pressing some of the chosen channels. In both contexts, the roleof theory is essential from conceptual as well as from predictivepoint of view.

Theory not only determines time scales of different processesand predicts ultrafast observables, but also finds conditionsunder which they can be experimentally realized (13). Moreover,the analysis of shaped laser pulses and the comparison withexperimentally optimized laser fields allows us to identify theunderlying processes and therefore to use optimal control (20–23) as the tool for analysis (1, 23).

In this contribution we address both aspects by showing whatwe can learn (i) from the pump–probe signals and (ii) from theoptimized pulse shapes. We also wish to show that for accuratesimulations of pump–probe signals and for developing newcontrol strategies for complex systems, the semiclassical limit ofthe Liouville formulation of quantum mechanics offers a suit-able theoretical approach (1, 14, 15, 23). This is based on theWigner–Moyal representation of the vibronic density matrixcombined directly with adiabatic and nonadiabatic moleculardynamics (MD) ‘‘on the fly’’ (without precalculation of theenergy surfaces) developed in one of our laboratories (14, 15,

23). This approach describes approximately quantum phenom-ena such as optical transitions by means of the averaged ensem-ble of classical trajectories. Moreover, the systematic introduc-tion of quantum effects is possible although computationallydemanding and still in development (24). Our ab initio Wignerdistribution approach is an appropriate choice to study ultrafastprocesses in elemental clusters with heavy atoms, for which in thefirst approximation the classical description of nuclear motion isacceptable and all degrees of freedom have to be consideredbecause usually these clusters do not contain a ‘‘chromophoretype’’ subunit and do not obey regular growth patterns.

At the same time, we wish to show the scope of our pump–probe negative-to-neutral-to-positive ion (NeNePo) spectros-copy, introduced by some of the authors (25), which is capableof resolving structural properties, geometry relaxation, IVR, andisomerization processes (13). This will be illustrated on theexample of bimetallic trimers and trimer-oxygen complexes.

We wish also to present a strategy for control of ultrafastprocesses applicable to complex systems. Many control experi-ments are based on evolutionary algorithms in a feedback loopproposed by Judson and Rabitz (26). By using an iterativeprocess, the method enables one to find an optimal pulse fromwhich, under the given conditions, the target system can bereached. The aim of these experiments was to achieve maximalyields for a selected objective. However, the major goal of ourtheoretical approach is to attain information about the photo-induced process itself, which we wish to address here. In thiscontext, a strategy for optimal control will be shown and appliedto control the isomerization process in Na3F2 cluster.

Dynamics and Ultrafast Observables in the Framework ofNeNePo SpectroscopyThe real-time investigation of intra- and intercluster and mo-lecular electronic and nuclear dynamics by femtosecond spec-troscopy during the geometric transformation along the reactioncoordinate is based on two steps: first on the preparation of thetransition state of the chemical reaction by the optical excitationof a stable species in a nonequilibrium nuclear configuration inthe pump step, and second on probing its time evolution bylaser-induced techniques such as fluorescence, resonant mul-tiphoton ionization, or photoelectron spectroscopy (2, 4). Anonequilibrium or transition state can also be produced byvertical photodetachment of stable negative ions (27, 28). Ver-tical one-photon detachment techniques were advanced by in-troducing the NeNePo pump–probe experiments (25). Theyallowed for probing of structural relaxation and isomerizationprocesses in neutral clusters as a function of the cluster size and

Conflict of interest statement: No conflicts declared.

This paper was submitted directly (Track II) to the PNAS office.

Abbreviations: IVR, internal vibrational relaxation; MD, molecular dynamics; NeNePo,negative-to-neutral-to-positive ion; OCT, optimal control theory; ZEKE, zero kinetic energyelectrons.

†To whom correspondence should be addressed. E-mail: [email protected].

© 2006 by The National Academy of Sciences of the USA

10594–10599 � PNAS � July 11, 2006 � vol. 103 � no. 28 www.pnas.org�cgi�doi�10.1073�pnas.0600151103

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1

Page 2: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

the atomic composition (13, 16). Moreover, the NeNePo spec-troscopy of clusters bridges ground-state dynamics of thesespecies with real-time investigation of chemical reactions, whichopens opportunities to study reactions of clusters toward inor-ganic and organic molecules and to control them by laser fields.

MethodsExperimental and Computational. The experimental setup for tem-perature-controlled ultrafast NeNePo spectroscopy is describedin ref. 18, and here only the main features will be outlined. Thecharged clusters formed in a sputtering process are steered intoa helium-filled quadrupole where clusters are thermalized. Aftermass selection, monodisperse clusters are then introduced intoa helium-filled (�1 Pa) octupole ion trap. The radio frequencyoctupole ion trap allows trapping of ions for seconds withoutsignificant loss. To achieve the temperature control, the trap isattached to a closed-cycle helium cryostat, which allows variationof temperature in the range between 20 and 350 K.

The femtosecond laser pulses were generated by a home-builtTi:sapphire oscillator of the Kapteyn–Murnane type that wascontinuously pumped by a 5-W Spectra-Physics MilleniaNd:YVO4 laser. Pulses were amplified with a Nd:YLF-laser-pumped Quantronix Odin multipass amplifier to yield 1-mJ,40-fs, 812-nm pulses at a 1-kHz repetition rate. The typical timeresolution obtained in the NeNePo measurements is 80 fs. Afterthe electron is photodetached the dynamics is probed by atime-delayed pulse that induces a two-photon ionization process.Cations formed then are extracted by electrostatic field andmass-analyzed with the final quadrupole mass filter. The cationicion current is recorded as a function of the pump–probe delaygiving rise to a transient NeNePo signal.

The simulations of the time-resolved NeNePo spectra werecarried out in the framework of our Wigner distribution ap-proach based on the propagation of the ensemble of classicaltrajectories carried out ‘‘on the fly.’’ Because here the dynamicsin the ground electronic states of neutral species is required, weemploy the gradient corrected density functional theory with theatomic orbitals (Gaussian basis sets) as the method of choice.For details on electronic structure calculations, cf. refs. 18and 29.

The simulations of the NeNePo pump–probe signals involvethree steps. (i) First an ensemble of initial phase space points isgenerated by sampling the Wigner distribution function corre-sponding to the initial state. This is done for the canonicalensemble invoking the harmonic approximation. The Wignerdistribution for each normal mode is given by

P�q, p� ��

� –hexp��

2�

–h��p2 � �2q2�� , [1]

where � � tanh(–h��2kBT), and � is the normal mode frequency,corresponding to the full quantum mechanical density distribu-tion. (ii) In the second step the initial phase space ensemble ispropagated on the neutral electronic state by solving the classicalequations of motion using forces calculated ‘‘on the fly.’’ Parallelto the propagation of the ensemble, the time-dependent energygaps between the neutral and cationic states are calculated foreach trajectory. (iii) The pump–probe signal is determined foreach trajectory according to

S�td� � limt3�

P22�2�� t�

� � dq0dp0 �0

d�1 exp����1 � td�

2

�pu2 �pr

2 �

exp���pr

2

–h2�–h�pr � V IP�q1��1; q0, p0���

2� exp��

�pu2

–h2�–h�pu

� VVDE�q1��1; q0, p0���2� P00�q0, p0� . [2]

In this expression �pu (�pr) and Epu � –h�pu (Epr � –h�pr) are thepulse durations and excitation energies for the pump and probestep with a time delay td. VIP(q1(�1; q0, p0)) labels the time-dependent energy gaps between the neutral and cationic groundstates calculated at coordinates q1(�1) on the neutral groundstate with initial coordinates and momenta q0 and p0 given by theanionic thermal Wigner distribution P00(q0, p0) and VVDE(q0) arethe vertical detachment energies of the initial ensemble. Thesignal is calculated by averaging over the whole ensemble. Thetime resolution of the signal is determined by the pump–probecorrelation function with the probe window located around thetime delay td, which is represented by the first exponential inEq. 2.

Results: NeNePo on Ag2Au��Ag2Au�Ag2Au� and Ag3O2��Ag3O2�

Ag3O2�. The ultrafast dynamics in the ground electronic state of

neutral Ag2Au induced by the photodetachment of the anionicspecies offers an opportunity to study the geometric relaxationfollowed by the ultrafast vibrational energy redistribution(IVR). The energy intervals necessary for detection of theseprocesses can be inferred from the scheme of the multistatefemtosecond dynamics presented in Fig. 1. Because the sim-ulations have been carried out for the initial ensemble at 20 K,it is justified to assume that only the linear isomer in theanionic state contributes because the second isomer is 0.36 eVhigher in energy (cf. Fig. 1). Moreover, due to the lowtemperature the harmonic approximation is valid, and theinitial conditions for the ensemble propagation were gener-ated by sampling the canonical Wigner distribution function asdefined in Eq. 1. For the detection, two scenarios can bedistinguished. The first involves the NeNePo-ZEKE (zerokinetic energy electrons) process in which no kinetic energy istransferred to the photodetached electron. The second sce-nario corresponds to the classical NeNePo experiment inwhich the excess energy is accounted for by integrating overthe populations of anionic and cationic states over the entireexcess of energy range. Usually, integrated signals do not allowdistinguishing different processes (cf. refs. 13 and 18).

The simulated and experimental signals are presented on theright side of Fig. 1. Because the probe pulse energy Epr � 7.7 eVexceeds the highest ionization potential of the neutral Ag2Auspecies (7. 49 eV; cf. Fig. 1), our NeNePo simulations predict thatthe signal corresponding to this energy will rise and becomeconstant at later times due to the contribution of the continuum.This allows probing the system while it leaves the transition stateregion reached by the photodetachment. Comparison betweenthe simulated and experimental signal presented in Fig. 1c showthe time scale for leaving the transition state region to be �1 ps.Reducing the energy of the probe pulse to 7.1 eV reveals anintermediate structure on the way between the linear andtriangular structure. The simulated NeNePo-ZEKE and exper-imental signal are in perfect agreement and allow detection ofthe geometrical relaxation with the onset at �500 fs. The signalreaches the maximum at �1 ps and has an offset at �2 ps. If theenergy of the probe pulse is further lowered to 6.1 eV, theappearance of the triangular structure and subsequent IVRprocess can be probed. This signal has its onset at �2 ps, reaches

Bonacic-Koutecky et al. PNAS � July 11, 2006 � vol. 103 � no. 28 � 10595

CHEM

ISTR

YSP

ECIA

LFE

ATU

RE

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1

Page 3: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

the maximum at �2.4 ps, and then decays slightly due to theincreasing delocalization of the phase space ensemble due to theIVR process. Good agreement between the simulated NeNePoand experimental signals (cf. Fig. 1a) indicates that in this casethe IVR process induced by the internal collision between theterminal Ag and Au atoms has been observed (18). It should bepointed out that, for the probe pulse energy of 6.1 eV, thecontribution of the continuum is negligible and therefore Ne-NePo-ZEKE and NeNePo simulated signals are very close toeach other (cf. two full lines in Fig. 1a). In summary, our jointtheoretical and experimental effort has shown that differentdynamical processes can be identified in the NeNePo spectra andthat their time scales can be determined, provided that theexperiment is carried out close to the ZEKE conditions shownin Fig. 1a.

Next, we show that NeNePo spectroscopy provides valuableinformation about dynamical processes in neutral species in thecontext of the cluster reactivity. We investigated previously theelectronic and structural aspects of adsorption of molecularoxygen on anionic silver clusters (9). Here the dynamics of thecomplex Ag3

�–O2 will be presented for the first time. The schemeof the multistate dynamics is presented in Fig. 2a. Anionicisomers (isomer AI with a linear Ag3 subunit, and isomer AIIwith a triangular Ag3 subunit) are very close in energy, and theexact energy sequence is difficult to predict theoretically, evenwith highly accurate methods. However, due to the pronounceddifferences in vertical detachment energies (VDEs) of bothisomers, they can be selected by appropriate choice of thepump-pulse energy. We assume that only isomer AI is populatedin the anionic ground state and use a canonical Wigner distri-bution function given in Eq. 1 at T � 20 K to generate the initialconditions. After the photodetachment with the pump-pulseenergy of 2.75 eV, a transition state in neutral Ag3O2 is reached.Three isomers are present for Ag3O2 as shown in Fig. 2. In thestructure of the global minimum (isomer I), the molecularoxygen bridges two silver atoms (cf. Fig. 2a). In the two

higher-lying isomers, molecular oxygen is bound to one silveratom, forming one and two bonds, respectively. Therefore, afterthe photodetachment a dynamical process dominated by the Ag3

relaxation from the linear to the triangular structure is induced.At later times the energy gained by the cluster isomerization ispartially transferred to molecular oxygen leading mainly to themixture of the two higher lying isomers (II and III). After 5 psthe global minimum (isomer I) is populated only by �10%. Thesnapshots of the dynamics in the neutral state are presented inthe Fig. 2c, showing the onset of the geometric relaxation of theAg3 subunit at 500 fs, the arrival to the triangular Ag3 subunit at1 ps, and the final ensemble of isomers after 5 ps. The simulatedNeNePo-ZEKE pump–probe signals for Ag3O2 are presented inFig. 2b. The signal for the probe pulse energy of 7.6 eV exhibitsa fast decay after �1 ps due to the escape from the initialFranck–Condon region reached upon the photodetachment.However, this signal starts to rise again after 2 ps because theglobal minimum with the ionization potential of 7.62 eV is beingpopulated at later times. This signal offers an opportunity toidentify the isomerization process leading to the global minimumin the ground state and also to determine its time scale. If theenergy of the probe pulse is lowered to 7.3 eV the onset of theisomerization of the Ag3 subunit starting at �500 fs can beselectively probed. Further lowering of energy to 6.8 eV allowsus to probe the IVR process induced after the intraclustercollision within the Ag3 subunit. It should be pointed out that incontrast to the Ag2Au, here the IVR process is extended also tothe O2 subunit and is responsible for the final populations of theisomers, because they mainly differ by the bonding of O2 subunit.

In summary, we expect that the future studies of the dynamicsof reactive complexes between metal clusters and small mole-cules may shed new light on the influence of the dynamics andparticularly IVR on their reactive (catalytic) properties. More-over, the population of the minimum with corresponding reac-tive centers can be controlled by tailored laser fields.

Fig. 1. Ag2Au��Ag2Au�Ag2Au NeNePo. (Left) Scheme of the multistate femtosecond dynamics for NeNePo pump–probe spectroscopy of Ag2Au��Ag2Au�Ag2Au with structures and energy intervals for pump and probe steps. (Right) Comparison between simulated (full line) and experimental (circles) pump–probesignals at different probe pulse energies.

10596 � www.pnas.org�cgi�doi�10.1073�pnas.0600151103 Bonacic-Koutecky et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1

Page 4: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

Concepts for Control of Ultrafast Processes inComplex SystemsExploration of coherence properties of laser radiation due toquantum mechanical interference effects led to the idea ofcontrolling the selectivity of product formation in a chemicalreaction by using ultrashort pulses with the proper choice of theirphase or duration and the delay between the pump and probe(20, 21). This initiated the shaping of pulses in the framework ofthe many-parameter optimal control theory (OCT) (22, 30, 31).Technological progress due to femtosecond-pulse shapers al-lowed introducing closed loop learning by Judson and Rabitz(26), which opened a possibility to control more complex sys-tems. The idea was to combine a femtosecond-laser system witha computer-controlled pulse shaper to produce tailored fieldsthat initiate the photochemical process and to use the learningalgorithm to optimize iteratively the field based on yield ob-tained from the experimental target. Such a black box is veryuseful but does not provide information about the nature of theunderlying processes that are responsible for the outcome. Thisis true in particular for complex systems, due to the manifold oflocal solutions within the multiparameter optimization schemethat are reachable depending on the initial conditions.

Therefore, for complex systems such as clusters, biomolecules,and their complexes, the important task is first to developtheoretical methods that allow one to design optimal laser pulsesand second to establish a connection between the underlyingprocesses. For such systems, precalculation of energy surfaces isimpractical, and therefore ab initio adiabatic and nonadiabatic

MD ‘‘on the fly’’ methods are particularly suitable. The Liouvillespace formulation of the OCT developed by Yan, Wilson,Mukamel, and their colleagues (32, 33), in particular its semi-classical limit in the Wigner representation, is suitable fordeveloping optimal control strategies for complex systems de-spite its limitations. Quantum effects, such as interferencephenomena, tunneling, and zero-point vibrational energy, arenot accounted for, and these limitations had to be respected bychoosing appropriate systems. Such a strategy for optimal con-trol (23) will be presented and applied for optimization of thelaser field that drives the photoisomerization process in theNa3F2 cluster into the second isomer avoiding the conicalintersection between the excited and electronic ground state.

Methods and ResultsNew Strategy for Optimal Control in Complex Systems. The appli-cation of OCT to optical control of the complex molecular andcluster species has been until now mainly prohibited by thedifficulties in the precalculation of multidimensional potentialsurfaces and the inability to perform the exact quantum dynam-ics simulations in these systems. Ab initio adiabatic and nonadia-batic MD ‘‘on the fly’’ represents a viable approach that canbypass these obstacles, identify properties necessary for assuringthe controllability of complex systems, and detect the mecha-nisms responsible for the obtained pulse shapes. For complexsystems it is still an open, central issue to determine theconditions under which optimal control involving more than oneelectronic state can be achieved (e.g., pump–dump control of

Fig. 2. Ag3O2��Ag2O2�Ag3O2

NeNePO. (a) Scheme of the multistate femtosecond dynamics for NeNePo pump–probe spectroscopy of Ag3O2��Ag3O2�Ag3O2

with structures and energy intervals for pump and probe steps. (b) Simulated NeNePo-ZEKE pump–probe spectra. (c) Snapshots of the dynamics in the neutralAg3O2 after the photodetachment.

Bonacic-Koutecky et al. PNAS � July 11, 2006 � vol. 103 � no. 28 � 10597

CHEM

ISTR

YSP

ECIA

LFE

ATU

RE

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1

Page 5: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

isomerization). The complicated potential-energy landscapes ofthe ground and excited electronic states can substantially differfrom each other, and therefore the existence of a connectingpathway between the initial state over excited state and thedesired objective in the ground state has to be insured. More-over, the optimal path must be found, and the method for nucleardynamics and pulse optimization should be computationallyrealistic. This calls for the development of new strategies for theoptimal control. An attractive possibility offers the concept ofthe intermediate target (cf. ref. 23) in the excited state. It isdefined as a localized ensemble (wavepacket) corresponding tothe maximum overlap between the forward propagated ensem-ble on the electronic excited state (starting from the initial state)and the backwards propagated ensemble from the objective inthe ground state at the optimal time delay between both pulses.In the case of the pump–dump control involving two phase-unlocked ultrafast fields in the weak response regime, we haveshown that the concept of the intermediate target leads to theseparate (uncoupled) optimization of the pump and dump fields,which is very advantageous from the computational point ofview. An appropriate formalism to implement this strategy isbased on the density matrix formulation of the OCT. It combinesthe Wigner–Moyal representation of the vibronic density matrixwith ab initio MD ‘‘on the fly’’ in the electronic excited and

ground states (cf. refs. 14 and 15). (For details, cf. refs. 1 and 23.)The envelopes of the optimal pump and dump pulses areobtained by solution of two coupled Fredholm-type integralequations, which require an iterative solution because the inte-gral kernels are coupled by the dump and pump field, respec-tively. The decoupling of these equations is possible in the shortpulse regime, as outlined in refs. 1 and 23. The pump-pulseoptimization involves the propagation on the excited state from� � 0 until � � td, forming the intermediate target. For thedump-pulse optimization, the dynamics in the ground state hasto be carried out by taking the intermediate target as an initialensemble and propagating in the ground state. Subsequently,both pump and dump pulses are determined by solving theFredholm-type integral equations.

We illustrate this strategy to optimize pump and dump pulsesfor driving the isomerization process in the Na3F2 cluster,avoiding the conical intersection between the ground and thefirst excited state and maximizing the population of the secondground-state isomer (cf. Fig. 3 Lower Left). For this purpose, theinitial ensemble corresponding to the 50-K canonical ensemblehas been generated by sampling the canonical Wigner distribu-tion function in the harmonic approximation (cf. Eq. 1). Theensemble has been propagated on the excited state for 300 fs.(For details of electronic structure and MD, cf. refs. 1 and 23.)

Fig. 3. Optimal pump–dump control. (Upper Left) The optimal pump and dump pulses. (Upper Right) Fourier transforms of the optimal pump and dump pulsesand the Franck–Condon profile for the first excited state corresponding to an excitation energy Te � 1.33 eV. Wigner transform of the optimal pump pulse isshown in Inset. (Lower Left) Scheme for pump–dump optimal control in the Na3F2 cluster with geometries of the two ground-state isomers and the transitionstate separating them, the conical intersection, and the intermediate target. (Lower Right) Snapshots of the dynamics obtained by propagating the ensemblecorresponding to the intermediate target after the optimized pump–dump at 250 fs.

10598 � www.pnas.org�cgi�doi�10.1073�pnas.0600151103 Bonacic-Koutecky et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1

Page 6: Ultrafast dynamics in atomic clusters: Analysis and control · 2006. 6. 30. · 10594–10599 PNAS July 11, 2006 vol. 103 no. 28 cgi doi 10.1073 pnas.0600151103. the atomic composition

To determine the intermediate target and the optimal time delayat which the dump should occur, the system has been dumped tothe ground state (in regular time steps of 25 fs) and propagatedsubsequently. The isomer II is optimally reached by the ensembledumped at td � 250 fs, and therefore this ensemble (cf. schemein Fig. 3 Lower Left) is chosen to construct the intermediatetarget. The optimal pump and dump pulses are presented in Fig.3 Upper Left. The optimal pump pulse exhibits two portions withdurations of 70 and 10 fs, respectively. The Fourier analysis of theoptimal pump pulse and the comparison with the Franck–Condon profile reveals that the excitation of low-lying vibra-tional modes at �1.2 eV of the isomer I is responsible for drivingthe dynamics to the intermediate target. The Wigner–Villetransform of the optimal pump pulse shows that this energeti-cally sharp transition corresponds to the first temporal portionof �70 fs of the pump pulse. In contrast, a very short secondportion after 80–90 fs is energetically much wider and corre-sponds to the tails of the Fourier transform that are symmetricwith respect to the 1.2 eV transition. The optimal dump pulse isvery short (�20 fs) leading to a short time window for thedepopulation of the excited state. The dynamics after the dumppulse illustrating the efficiency of the optimized pulses is pre-sented in Fig. 3 Lower Left. It can be clearly seen that the phasespace density is localized in isomer II after �450 fs. The achievedpopulation is �78%.

In summary, our strategy for optimal control based on theconcept of the intermediate target allows us to design pump anddump pulses that are able to drive the isomerization process andto populate the desired ground-state isomer. This is particularly

important in the context of the control of photochemistry and fordesign of molecular switches that can be driven by laser fields.Furthermore, the combination of OCT with the ab initio MD ‘‘onthe fly’’ allows one to gain insight into the mechanism governingthe control of the dynamics in complex systems and to establisha connection between the spectral features of the optimal pulsesand the underlying dynamical processes, which means the use oftailored laser fields as a tool for analysis. In the future, one otherpossible approach may be through kinetics, based on masterequations (34).

Summary and Outlook. We have shown that by joint theoreticaland experimental effort to analyze the ultrafast processes inatomic clusters in size regime in which ‘‘each atom counts,’’different ultrafast processes can be monitored and their timescales can be determined. They include bond-breaking, geomet-ric relaxation of different nature, IVR, isomerization, andreaction channels. Moreover, in the framework of our strategyfor tailoring laser fields, the selected processes can be drivensuch as isomerization toward one of the isomers, or the chosenreaction channel for which particular bond breaking or new bondrearrangements promote the emanation of reaction products.We have also shown that the optimal control schemes can beused as analysis, which is an important conceptual issue with apromising perspective for applications in complex systems suchas biomolecules and clusters, or even their complexes.

This work was supported by Deutsche Forschungsgemeinschaft SFB 450,‘‘Analysis and Control of Ultrafast Photoinduced Reactions.’’

1. Bonacic-Koutecky, V. & Mitric, R. (2005) Chem. Rev. 105, 11–66.2. Zewail, A. H. (1994) Femtochemistry (World Scientific, Singapore).3. Manz, J. & Woste, L., eds. (1995) Femtosecond Chemistry (VCH, Weinheim,

Germany), Vols. 1 and 2.4. Zewail, A. H. (2000) J. Phys. Chem. A 104, 5660–5694.5. Brixner, T. & Gerber, G. (2003) ChemPhysChem 4, 418–438.6. Bonacic-Koutecky, V., Fantucci, P. & Koutecky, J. (1991) Chem. Rev. 91,

1035–1108.7. Castleman, A. W., Jr., & Bowen, K. H. (1996) J. Phys. Chem. 100, 12911–12944.8. Jortner, J. (1997) Faraday Discuss. 108, 1–22.9. Landman, U., Barnett, R. N., Cleveland, C. L. & Chen, H.-P. (1992) Int. J. Mod.

Phys. B 6, 3623–3642.10. Haberland, H., ed. (1994) Clusters of Atoms and Molecules: Theory, Experiment,

and Clusters of Atoms, Springer Series in Chemical Physics (Springer, Berlin),Vol. 52.

11. Ekardt, W., ed. (1999) Metal Clusters (Wiley, Chichester, U.K.).12. Jena, P., Khanna, S. N. & Rao, B. K., eds. (1999) Cluster and Nanostructure

Interfaces (World Scientific, Singapore).13. Hartmann, M., Pittner, J., Bonacic-Koutecky, V., Heidenreich, A. & Jortner,

J. (1998) J. Chem. Phys. 108, 3096–3113.14. Hartmann, M., Pittner, J. & Bonacic-Koutecky, V. (2001) J. Chem. Phys. 114,

2106–2122.15. Hartmann, M., Pittner, J. & Bonacic-Koutecky, V. (2001) J. Chem. Phys. 114,

2123–2136.16. Mitric, R., Hartmann, M., Stanca, B., Bonacic-Koutecky, V. & Fantucci, P.

(2001) J. Phys. Chem. A 105, 8892–8905.17. Vajda, S., Lupulescu, C., Merli, A., Budzyn, F., Woste, L., Hartmann, M.,

Pittner, J. & Bonacic-Koutecky, V. (2002) Phys. Rev. Lett. 89, 213404.

18. Bernhardt, T. M., Hagen, J., Socaciu, L. D., Mitric, R., Heidenreich, A., LeRoux, J., Popolan, D., Vaida, M., Woste, L., Bonacic-Koutecky, V. & Jortner,J. (2005) ChemPhysChem 6, 243–253.

19. Henriksen, N. E. & Engel, V. (2001) Int. Rev. Phys. Chem. 20, 93–126.20. Tannor, D. J. & Rice, S. A. (1985) J. Chem. Phys. 83, 5013–5018.21. Shapiro, M. & Brumer, P. (1986) J. Chem. Phys. 84, 4103–4104.22. Peirce, A. P., Dahleh, M. A. & Rabitz, H. (1988) Phys. Rev. A 37, 4950–4964.23. Mitric, R., Hartmann, M., Pittner, J. & Bonacic-Koutecky, V. (2002) J. Phys.

Chem. A 106, 10477–10481.24. Donoso, A. & Martens, C. C. (2001) Phys. Rev. Lett. 87, 223202.25. Wolf, S., Sommerer, G., Rutz, S., Schreiber, E., Leisner, T., Woste, L. & Berry,

R. S. (1995) Phys. Rev. Lett. 74, 4177–4180.26. Judson, R. S. & Rabitz, H. (1992) Phys. Rev. Lett. 68, 1500–1503.27. Neumark, D. M. (2001) Annu. Rev. Phys. Chem. 52, 255–277.28. Wenthold, P. G., Hrovat, D., Borden, W. T. & Lineberger, W. C. (1996) Science

272, 1456–1459.29. Hagen, J., Socaciu, L. D., Le Roux, J., Popolan, D., Bernhardt, T. M., Woste,

L., Mitric, R., Noack, H. & Bonacic-Koutecky, V. (2004) J. Am. Chem. Soc. 126,3442–3443.

30. Kosloff, R., Rice, S. A., Gaspard, P., Tersigni, S. & Tannor, D. J. (1989) Chem.Phys. 139, 201–220.

31. Rice, S. A. (1992) Science 258, 412–413.32. Yan, Y. J., Gillian, R. E., Whitnell, R. M., Wilson, K. R. & Mukamel, S. (1993)

J. Phys. Chem. 97, 2320–2333.33. Krause, J. L., Whitnell, R. M., Wilson, K. R. & Yan, Y. J. (1995) in Femtosecond

Chemistry, eds. Manz, J. & Woste, L. (VCH, Weinheim, Germany).34. Kunz, R. E., Blaudeck, P., Hoffmann, K. H. & Berry, R. S. (1998) J. Chem. Phys.

108, 2576–2582.

Bonacic-Koutecky et al. PNAS � July 11, 2006 � vol. 103 � no. 28 � 10599

CHEM

ISTR

YSP

ECIA

LFE

ATU

RE

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 1,

202

1