Initial state laser control of curvecrossing reactions using the Rayleigh–Ritz variational procedure Peter Gross, Ashish K. Gupta, Deepa B. Bairagi, and Manoj K. Mishra Citation: J. Chem. Phys. 104, 7045 (1996); doi: 10.1063/1.471421 View online: http://dx.doi.org/10.1063/1.471421 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v104/i18 Published by the American Institute of Physics. Additional information on J. Chem. Phys. Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions Initial state laser control of curve-crossing reactions using the Rayleigh–Ritz variational procedure Peter Gross Tata Institute of Fundamental Research, Homi Bhabha Road, Bombay 400 005, India Ashish K. Gupta, Deepa B. Bairagi, and Manoj K. Mishra Department of Chemistry, Indian Institute of Technology, Powai, Bombay 400 076, India ~Received 21 June 1995; accepted 18 January 1996! A new two-step procedure for laser control of photodissociation is presented. In the first step of the procedure, we show that control of photodissociation product yields can be exerted through preparation of the initial wave function prior to application of the photodissociation field in contrast to previous laser control studies where attention has focused on the design of the field which induces dissociation. Specifically, for a chosen channel from which maximum product yield is desired and a given photodissociation field, the optimal linear combination of vibrational eigenstates which comprise the initial wave function is found using a straightforward variational calculation. Any photodissociation pulse shape and amplitude can be assumed since the Schrödinger equation is solved directly. Application of this method to control of product yields in the photodissociation of hydrogen iodide is demonstrated. The second step of the control procedure involves the preparation of the coherent superposition of discrete levels obtained from the previous step; design of the preparatory field can be done analytically for two or three level systems as demonstrated here or with other well-studied iterative field design methods. © 1996 American Institute of Physics. @S0021-9606~96!01516-1# I. INTRODUCTION In recent years there has been a renewed interest in laser control of chemical reactions. From the theoretical side, a variety of approaches or schemes for laser field design have emerged including the optimal control method introduced by Rabitz and co-workers,1,2 the pump-and-dump scheme of Rice and Tannor,3,4 and the coherent control method advocated by Brumer and Shapiro.5 The latter method has in particular attracted attention because, unlike the optimal control method for example, the fields employed are simple multicolor continuous wave ~cw! fields, and it has also been attempted in the laboratory with some success.6 Although the coherent control method and its variations have relied on perturbation theory ~therefore resulting in presumably small total yields although control of relative product yields, or the branching ratio, may be very good!, recently these authors have published a study of two-color control of diatomic dissociation in intense fields, i.e., the molecule–field interaction is treated nonperturbatively.7 In this work we develop a very simple extension of the perturbative coherent control method whereby an optimal ~small! superposition of vibrational eigenstates is prepared in the ground electronic potential. This superposition is defined as the initial wave function which is then subjected to the photodissociation pulse which results in selective dissociation out of a desired channel. The ‘‘control part’’ of this scheme is entirely encompassed in the preparation of the initial wave function; the photodissociation pulse is assumed to be fixed. Finding the optimal superposition state for a desired product yield objective is done using the Rayleigh– Ritz variational procedure. Although our method is similar in spirit to one of the Brumer and Shapiro coherent control J. Chem. Phys. 104 (18), 8 May 1996 variants,8 our method is more general in that it can accommodate any photodissociation pulse shape or intensity. The underlying motivation for this study is the fact that photodissociation processes have been shown to be influenced ~sometimes dramatically! by the initial state prior to dissociation. Examples include photodissociation of HCl ~Refs. 9, 10! and HI,11 where the product branching ratios were found to be highly sensitive to the initial vibrational state of the ground electronic curve. This vibrationallymediated photodissociation technique has been exploited in bond-selective photodissociation of HOD where the relative yields of the resulting products ~H1OD or OH1D! can be controlled to a large degree through selective IR excitation of certain vibrational modes of the ground surface before application of the UV photodissociation pulse.12–15 Also, a theoretical control study has been presented for the HOD system where a non-stationary vibrational state is prepared with an intense IR pulse and then subsequently dissociated with a weak UV pulse.16 Like the studies mentioned above, our control method of photodissociation reactions is through manipulation of the vibrational states prior to application of the photodissociation pulse. In particular, we advocate preparation of a nonstationary vibrational state as in Ref. 16. However, our method is more general and systematic in that one determines the best nonstationary state for a given control objective. This nonstationary state is comprised of only those eigenstates which we choose to be a part of the superposition. Usually, only a few ~two or three! eigenstates will be included in the initial wave function prior to photodissociation; experimental preparation of this coherent two or three level superposition may be possible in the laboratory. 0021-9606/96/104(18)/7045/7/$10.00 © 1996 American Institute of Physics Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 7045 Gross et al.: Laser control of curve-crossing reactions 7046 We emphasize that the method presented here has redefined the control problem from designing the photodissociation pulse which results in the desired products to preparing a coherent superposition of a few discrete ~bound! levels. In many cases this might be preferable to direct design of the photodissociation pulse because the traditional control problem, i.e., the problem of finding the field which produces a particular objective, now focuses solely on the preparation of a well-defined coherent superposition of a few discrete levels. Iterative control methods for field design, such as optimal control theory,2 would obviously be far more computationally tractable for designing fields involving a few bound discrete level system rather than photodissociation systems involving continuum states and/or multidimensional potential energy surfaces. Furthermore, the dynamics of these systems under the influence of laser fields ~including analytical solutions for two and three level systems! and controllability of discrete level systems has been well-studied.17–21 The balance of the paper is as follows: In Sec. II, the nonperturbative initial state control method is described and its relation to the Rayleigh–Ritz variational principle emphasized. In Sec. III, the method is applied to photodissociation of hydrogen iodide which possesses two dissociation pathways leading to either ground state or spin–orbit excited iodine. Curve-crossing systems have been used in the past for testing control schemes because of their relative simplicity from both a theoretical,22,23 and experimental perspective.24 Also in this section are some calculations to get a ‘‘feel’’ for the sensitivity of product branching ratios and yields in HI photodissociation as a function of initial vibrational state. We then present results for multicolor photodissociation pulses and compare product yields using unoptimized and optimized initial wave functions. It is shown that, by using the variational procedure to obtain the optimal combination of vibrational states, significant enhancement of product yields from either the I~2P 3/2! or I*~2P 1/2! channels is produced as desired. In Sec. IV, we present an analytical noniterative method for determining the preparatory field which provides the necessary coherent superposition of ~here three! vibrational eigenstates. This method, which invokes the rotating wave approximation ~RWA! in order to solve the time-dependent Schrödinger equation ~TDSE!, determines the amplitudes and phases of two cw pulses which induce transitions between the relevant vibrational states. Finally, Sec. V provides a summary and future extensions. where Û(T,0) is the ~not necessarily unitary! propagator and c~0! is the initial wave function to which the photodissociation pulse is applied. The time-integrated flux, which is directly related to the product yield, is E T 0 dt ^ ĵ & t 5 E T 0 dt ^ c ~ t ! u ĵ u c ~ t ! & , where the flux operator is defined as ĵ5 1 @ p̂ d ~ r2r d ! 1 d ~ r2r d ! p̂ # . 2m E T 0 dt ^ ĵ & t 5 E T 0 dt ^ c ~ 0 ! u Û † ~ t,0! ĵÛ ~ t,0! u c ~ 0 ! & , ~5! or E T 0 dt ^ ĵ & t 5 ^ c ~ 0 ! u F̂ u c ~ 0 ! & , ~6! where the operator F̂ is defined as F̂5 E T 0 dtÛ † ~ t,0! ĵÛ ~ t,0! . ~7! As mentioned previously, control over photodissociation product yields is sought through preparation of the initial wave function c~0! as a coherent superposition of eigenstates of the ground electronic potential. ~Since rotational motion is ignored in the present work, eigenstates of the ground potential refer only to vibrational eigenfunctions.! Thus, the propagator Û(T,0) in the previous equations is predetermined for all time t50 to t5T; we can only manipulate c~0! in order to maximize *T0 dt ^ ĵ & t and therefore the product yield. To do this, the Rayleigh–Ritz variational method is employed to find the extreme eigenvalues of the operator F̂ defined in Eq. ~7!. Note that F̂ is bounded from below and from above; if the initial wave function is normalized such that ^c~0!uc~0!&51.0, then the lower and upper extremes are 0.0 and 1.0. We proceed by first expanding c~0! in a basis of ~M11! vibrational eigenfunctions, M c~ 0 !5 ( c mf m . ~8! We then substitute the above basis set expansion into Eq. ~6! and construct the matrix elements, II. FORMULATION The solution of the time-dependent Schrödinger equation, ~1! at some time T can be expressed as c ~ T ! 5Û ~ T,0! c ~ 0 ! , ~4! Here m is the reduced mass along the reaction coordinate, r d is the grid point where the flux is evaluated, and p̂ is the momentum operator. Using Eq. ~2!, the time-integrated flux in Eq. ~3! becomes m50 ] c 5Ĥ ~ t ! c , i\ ]t ~3! ~2! F kl 5 ^ f k u F̂ u f l & , ~9! of the F matrix. Diagonalization of F provides the desired maximum eigenvalue F max which is equal to the maximum product yield ~flux!. The corresponding eigenvector, cmax , is the set of coefficients defined in Eq. ~8! which define the initial wave function c~0! which leads to the product yield F max . Since for experimental and practical reasons one would restrict the basis set expansion in Eq. ~8! to only the J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions Gross et al.: Laser control of curve-crossing reactions 7047 lowest vibrational eigenstates, the matrix F will be quite small in most cases of interest and therefore computationally trivial to diagonalize. It is important to note that to obtain the matrix elements F kl it is not necessary to compute the entire propagator Û(T,0). To see this, we first replace the integral in Eq. ~7! with the following summation ~as we do in actual computations!: Nt F̂5Dt ( n50 ~10! Û † ~ nDt,0! ĵÛ ~ nDt,0! , where the total time of propagation is divided up into N t 11 equally spaced time points and N t •Dt5T. Using Eq. ~10!, the matrix elements become @cf. Eq. ~9!# FIG. 1. Diabatic potential energy curves of HI. Nt F kl 5Dt ( n50 ^ f k u Û † ~ nDt,0! ĵÛ ~ nDt,0! u f l & , ~11! or ^ck~nDt!u ĵucl~nDt!&5 Nt F kl 5Dt ( can be evaluated as n50 ^ c k ~ nDt ! u ĵ u c l ~ nDt ! & , ~12! where c k (nDt) and c l (nDt) are solutions of the TDSE with the chosen photodissociation pulse assuming the initial conditions c~0!5f k , f l , i.e., the initial wave function is a vibrational eigenstate. Thus, if the expansion in Eq. ~8! contains three eigenstates, then we need only propagate the TDSE from t50 to t5T three times rather than construct the entire propagator Û(T,0). The nonperturbative initial state control procedure can be outlined as follows: ~1! Choose a photodissociation pulse and time interval T for which the TDSE is to be solved. Note that the length of the pulse and the total propagation time T need not be the same; if desired, one may propagate the TDSE after the pulse has turned off in case there remains some molecules with sufficient energy which have not yet dissociated when the pulse is shut off. There are no restrictions on the field intensity or shape of the field since the TDSE is solved numerically. ~2! Choose the eigenstates of the ground potential surface which will be included in the variational calculation. As in other similar calculations, the larger the basis set expansion, the ‘‘better’’ the results. In the initial state control method, ‘‘better’’ results means better product yields. However, in practice only a few ~two or three! eigenstates can reasonably be included in the expansion due to experimental constraints of preparing a large coherent superposition of eigenstates prior to application of the photodissociation pulse. ~3! Propagate the TDSE M 11 times from t50 to t5T for each initial condition c~0!5fm , m50,1...,M in the basis set expansion @Eq. ~8!#. During propagation, the matrix elements F kl must be accumulated according to Eq. ~12!. The most efficient way to do this is to propagate all M 11 wave functions c~0!5f0 ,f1 ,...,fM simultaneously. At each time step nDt, the matrix elements F kl F ]c* i\ k ~nDt ! cl~nDt! 2m ]r 2c* k ~nDt ! G ]cl~nDt! , ]r r ~13! d where the r d subscript indicates that the wave functions and their derivatives are evaluated at the flux indicator point r d located in the asymptotic region of the potential. In practice, the derivatives were evaluated with a ninepoint differentiation formula. Since the F matrix is Hermitian, it is necessary to compute only the upper or lower triangular elements of F. ~4! Diagonalize F and pick the largest eigenvalue and corresponding eigenvector which represents the superposition of eigenstates that provides the highest desired product yield as discussed earlier. III. APPLICATION TO PHOTODISSOCIATION OF HI The hydrogen iodide system ~HI! used in this work is modeled as a rotationless oscillator. Field–molecule interaction is within the semiclassical dipole approximation and the laser field is assumed to be linearly polarized along the molecular axis. The five relevant potential energy curves are shown in Fig. 1. All potential energy parameters, diabatic couplings, and transition dipoles are taken from Ref. 25. ~Note that the diabatic coupling strength listed in Table I of Ref. 25 between the 1P1 and 3S1 states is an order of magnitude too large, although the correct value is shown in Fig. 2 of the same reference.! The TDSE is set up exactly as in Ref. 11 except here the ground electronic state potential is included in the electronic Hamiltonian matrix. The TDSE was solved numerically using the split-operator fast Fourier transform ~FFT! method.26 A total spatial grid length of 9.0 spanning from 1.0 to 10.0 ~atomic units used unless otherwise noted! was used and divided into 256 equally spaced grid points. ‘‘Ramp’’ ~linear! optical potentials were included at the ends of the grid ranging from r58.5 to r510.0 with a maximum height of 0.05 in order to avoid spurious wave packet reflection. The flux point r d on all curves was J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 7048 Gross et al.: Laser control of curve-crossing reactions FIG. 2. Photodissociation branching ratio @I*~2P 1/2!/I~2P 3/2!# vs. laser excitation energy ~v!. The field employed is A cos ( v t), where A50.01 atomic units and the pulse length is 100 optical cycles. set at 8.0. The vibrational eigenfunctions of the ground state potential were computed using the Fourier grid Hamiltonian ~FGH! method.27 In order to test our program and to demonstrate the sensitivity of the photodissociation branching ratio with respect to initial vibrational state, we repeated the calculations in Ref. 11 with the difference that we do not assume weak-field conditions whereby the initial wave function can simply be projected ‘‘upstairs’’ onto the upper surfaces and propagated without explicit inclusion of the time-dependent field in the Hamiltonian. Figure 2 presents the product branching ratio @I*~2P 1/2!/I~2P 3/2!# vs frequency ~excitation energy! for c~0!5f0 , f1 , and f2 initial conditions. The photodissociation pulse here is cw with amplitude 0.01 and a length of 100 optical cycles. The branching ratio is determined by computing the total time-integrated fluxes out of channels leading to I*~2P 1/2! ~3S1 and 3P0! and I~2P 3/2! ~1S0 , 3P1 , and 1P1!. @Note that in the multicurve problem here, flux must be evaluated for all five components of the wave function corresponding to the five potential curves. Thus, the flux operator ĵ defined in the previous section refers to flux out of the I~2P 3/2! channel ~J 1! or the I*~2P 1/2! channel ~J 2!.# Our results qualitatively agree with those in Ref. 11 although they do not match exactly because the field strength used here is outside the perturbation regime. Note also, as in Ref. 11, there is strong evidence of sensitivity to initial vibrational quantum number. FIG. 3. Product yield ~flux! from ~a! I~2P 3/2! and ~b! I*~2P 1/2! channels vs v ~same field used for Fig. 2 results employed!. Results are shown assuming the initial wave function c~0! is one of the three lowest vibrational eigenstates f0 , f1 , f2 . In our control scheme the control objective is not the branching ratio itself but rather the photodissociation yields out of each channel. Thus a more relevant frame of reference for our purposes is a plot of fluxes or yields from both channels as a function of laser frequency. Using the same pulse as in Fig. 2, we show in Fig. 3 product yields from both channels J 1 and J 2 . Again, as in Fig. 2, there is a significant sensitivity with respect to initial vibrational quantum number. As mentioned previously, in our scheme we seek to control the product yields by finding the best superposition of vibrational eigenstates for the initial wave function c~0!. In all calculations presented below we include only the lowest three vibrational eigenfunctions in our basis set expansion @M 52 in Eq. ~8!#. Instead of employing a single color field as in Figs. 2 and 3, we employ here multicolor fields of the form 2 E ~ t ! 5A ~ t ! ( p50 cos~ v 2 v p,0! t, ~14! where A(t) is the amplitude function, v is the primary photodissociation frequency, and v p,05(E p 2E 0 )/\ is the energy level difference between the pth vibrational energy and the ground vibrational energy. Although the theory presented in Sec. II is valid for any field, the above form was chosen J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions Gross et al.: Laser control of curve-crossing reactions FIG. 4. Product yield ~flux! from ~a! I~2P 3/2! and ~b! I*~2P 1/2! channels vs. v employing a three-color field E(t)5A(t)(2p50 cos@( v 2 v p,0)t# for both unoptimized ~open circles! and optimized ~filled circles! initial wave functions. Here, constant amplitude fields are used, i.e., A(t)5A. because all the cw components resonantly excite to the same energy in the upper electronic continuum states, i.e., v induces a resonant transition from v 50 to energy E in the continuum; v2v1,0 induces a transition from v 51 to E, etc. This then allows for control over degenerate states at energy E which asymptotically correspond to J 1 or J 2 products.8 Here, however, there are no restrictions on the pulse intensity or envelope, and thus more realistic pulse shapes such as Gaussians may be employed. Figure 4 shows results for both unoptimized ~open circles! and optimized ~filled circles! computed from the initial state control scheme. By unoptimized we mean that we have chosen the vibrational eigenfunction f0 , f1 , or f2 as the initial state c~0! which provides the maximum yield for the product sought. In other words, only one of the three lowest vibrational states ~the one which provides the best product yield out the desired channel! is initially populated in the unoptimized results. For the optimized results we allow for mixing of the lowest three vibrational eigenfunctions as described above. Figures 4~a! and 4~b! show results assuming a cw field where A(t)5A50.01 in Eq. ~14!. All results are for the same set of primary frequencies ~v! in the range 38 000–48 000 cm21. In Fig. 4~a!, J 1 is maximized, and in Fig. 4~b!, J 2 is maximized. Note that the optimized results are significantly better than the unoptimized ones, in some cases almost 50% better. 7049 FIG. 5. Same as Fig. 4 except using a three-color Gaussian field, i.e., A(t)5A exp@2a(t2t 0 ) 2#. In Figs. 5~a! and 5~b! we show similar results as before except here the amplitude function A(t) is a Gaussian; A(t)5A exp@2a(t2t 0 ) 2#. The maximum amplitude A is 0.01, the width is a50.000 002, and the mean time t 0 is ~50!~2p/v!. Note that in general the overall yields are lower in the case of Gaussian pulses ~compare the y-axes of Figs. 4 and 5!. This is largely due to the overall lower field fluence in the Gaussian case. However, as in the cw field case, improvement using the optimized initial state is still significant. IV. PREPARATORY FIELD Here we address the issue of preparing the initial coherent superposition of states resulting from the above proposed variational calculation. The initial wave function to which the photodissociation pulse is applied is, for the example given in the previous section where only the three lowest vibrational eigenstates were included in the variational calculation, 2 c ~ 0 ! 5 ( c opt m fm , m50 ~15! where c opt m are the ~complex! optimal coefficients of the vibrational eigenstates fm determined from the diagonalization of F as described in Sec. II. The problem is to find a preparatory field which gives this superposition. We take the preparatory field form to be J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions Gross et al.: Laser control of curve-crossing reactions 7050 E prep~ t ! 52A 1 cos~ v 10t1 d 1 ! 12A 2 cos~ v 21t1 d 2 ! , ~16! where the field frequencies v nm 5( e n 2 e m )/\ are chosen such that the first cw component is resonant with the v 50→ v 51 transition, and the second cw component is resonant with the v 51→ v 52 transition. The amplitudes are assumed to be sufficiently weak such that the conditions um10uA 1/\!v10 and um21uA 2/\!v21 are fulfilled ~here m nm 5 ^ f n u m̂ u f m & 5 m mn are the transition dipole matrix elements!. Furthermore, the cw components should not appreciably excite nonresonant transitions @v 5m→ v 5n:(m,n)Þ(0,1) or ~1,2!#, i.e., the condition ~ m nm ! 2 ~ A 2 ! 2 ~ m nm ! 2 ~ A 1 ! 2 !1 2 2 2, 2 \ ~ v nm 2 v 21! 1 ~ m nm ! ~ A 2 ! \ ~ v nm 2 v 10! 2 1 ~ m nm ! 2 ~ A 1 ! 2 ~17! 2 must be satisfied. Roughly, pulse lengths in the picosecond regime and cw amplitudes of '1024 a.u. will fulfill the above conditions for HI and other similar systems. With these assumptions, the RWA may be invoked in solving for the dynamics of the discrete level system whereby all nonresonant transitions are ignored. The goal is to determine the field parameters A 1 , A 2 , d1 , and d2 which provide the required initial state wave function c~0!. We define the optimal coefficients determined from the variational procedure c opt m [ c m ( t ), i.e., the coefficients of the lowest three vibrational eigenstates resulting from application of the preparatory laser pulse given by Eq. ~16! for a time duration t. With the aforementioned weak-field assumptions and invoking the RWA, the solution of the TDSE in the eigenstate basis for the three coefficients for a pulse of length t is, assuming that before application of the preparatory field E prep(t) all vibrational population is in the ground vibrational state @uc 0~0!u251.0#, c 0 ~ t ! 5exp~ 2i e 0 t /\ ! SD c 1 ~ t ! 5exp~ 2i e 1 t /\ ! S D c 2 ~ t ! 5exp~ 2i e 2 t /\ ! 1 $ a 1 u a u 2 @ cos~ Aat /\ ! 21 # % , a ~18! ia * Aa sin~ Aat /\ ! , S D a *b * $ cos~ Aat /\ ! 21 % , a ~19! ~20! b5A 2m21 exp~i d 2!, and where a5A 1m10 exp~i d 1!, a5u a u 2 1 u b u 2 . Using the fact that ^c~0!u F̂ u c~0!& in Eq. ~6! does not depend on the overall phase factor of the initial wave function @i.e., c~0!→e i b c~0! does not change the expectation value# and performing the necessary algebra, Eqs. ~18!–~20! can be solved for the required cw field amplitudes and phases, A 15 A 25 \ u m 10u t A11M 2 cos21 $ 12 @~ 11M 2 ! /M # u c 2 ~ t ! u % , u m 10u M A1 , u m 21u ~22! p e 1t 2 u 1 1q 1 , d 15 2 h 12 b 2 2 ~21! \ ~23! d 25 p 2 d 12 h 12 h 22 b 2 e 2t 2 u 2 1q 1 1q 2 . \ ~24! Here, M 5 u c 2 ( t ) u /[12 u c 0 ( t ) u ], b52@u01e0t/\#, and u0 , u1 , and u2 are the phases of the desired coefficients c opt 0 iu1 [ c 0 ( t ) 5 u c 0 ( t ) u e i u 0 , c opt , and c opt 1 [ c 1( t ) 5u c 1( t )ue 2 [ c 2 ( t ) 5 u c 2 ( t ) u e i u 2 , respectively. The remaining parameters are defined as follows: h150 for m10.0 or h15p for m10,0, h250 for m21.0 or h25p for m21,0, q 15p for Re$c 1~t!%,0 or q 150 for Re$c 1~t!%.0, and q 25p for Re$c 2~t!%,0 or q 250 for Re$c 2~t!%.0. Finally, if Re$c 0~t!%,0 then all coefficients should be multiplied by 21 prior to solving for the field parameters in Eqs. ~21!–~24!. The coefficients thus obtained with the pulse of the form in Eq. ~16! using the field parameters defined by Eqs. ~21!–~24! are the optimal coefficients in Eq. ~15! to within a common unimportant phase factor @e i b or e i( b 1 p ) #. The above direct analytical solution can be used only for preparation of a maximum of three coherent states. @For a desired coherent superposition of two states, f0 and f1 , just the first resonant cw component in Eq. ~16! is sufficient to prepare any superposition, and the solutions for c 0~t! and c 1~t! as well as the required field parameters, A 1 and d1 , may be easily derived as a special case of Eqs. ~21!–~24! and Eqs. ~18!–~20! above.# For more than three states, one must resort to numerical field design methods such as optimal control theory.2 However, other noniterative numerical methods have been recently developed which can ‘‘scan’’ through field parameter space ~cw component amplitudes, frequencies, and relative phases between field components! and determine the wave function quickly for a wide range of field parameters.28,29 Such methods could be employed for determining the preparatory field if c~0! is comprised of more than three eigenstates. We note, however, from a practical point of view, the preparation of more than three coherent states would entail more complex preparatory fields ~e.g., more than two phase-locked cw pulses!, and thus purely from an experimental perspective one may not want to prepare a superposition of more than three coherent states. Therefore, if coherent preparation of three states is the upper limit of experimental viability, then theory provides a complete noniterative solution to the field design problem as outlined in this paper. J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions Gross et al.: Laser control of curve-crossing reactions V. CONCLUSION We have presented a scheme for laser control of photodissociation processes which entails preparation of the initial wave function as a coherent superposition of a few bound vibrational eigenstates before application of a given photodissociation pulse. The optimal linear combination of eigenstates is found using the Rayleigh–Ritz variational procedure. This scheme is extremely simple to use in that it requires only a standard wave packet propagation routine and a subroutine to diagonalize small Hermitian matrices. Preparation of the coherent superposition of states, at least for two or three levels, can be done using a simple field of one or two cw pulses, respectively. Furthermore, assuming the pulses are sufficiently weak so that the RWA is valid, it has been demonstrated that the required amplitudes and phases of the pulses which result in the desired superposition can be found analytically. Thus, assuming a basis of no more than three eigenstates for which an optimal linear combination is determined by the proposed variational procedure, the field design prescription is noniterative and direct using the proposed control scheme. The use of larger basis sets in the variational calculation, i.e., allowing for a superposition of four or more vibrational eigenstates in the initial wave function prior to dissociation, will likely require iterative optimization methods for designing the preparatory field, e.g., optimal control. However, solving the TDSE for a few level discrete system, even many times, will certainly be less costly than designing the photodissociation pulse itself via iterative methods since this would entail repeated wavepacket propagations on the relevant dissociative electronic curves. Future work will include application of this control scheme to other curve-crossing systems30 such as IBr and HCl as well as to reactive systems such as HOD where initial state preparation prior to photodissociation has been theoretically and experimentally shown to be very influential in terms of product yields. Finally, in relation to the issue of experimental preparation of the superposition of vibrational eigenstates, it would be worthwhile to determine just how sensitive the results are to minor deviations in the optimal eigenstate amplitudes. Sensitivity analysis might prove very 7051 useful in determining how much allowance for error can be reasonably tolerated in an actual experimental preparation of the initial wave function. ACKNOWLEDGMENTS M.K.M. acknowledges support from the Board of Research in Nuclear Sciences ~BRNS Grant No. 37/9/94-R1DII/787! from the Department of Atomic Energy, India. D.B.B. acknowledges support from CSIR Junior Research Fellowship No. 9/87~143! 92/EMR-I. W. S. Warren, H. Rabitz, and M. Dahleh, Science 259, 1581 ~1993!. H. Rabitz and S. Shi, Adv. Mol. Vib. Coll. Dynamics 1A, 187 ~1991!. 3 S. A. Rice and D. J. Tannor, Adv. Chem. Phys. 70, 441 ~1988!. 4 J. Somloi and D. J. Tannor, J. Phys. Chem. 99, 2552 ~1995!. 5 P. Brumer and M. Shapiro, Annu. Rev. Phys. Chem. 43, 257 ~1992!. 6 S.-P. Liu, S. M. Park, Y. Xie, and R. J. Gordon, J. Chem. Phys. 96, 6614 ~1992!. 7 Z. Chen, P. Brumer, and M. Shapiro, Chem. Phys. Lett. 228, 289 ~1994!. 8 P. Brumer and M. Shapiro, Chem. Phys. Lett. 126, 541 ~1986!. 9 S. C. Givertz and G. G. Balint-Kurti, J. Chem. Soc. Faraday Trans. II 82, 1231 ~1986!. 10 I. H. Gersonde, S. Hennig, and H. Gabriel, J. Chem. Phys. 101, 9558 ~1994!. 11 C. Kalyanaraman and N. Sathyamurthy, Chem. Phys. Lett. 209, 52 ~1993!. 12 D. G. Imre and J. Zhang, Chem. Phys. 139, 89 ~1989!. 13 Y. Cohen, I. Bar, and S. Rosenwaks, J. Chem. Phys. 102, 3612 ~1995!. 14 R. L. Vander Wal, J. L. Scott, F. F. Crim, K. Weide, and R. Schinke, J. Chem. Phys. 94, 3548 ~1991!. 15 F. F. Crim, Annu. Rev. Phys. Chem. 44, 397 ~1993!. 16 B. Amstrup and N. E. Henriksen, J. Chem. Phys. 97, 8285 ~1992!. 17 T. C. Kavanaugh and R. J. Silbey, J. Chem. Phys. 98, 9444 ~1993!. 18 F. T. Hioe, in Lasers, Molecules, and Methods, edited by J. O. Hirschfelder, R. E. Wyatt, and R. D. Coalson ~Wiley, New York, 1989!. 19 S. H. Tersigni, P. Gaspard, and S. A. Rice, J. Chem. Phys. 93, 1670 ~1990!. 20 G. M. Huang, T. J. Tarn, and J. W. Clark, J. Math Phys. 24, 2608 ~1983!. 21 V. Ramakrishna, M. V. Salapaka, M. Dahleh, H. Rabitz, and A. Pierce, Phys. Rev. A 51, 960 ~1995!. 22 P. Gross, D. Neuhauser, and H. Rabitz, J. Chem. Phys. 96, 2834 ~1992!. 23 C. K. Chan, P. Brumer, and M. Shapiro, J. Chem. Phys. 94, 2688 ~1991!. 24 H. K. Haugen, E. Weitz, and S. R. Leone, J. Chem. Phys. 83, 3402 ~1985!. 25 I. Levy and M. Shapiro, J. Chem. Phys. 89, 2900 ~1988!. 26 R. Kosloff, J. Phys. Chem. 92, 2087 ~1988!. 27 C. C. Marston and G. G. Balint-Kurti, J. Chem. Phys. 91, 3571 ~1989!. 28 P. Gross, A. K. Gupta, D. B. Bairagi, and M. K. Mishra, Chem. Phys. Lett. 236, 8 ~1995!. 29 A. K. Gupta, P. Gross, D. B. Bairagi, and M. K. Mishra, Chem. Phys. Lett. ~submitted!. 30 D. B. Bairagi, A. K. Gupta, P. Gross, and M. K. Mishra ~in preparation!. 1 2 J. Chem. Phys., Vol. 104, No. 18, 8 May 1996 Downloaded 01 Mar 2012 to 14.139.97.79. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions
© Copyright 2025 Paperzz