reactive molecular collisionslight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive...

33
Ann.Re~.Phys. Chern, 1980. 31:401-33 REACTIVE MOLECULAR COLLISIONS 2712 Robert B. Walker Theoretical Division, University of California, Los Alamos Scientific Labora- tory, Los Alamos, New Mexico 87545 John C. Light Department of Chemistry and the James Franck Institute, University of Chicago, Chicago, Illinois 60637 INTRODUCTION Thereactive molecular collision is the fundamental microscopicevent. that underlies the chemical reaction. The ability to describe this process from a knowledge of forces operating at the molecular level has long been the goal of the theoretical dynamicist. The field of reactive molecular collisions, and fields related to it, have been reviewed frequently within the last five years. For the newcomer, two excellent textbooks on molecular collision dynamics provide a broad background of the subject: the introductory text, Molecular Reaction Dynamics, by Levine & Bernstein (1) and the more comprehen- sive text by Child (2), Molecular Collision Theory. In addition, there are several volumes whose chapters consist of excellent reviews of specific topics relevant to reactive collisions. The most recent of these, and the most comprehensive to date, is Bernstein’s handbook for the experi- mentalist, Atom-Molecule Collision Theory [Reference (3), abbreviated here to AMCT].Every topic we consider in this review is represented by one or more chapters in Bernstein’s book. In addition to Bernstein’s volume, equally excellent review material (although now somewhat older) is contained in two volumes edited by Miller, Dynamics of Molecular Collisions, Part A [Reference (4), abbreviated DMCA],and Part B [Reference (5), abbreviated DMCB]. 1The US Government has the fight to retain a nonexclusive,royalty-free license in and to any copyright covering this paper. 401 www.annualreviews.org/aronline Annual Reviews

Upload: others

Post on 24-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

Ann. Re~. Phys. Chern, 1980. 31:401-33

REACTIVE MOLECULARCOLLISIONS

2712

Robert B. WalkerTheoretical Division, University of California, Los Alamos Scientific Labora-tory, Los Alamos, New Mexico 87545

John C. LightDepartment of Chemistry and the James Franck Institute, University of Chicago,Chicago, Illinois 60637

INTRODUCTION

The reactive molecular collision is the fundamental microscopic event.that underlies the chemical reaction. The ability to describe this processfrom a knowledge of forces operating at the molecular level has longbeen the goal of the theoretical dynamicist.

The field of reactive molecular collisions, and fields related to it, havebeen reviewed frequently within the last five years. For the newcomer,two excellent textbooks on molecular collision dynamics provide abroad background of the subject: the introductory text, MolecularReaction Dynamics, by Levine & Bernstein (1) and the more comprehen-sive text by Child (2), Molecular Collision Theory. In addition, there areseveral volumes whose chapters consist of excellent reviews of specifictopics relevant to reactive collisions. The most recent of these, and themost comprehensive to date, is Bernstein’s handbook for the experi-mentalist, Atom-Molecule Collision Theory [Reference (3), abbreviatedhere to AMCT]. Every topic we consider in this review is represented byone or more chapters in Bernstein’s book. In addition to Bernstein’svolume, equally excellent review material (although now somewhatolder) is contained in two volumes edited by Miller, Dynamics of

Molecular Collisions, Part A [Reference (4), abbreviated DMCA], andPart B [Reference (5), abbreviated DMCB].

1The US Government has the fight to retain a nonexclusive, royalty-free license in andto any copyright covering this paper.

401

www.annualreviews.org/aronlineAnnual Reviews

Page 2: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

402 WALKER & LI(;HT

Our review necessarily does not cover several topics that are insepara-ble from reactive molecular collision theory. These include the vastbody of sophisticated and detailed experimental data, upon which thetheory must be built. Experiments are discussed in many reviews,including Bernstein’s AMCT chapter (6). The most recent experimentalreview is given by Levy (7), and includes a truly exhaustive compilationof data. for every atom-diatom system studied in recent times.

We also cannot do justice to describing the efforts of the quantumchemist to generate the potential energy surfaces upon which all dy-namical theories critically depend. In his AMCT review, Schaefer (8)describes considerable progress with ab initio methods. The dynamicist’sneed to evaluate points on potential surfaces often and cheaply promptsour interest in semiempirical methods, such as the method of diatomics-in-molecules (DIM). A detailed description of the DIM method presented by Kuntz (9) in an AMCT review, and the earlier DMCBreview by Kuntz (10) describes how reaction dynamics is affected specific properties of potential energy surfaces.

Information theoretic concepts now permeate almost every aspect ofmolecular collision dynamics, including reaction. This is an extremelyvigorous field, and reviews appear on a yearly basis. The most recent isagain an AMCT review, a chapter by Levine & Kinsey (11). Levine (12)has also given a review in the 1978 volume of the Annual Review ofPhysical Chemistry series. We consider other features of reaction dy-namics here.

After describing what we do not cover in this review, we now turn ourattention to the topics we do plan to discuss. There are three majorsections in this review: the first describes the classical dynamical ap-proach to reactions; the second concentrates on quantum dynamics; thethird section considers some special topics.

Most dynamical treatments of reactive collisions use classical mecha-nics-assembling the global dynamics by calculating ensembles oftrajectories evolving on the appropriate potential surface(s). We woulddo the calculations quantum mechanically, except that at present theonly practical quantum technology for reaction dynamics that retainsthe full mathematical dimensionality of the problem is applicable onlyto the simplest systems. By assuming that nuclear motion can be treatedclassically, we opt instead for a practical (but at least well-defined)dynamical approximation. The classical approach is documented inmany reviews: The most recent are the AMCT reviews by Pattengill(13) for inelastic rotational scattering and by Truhlar & Muckerman(14) for reactive scattering. The earlier DMCB review by Porter & Raft(15) presents an excellent discussion of classical methodology applied

www.annualreviews.org/aronlineAnnual Reviews

Page 3: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 403

atom-diatom systems. This review describes in detail how to selectinitial conditions for trajectories; the Truhlar & Muckerman reviewgives a more detailed discussion for analyzing final conditions.

Although classical mechanics is itself well defined, its applications tomolecular reactions and the interpretation of its results evolve withexperience. Here we briefly consider the theory as it is most oftenapplied (to atom-diatom systems), and we then discuss the problems expect to encounter as these methods are extended to more complicatedsystems.

Exact quantum scattering calculations are limited by basis-set consid-erations to calculations in one physical dimension. There is only ahandful of converged three-dimensional calculations. Several reviews ofquantum collision methods are contained in Bernstein’s AMCThandbook. The review by Secrest (16) compares several numericalmethods cormnonly used to solve the coupled equations characteristic ofquantum collision theories. Light (17) presents a formal introduction the reactive problem, and Wyatt (18) presents a detailed comparison methods that have been applied in actual calculations. Wyatt (19) alsopresents a separate discussion of various approximate quantum ap-proaches to reactive scattering. More recently, Connor (20) has reviewedreactive molecular collision calculations up to early 1979. This reviewcontains extensive literature references. Earlier reviews of quantumreactive scattering include the 1976 review by Truhlar & Wyatt (21),concentrating on the H+H: reaction, and the 1975 review by Micha(22), covering quantum reactive scattering.

Exact, fully three-dimensional quantum reactive calculations are stillsufficiently difficult that very little new work has been reported in thelast year. Here we concentrate on approximation methods, becausedevelopment of reliable approximate techniques is essential if we are totreat reactions quantum mechanically.

The final section briefly describes progress in several allied fields.There is considerable activity in statistical theories of reaction, includingtransition state theories and phase space theories. Truhlar (23) hasrecently reviewed transition state theories, and Pechukas’ DMCB review(24) presents a general and very readable discussion of statistical reac-tion theories. At higher collision energies, collision-induced dissociation(CID) becomes a significant alternative channel for reaction. A reviewof the classical treatment of CID processes is given in AMCT by Kuntz(25), and the quantum treatment is presented there by Diestler (26).Finally, dynamicists have become increasingly interested in whether ornot intense electromagnetic fields can affect reaction dynamics. Weconclude this review by discussing recent developments in this area.

www.annualreviews.org/aronlineAnnual Reviews

Page 4: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

404 WALKER & LIGHT

CLASSICAL TRAJECTORY CALCULATIONS

Far more calculations of dynamical processes employ a classical trajec-tory approach than any other. Classical trajectory methods have beenapplied extensively to molecular systems ever since modem electroniccomputers have been available. The ease with which most problems canbe adapted to the modem minicomputer environment ensures theviability of the method for the indefinite future. Almost any reviewarticle describing theoretical treatments of molecular dynamics em-phasizes the contribution of classical mechanics. The classical approachitself has been reviewed often; in 1974 Porter (27) reviewed the applica-tion of classical methods from their inception up to the early 1970s. Inthe introduction, we mentioned several of the most recent reviews ofclassical methods (13-15). Classical calculations were also reviewed Connor (20).

The trajectory approach to dynamics earns its popularity because it isthe only method currently available that is general enough to treat thewide range of problems facing the dynamicist. Trajectory calculations ofatom-diatom reactions now appear routinely; the treatment today isbasically that originally described by Karplus et al (28) in 1965.

In the last year alone, many trajectory calculations have appeared.Because of its basic importance to the dynamics community, calcula-tions for the H+H2 exchange reaction continue to receive frequentattention. Osherov et al (29) provide a sketchy account of calculationsfor reaction from the initial v = 1 excited state of H2. Mayne & Toen-nies (30) and Mayne (31) also report calculations for the 2reaction, using the most recent and accurate HHH potential surface ofTruhlar & Horowitz (32). This surface is an analytic fit to the ab initiocalculations of Liu (33) and Siegbahn & Liu (34).

Trajectory calculations are frequently used to compare theoreticaldynamics with experimental results, at least at the rate constant level.Several calculations in this category have been reported in the last year.In a continuing study of the He +H~- reaction, Zuhrt et al (35) considerthe effectiveness of reactant translational, vibrational, and rotationalenergy in promoting reaction. Gittins & Hirst (36) report calculationsfor a single surface model of the N+ + H2 reaction, in which the surfacehas an extremely deep well for C2~ geometries. Schinke & Lester (37)report calculations for the O(3P)+ 2 reaction, using t heir own ab i nitiopotential surface. Luntz ei al (38) report a study of the O(~D)+H2reaction, in which both trajectory results and experimental productrotational state distributions differ significantly from phase space theorypredictions. Herbst et al (39) studied the ion-molecule reaction O- + 2

www.annualreviews.org/aronlineAnnual Reviews

Page 5: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 405

at relatively high energies, using the multisurface "trajectory-surface-leaking" model of Preston & Cohen (40). Jaffe (41) reports theoreticalrate constants for the atmospherically important CIO + O reaction, andPersky (42) reports rate constants, isotope effects, and energy disposalfeatures for the CI + HD reaction.

Trajectory calculations are also used to investigate the relationshipbetween scattering dynamics and potential surface effects or statisticaltheories. Polanyi & Schreiber (43) use trajectory results to test informa-tion theoretic predictions of branching ratios for the F + HD reaction.Polanyi & Sathyamurthy (44, 45) use trajectories to study the relation-ship between features of the potential surface and dynamics for modelsystems. Duff & Brumer (46) use trajectories as a tool for investigatingstatistical behavior in two reactions involving long-lived complexes, theK + NaC1 and H + IC1 reactions.

In the future we will undoubtedly see an increasing number ofcalculations involving more than three nuclei. Recent examples are thefour-center reactions of HBr + C12 and HBr + BrC1, studied by Brown etal (47). In this calculation, both reactants and products are diatomicmolecules, but calculations are more difficult to interpret when tri-atomics and higher polyatomics are considered. We discuss some ofthese additional difficulties later, but a flavor for the problems involvedcan be detected in the unimolecular decomposition reaction studies ofHase et al (48) for C2H5 and of Grant & Bunker (49) for ethane.

Classical vs Quantum Studies

Apart from the practical problems that must be faced when implement-ing a classical trajectory calculation, there is always the conceptualquestion: How reliable is classical mechanics for treating nuclear mo-tion in molecular systems? Truhlar & Muckerman (14) present a discus-sion of this point in their review. At the most detailed dynamical level, afailure of classical mechanics can usually be attributed to circumstancesin which the important dynamics is associated with motion char-acterized by a large deBroglie wavelength. For this reason, classicalmechanics does poorly at predicting reaction thresholds and missesresonance features altogether. On the other hand, when the dynamics ofinterest involves less detail (more averaging), classical mechanics tendsto do a better job.

Of course, the best way to see how well a classical calculation worksis to solve the problem both classically and quantum mechanically.Classical vs quantum studies are very popular in the literature. Becauseexact quantum reactive calculations are primarily restricted to collinear

www.annualreviews.org/aronlineAnnual Reviews

Page 6: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

406 WALKER ~ LIGHT

geometries, virtually all collinear classical calculations are performednowadays purely to provide a data base for classical vs quantumcomparisons. Recent examples of this type of work may be found inpapers by Truhlar and co-workers (50-53) for the collinear I+H2,H+C12, H+H2, F+H2, and CI+H2 systems, and in the papers ofConnor and co-workers (54-56) for the collinear X+F2 (X=Mu, H, D,T) system. Hutchinson & Wyatt (57) have also studied collinear F+H2 reaction to investigate both classical vs quantum andclassical vs statistical behavior. In connection with statistical behavior inthis system, see also the related paper by Duff & Brumer (58).

Practical Considerations

Having decided to use classical trajectories to study a particular dy-namical system, we must still face several practical considerations. Forthe basic single surface A+BC reaction, the methodology (13-15) known, and the problems discussed below are generally not serious. Butfor polyatomic systems, many of these problems must be reconsidered[see References (48, 49) and references cited therein].

We consider classical trajectory calculations in three stages.

1. We must decide how to specify initial conditions for each trajectory.2. The system is allowed to evolve according to the classical equations

of motion.3. We decide when each trajectory is over and the final coordinates

and moments are analyzed to see what has happened.

Each stage of the c:dculation presents its own unique problems. Wenow consider some of these problems, as they occur for A + BC reactivesystems.

1NITIAL CONDITIONS The conventional wisdom in molecular scatteringcalculations (whether reactive or not) is to prepare the molecular speciesso that its rotational and vibrational action variables correspond tointegral quantum numbers (times/~) for the associated quantum system.Calculations done this way are termed quasiclassical trajectory calcu-lations; this is the method almost universally used in molecular scatteringsystems. It is not difficult to prepare initial conditions this way fordiatomic molecules; however, in treating triatornic and higher poly-atomic species, this part of the problem becomes increasingly difficultbecause nonseparability of the internal molecular degrees of freedombecomes important. At higher energies it may no longer be possible tofind good classical action variables for such systems.

A second problem with initial conditions is statistics. After quantizingthe classical action variables, the uncertainty principle requires all

www.annualreviews.org/aronlineAnnual Reviews

Page 7: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 407

values of the conjugate angle variables to be equally likely. The com-mon practice is to select these variables randomly in a multidimensionalMonte Carlo integration procedure. The Monte Carlo approach isnormally preferred to nonrandom initial condition selection methods(15) because (a) the associated statistical error is easily evaluated (b) the accuracy of the Monte Carlo method can be systematicallyimproved by running more trajectories, thus accumulating all the re-suits. (If insufficient accuracy has been obtained in the nonrandomselection method and a greater number of trajectories has to be in-tegrated, then the trajectories already computed are discarded.)

Statistical problems arise when the desired dynamical event occurswith a low probability. In such cases, many trajectories are requiredbefore enough of interest are found (for example, reactive trajectories atlow collision energies). Several strategies can then be employed toimprove the initial condition sampling problem. These strategies includeimportance sampling procedures, stratified sampling procedures, andthe nonrandom initial condition selection method. These methods arediscussed in detail in recent reviews (13-15); discussion and use of theimportance sampling method may be found in several papers (43,60-61). Furthermore, Suzukawa & Wolfsberg (62) recently presentedevidence for the improved convergence properties of the nonrandominitial condition selection method.

INTEGRATING TRAJECTORIES Once an initial condition algorithm hasbeen selected, integrating the equations of motion is usually straightfor-ward. In terms of computer time, however, it is by far the mosttime-consuming part of the study. Many numerical methods for in-tegrating the usual differential equations are discussed in earlier reviews(14, 15).

Collisions typically begin and end at separations large enough so thatthe interaction between unbound particles is negligible. Unfortunately,the interesting dynamics (reaction and energy transfer) occurs duringthe relatively short time that the particles are close together. Therefore,a significant part of the trajectory effort is expended in the nearlynoninteracting region. Billing & Poulsen (63) propose an asymptoticperturbation theory to reduce the expense of bringing the collisionpartners close together. Their approach uses action angle variables andis therefore more difficult to apply to polyatomic molecules.

A problem associated with trajectory integration arises whenever theinteresting dynamics occurs slowly on a time scale defined by the fastestmolecular motions followed. In such cases, a very large number oftrajectory integration steps become necessary. A standard procedure toverify the accuracy of (selected) individual trajectories is to run them

www.annualreviews.org/aronlineAnnual Reviews

Page 8: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

408 WALKER & LIGHT

backwards from their final conditions to see how well the initial condi-tions are reproduced. Duff & Brumer (46, 58, 64-66) point out thatmany classical systems exhibit instabilities (exponential divergence be-tween trajectories initially close together in phase space) that precludereliable integration for extremely long times--times so long that trajec-tories cannot be back-integrated successfully. Related to this problemare questions of ergodic and stochastic behavior in molecular systems(64, 65, 67-72), which are beyond the scope of this review. Thesequestions are now entering the reactive collision literature (46, 57, 58)and more work in this area is sure to follow in the 1980s.

FINAL CONDITIONS Although it is straightforward to analyze classicaltrajectory data and obtain estimates of averaged quantities such as rateconstants and vibrational energies, there are no reliable methods forobtaining accurate quantum state-to-state cross sections from thesedata. Classical mechanics samples a continuous distribution of finalaction variables, whereas quantum mechanics produces only discretedistributions of action variables. To express the classical data in aquantum mechanical language, we must somehow "’discretize’" theclassical distribution. The rare trajectory with quantized final conditionsis of central importance to the semielassical S-matrix theory of collisions(14, 73-75); unfortunately, as the phase space of the system increases,the task of finding these very special trajectories becomes insurmount-ably difficult.

Several methods are commonly used to quantize the final conditionsobtained from trajectories. One is the histogram method, which takesthe final quantum state as the state with quantum numbers closest tothe actual nonintegral action variables obtained classically. A secondapproach finds the quantum final state distribution which reproduces(some of) the moments of the distribution obtained classically. Bothmethods have been reviewed in detail (13, 14). A third approach, calledsmooth sampling, assigns trajectories to more than one quantum bin inproportion to the difference between the actual final actions and thenearest integral actions (55, 76-79). How well these procedures actuallywork in practice is usually the central objective of the classical vsquantum comparison studies referred to earlier (50-57).

Calculations Invoh~ing Polyatomics

Calculations involving triatomic and higher polyatomic species are ofincreasing interest. In these larger molecules the distribution of energyamong the molecular vibrational degrees of freedom may affect dy-namical processes such as reaction and energy transfer. The problem of

www.annualreviews.org/aronlineAnnual Reviews

Page 9: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 409

quantizing final states (and also of preparing initial states) becomesmuch more difficult when molecules more complicated than diatomicsare treated. Unfortunately, the quantity we most want to study inpolyatomics is the energy distribution within the molecule.

In order to minimize the computational expense of trajectories, wewant to use a coordinate system in which the equations of motion areefficiently formulated. It has been recognized for some time (27, 80) thatthe most efficient coordinate system is the space-fixed (SF) Cartesianreference frame. The task of preparing initial conditions and analyzingfinal conditions requires the canonical transformation between SFCartesian coordinates and coordinates which directly express the de-grees of freedom of the polyatomic in more separable terms. First, wewant to separate the overall dynamics into translational, rotational, andvibrational components. Separating overall translational motion is easy.However, Coriolis interactions make it impossible to separate cleanlythe rotational and vibrational motions of a polyatomic molecule. Byviewing the motion of the polyatomic from a body-fixed (BF) referenceframe satisfying the Eckart conditions (81-83), we minimize the Corio-lis coupling between overall rotation (carried completely by the BFframe) and vibration. Before and after collision, the total energy of thepolyatomic is constant in time, and so it is always possible to define thetotal energy transfer in a collision. However, the Coriolis interactioncomplicates apportioning this energy transfer between rotational andvibrational motions, because the rotational and vibrational energies ofthe polyatomic vary with time. By time averaging (83) the rotational andvibrational energies over several (of the lowest frequency) vibrationalperiods, we can define rotational and vibrational energy transfers.

At this point, we can now turn our attention to distributing theinternal energy among the 3N-6 vibrational degrees of freedom. Think-ing quantum mechanically about the states of a polyatomic, we areinclined to speak of internal excitations in terms of the system normalmodes. However, the idea of normal modes to the practitioner ofquantum mechanics differs significantly from the classical definition ofnormal modes. Classically, normal modes refer to the coordinate repre-sentation in which there are no off-diagonal quadratic terms in themultidimensional Taylor series expansion of the molecular force fieldabout the equilibrium conformation of the molecule. The practitioner ofquantum mechanics, in referring loosely to normal modes, really meansthose molecular degrees of freedom in which excitations are maintainedindefinitely in the absence of perturbations (classical constants of mo-tion). These degrees of freedom are classical normal modes only in thelimit of infinitesimally small excitations. As soon as finite excitations are

www.annualreviews.org/aronlineAnnual Reviews

Page 10: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

410 WALTZER ~ LIGHT

allowed classically (even excitations so small that they correspond toquantum zero point energies), then normal modes are no longer classicalconstants of motion. Instead, the true constants of motion are expressedclassically in terms of the "gc~od" actions and action angle variables.Finding these variables is a subject of intense activity beyond the scopeof this review. For recent work in this area, the reader should consultthe papers of Schatz and co-workers (84-87), Marcus and co-workers(88-91), Miller and co-workers (92-94), Sorbic (95), Sorbic & (96), Colwell & Handy (97), and Jaffe & Reinhardt (98). These workershave concentrated on triatomics; at present the methods appear practi-cal enough that it shoul.d be possible to perform quasiclassical trajectorycalculations for systems in which the most complicated polyatomic is atriatomic. How these methods will perform for higher polyatomics ispresently unknown. Without a definition of good action variables,trajectory calculations will be limited in their ability to describe the roleof energy distribution within a polyatomic molecule.

QUANTUM REACTIVE CALCULATIONS

Exact Methods

The development of practical, methods capable of providing a dynami-cally accurate (quantum) description of molecular reactive scatteringprocesses has been a field of considerable activity within the last 15years. At first, accurate quantum calculations were restricted to A + BCreactions in the one-dimensional collinear (1 D) world, whereas there arenow several accurate, fully three-dimensional (3D) quantum calcula-tions. Even in the last four years, progress in quantum reactive scatter-ing has been reviewed often. Because of the central role the H+H2exchange reaction has played in this area, Truhlar & Wyatt reviewedboth theoretical and experimental work on this reaction in 1976 (21).Further progress and some results for the 3D H+H2 and F+H2reactions were again presented by Wyatt (99) in 1977. The 1979 reviewof reactive calculations by Connor (20) considers primarily the results quantum collinear calculations, and Light’s 1979 review (17) presents formal introduction to the quantum reactive problem and discussesseveral formal considerations appropriate to the quantum reactive prob-lem. By far the most comprehensive review of theoretical methodsavailable to date is that of Wyatt (18), who presents a comparativediscussion of similarities and differences in current 3D quantum scatter-ing theories. After a brief introduction, we consider in this section bothexact collinear and exact three-dimensional quantum approaches andpoint out newly proposed methods that may extend our capabilities in

www.annualreviews.org/aronlineAnnual Reviews

Page 11: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 411

the future. Only very limited work has been done on quantum systemsmore complicated than the single surface A + BC reaction, and so herewe consider only this system.

Although other methods have been tried, the most productive currentapproach for accurate quantum calculations of chemical reactions con-sists of a coupled-channel solution of the Schrrdinger equation, writtenin natural collision coordinates (NCC). In the coupled-channel for-malism, all degrees of freedom except one are expanded in a target basisset of square integrable functions. The coupled-channel equations arethen solved numerically for motion in the final degree of freedom,referred to variously as either the reaction, propagation, scattering, ortranslational coordinate. Algorithms for solving the coupled equationstypically require computer times that go asymptotically as N3, where Nis the number of channels in the wavefunction expansion. This N3

dependence on basis-set size defines one primary goal of any closecoupling theory of quantum reactive scattering--to devise a methodthat keeps the basis set as small as possible.

Of course, any quantum close-coupling study, reactive or not, requiresa large basis expansion if particles with large masses, strong couplingand/or high energies are involved. Quantum dynamics is mostly con-cerned with light-mass low-energy systems anyway, because these pre-sumably show the most pronounced quantum effects.

There are, however, problems peculiar to reactions that require spe-cial strategies for minimizing basis-set size requirements. One suchstrategy is the choice of natural collision coordinates. Because of thedifference in asymptotic motions for reactants and products, we mustuse different coordinates for them. Natural coordinates transform (atleast some of) these coordinates smoothly. Their use also minimizes thetarget basis-set size because they can be chosen to represent the "natu-ral" motions of the three-body system (translation, rotation/bending,and vibration) everywhere along the scattering coordinate. The more"natural" the collision coordinates are, the more effectively we canchoose a small internal basis to represent bound-like motion.

The second common strategy is prompted by the following considera-tion. For reactants, we want an internal basis that efficiently representsthe molecular BC vibrational and rotational motions, whereas for prod-ucts, the internal basis should be equally efficient in representing the(different) vibrations and rotations of the (AB or AC) productmolecule(s). Consequently, it is common practice to deform the internalbasis, either continuously or by segments, along the reaction coordinate.This procedure is unnecessary in nonreactive systems, where the defor-mation of the target induced by the projectile is much less severe.

www.annualreviews.org/aronlineAnnual Reviews

Page 12: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

412 WALKER ~ LIGHT

Dynamical approximations for reactive scattering are interesting in sofar as they effect a reduction in the number of coupled equations to besolved simultaneously. Approaches such as these, including reactiveversions of the coupled states (CS) or infinite order sudden (IOS)approximations, are discussed later.

COLLINEAR CALCULATIONS The most drastic (and most effective) wayto reduce the size of the reactive basis set is to restrict the nuclei tomove along a common line. Today, collinear quantum calculations forthe A + BC reaction are commonplace and quite inexpensive. So long asthe mass combinations and energies of interest do not require too largea basis set, the quantum calculation is often cheaper than the corre-sponding quasiclassical trajectory calculation would be. For certainmass combinations (H-L-H, L-H-I-L and H-H-H, where L=light,H=heavy), large basis expansions are necessary in an NCC close-coupling approach. The H-L-H mass combination is especiallytroublesome because it generates an extreme (mass dependent) skewingof the collinear configuration space. The NCC in such a skewed spacedo not represent effectively the natural motions of the triatomic in thestrong interaction region; large basis expansions result from the coordi-nate defect. The L-H-H and H-H-H mass combinations are difficultbecause of the preponderance of heavy nuclei.

Collinear calculations, in spite of the severity of the dynamicalapproximation, are important for a variety of reasons. New coordinatesystems and new theoretical approaches are most easily explored forcollinear systems. For example, McNutt & Wyatt (100) discuss a gener-alized hyperbolic coordinate system, an extension of coordinates sug-gested by Witriol et al (101), which are not multivalued for ABCconfigurations typical of A+ B + C. Coordinates such as these may beuseful for coupled-channel treatments of collision-induced dissociation.A variety of alternative theoretical treatments of quantum reactivescattering have been introduced for collinear systems. Garrett & Miller(102) and Adams & Miller (103) have proposed an exchange kernelapproach that avoids the use of natural collision coordinates at theexpense of solving coupled integro-differential equations instead ofordinary coupled equations. Rabitz et al (104) and Askar et al (105)treat the collinear reactive problem using the finite-element method.This method is widely used in other fields (106), and also avoids usingnatural collision coordinates. Askar et al (107) also suggest that combination of the finite-element method (in the strong interactionregion) with the coupled-channel method (in the asymptotic region)may be computationally advantageous.

www.annualreviews.org/aronlineAnnual Reviews

Page 13: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 413

A variety of numerical methods have been proposed to solve reactivecoupled-channel equations. We are naturally inclined to prefer theR-matrix propagation method (108-111). It is a fast and generalnumerical method, easy to program, stable to the inclusion of closedchannels, and automatically produces unitary S-matrices with arbitrarilysmall basis expansions. There are, of course, several alternative meth-ods, and the one actually used is mostly a matter of convenience. Thosecommonly used include Gordon’s (112) propagation method, the in-tegral equation method proposed by Adams et al (113), and the state-path-sum method of Manz (114, 115) and Connor et al (116). state-path-sum method and the R-matrix propagation method are simi-lar in that they are both invariant imbedding solutions (16) to theSchr/Sdinger equation; both methods obtain the global dynamics byassembling solutions of many boundary value problems, each localizedin small regions (sectors) of configuration space. Recently, Basilevsky Ryaboy (117) have also presented a sector-oriented method for thecollinear reactive problem; to us, their approach seems very similar toour own. Because they do not factor the scattering wavefunction optim-ally (118), their coupled equations are complicated by first derivativeoperators. They then introduce several approximations, finally obtaininga tractable method. With comparable numerical effort, the R-matrixmethod gives accurate solutions.

We should also note that several new approaches for solving coupled-channel equations have been proposed. These have been proposed fornonreactive systems, and it would be interesting to investigate themwithin a reactive framework. Johnson (119) presented a renormalizedNumerov algorithm that shares many of the computational advantagesof the multichannel version of his log-derivative method (120, 121).Thomas (122) proposed an iterative method to solve coupled equations,which should be useful when extremely large basis expansions (on theorder of 200 channds) are necessary. His method obtains each colunmof the S-matrix independently. Stechel et al (123) and Schmalz et (124) proposed an R-matrix method using nonorthogonal coordinatesand a nonorthogonal basis to study the charge exchange problem. Thisproblem is interesting for quantum reactions because of its similarity tothe H-L-H mass combination. Finally, Leasure & Bowman (125) havedescribed a method based upon solving the Laplace transform of theSchrrdinger equation.

Because collinear quantum calculations are feasible, much work inthis area is devoted to comparing exact quantum dynamics to dynamicsobtained by approximate methods. We have already mentioned thepopularity of these calculations for estimating the reliability of qua-

www.annualreviews.org/aronlineAnnual Reviews

Page 14: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

414 WALKER ~: LIGHT

siclassical trajectory calculations (50-57). In addition, exact collineardynamics is useful in evaluating agreement with information theorymethods, transition state theories, and phase space theories. Much ofthis work has been described in Cormor’s review article (20) and discussed later in this review.

THREE-DIMENSIONAL CALCULATIONS A fully converged coupled-channelcalculation of the 3D A+ BC reaction still represents the major chal-lenge to the dynamicist. Theoretical approaches to the problem havebeen described by Wyatt and co-workers [see References (18, 99) andthe references cited therein], Schatz & Kuppermann (127), and Lightand co-workers (128, 129). These methods have in turn each producedconverged calculations only for the low energy H+ H2 reaction (130-134); it is gratifying that all calculations for this reaction are in substan-tial agreement with each other, especially at threshold energies.

Only a few accurate 3D studies have been reported since Wyatt’sreviews (18, 99). Using a reactive CS formalism, Redmon & Wyatt (135)have studied the F + H2 reaction, with emphasis on resonance behavior.The pronounced collinear resonance behavior is significantly broadenedin the 3D reaction. Wyatt (136) observed 3D F+H2 resonance mani-festations at collision energies consistent with recent experimental ob-servations of Sparks et al (137) for the reactive differential cross section.Both the theoretical and experimental observations represent state-of-the-art achievements, and the similarities of these observations is re-markable.

In addition to these F+H2 results, Redmon (138) presented pre-liminary (unconverged) results for the extremely difficult H +02 reac-tion. The H+H2 calculation of Walker et al (134) compared results calculations on the Porter & Karplus (139) 3 potential s urface with t heSiegbahn & Liu (34) and Truhlar & Horowitz (32) surfaces. Low energytotal reactive cross sections for this surface were also given. Clary &Nesbet (140, 141) have used these programs to investigate the evolutionof entropy in the product internal state distribution along the reactioncoordinate, with emphasis on applying information theory to 3D reac-tions.

There will not be, in the near future, many converged 3D quantumreactive calculations. Basis-set size limitations preclude the accuratestudy of reactions involving nuclei other than hydrogen and its isotopes.These few accurate 3D calculations will serve as reference points uponwhich to investigate approximations capable of providing dynamics ofsatisfactory accuracy at computationally feasible expense.

www.annualreviews.org/aronlineAnnual Reviews

Page 15: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 415

Approximate Quantum Theories of Reactive Scattering

Due to the formidable formal and computational problems associated¯ ~with exact quantum treatments of reactive scattering, it is not surprisingthat there is considerable current research devoted to finding approxi-mate quantum treatments that are relatively simple and accurate. Wyatt(19) has written an excellent review of the status of these efforts as late 1978. Although we shall concentrate here on results of the past year,because of the importance of the area, we shall review some of theearlier work as well. Most of these efforts fall into three basic cat-egories: (a) distorted wave Born approximations (DWBA) or adiabaticperturbation approximations, (b) overlap or Franck-Condon models,and (c) dimension-reducing approximations such as Jz-conserving(centrifugal sudden, coupled states, CS are aliases) or other suddenapproximations that utilize exact quantum methods on decoupled Ham-iltonians. In addition, we discuss a few other approaches that do notreadily fit into the above categories.

DISTORTED WAVE BORN APPROXIMATION Although the DWBA was de-veloped for reactive scattering and used earlier, Choi & Tang (142) 1974 showed that the integrals over the Euler angles in the matrixelements could be done analytically, thus greatly reducing the effortrequired to apply the approximation. We briefly review their formula-tion here before commenting on recent results. The DWBA is based onthe exact equation for the differential cross section in terms of theT-matrix elements

do~, i~,~ k~

where ~ and fl refer to channels (A + BC) and (Co-AB), respectively,and the T-matrix element is defined as

where X~-) is the solution of the Schr6dinger equation (with incomingboundary conditions) with the approximate Hamiltonian

H/s= H- V~ 3.

and ~k~(+) is the solution (with outgoing boundary conditions) of thecomplete Hamiltonian, H. To be exact, %(+) must contain all accessibleinternal states, although X~-) may, with a judicious choice Of V~, containonly a single internal state. The DWBA consists, then, of replacing the

www.annualreviews.org/aronlineAnnual Reviews

Page 16: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

416 WALKER ~ LIGHT

exact scattering wavefunction, ~2+~, by a simpler approximate wave-function satisfying the same boundary conditions. In practice ff~+~ hasbeen replaced by X~+~, a wavefunction corresponding to a single inter-nal state only. Thus the DWBA contains the coupling potential only tofirst order, although both X~÷~ and Xg-~ may correspond to waves withfull potential scattering and adiabatically distorted internal states. Withthe Euler angle integrations performed analytically, Choi & Tang (142)reduced the numerical problem from an extremely lengthy five-dimensional integral to a much more manageable sum of three-dimensional integrals. The calculations are nevertheless nontrivial, be-cause for each state-to-state cross section at a given scattering angle, (a)the adiabatic internal states must be determined at each value of thescattering coordinate, (b) the elastic scattering wavefunction must determined for this state, and (c) the three-dimensional integrals mustbe done for all angular momentum components of the initial and finalorbital angular momenta.

Remarkably, this approximation works very well for the H +H2 (143)and D+H~ (144) reactions at low energies. It has been applied to theF+H2 reaction (145) using both unperturbed and adiabatically per-turbed internal states, but with H2 initially in the ground (v=0, j=0)state. They found that the adiabatically perturbed internal states pro-duced much larger reactive cross sections, but that the product internalstate distribution was largely unaffected. Although the internal energydistribution for the related F+D2 reaction appears to be in goodagreement with the experimental results (146), no quantitative measureof accuracy is available.

Clary & Connor (147) used the DWBA on the H+F2 (t~=0, j=0)reaction and compared it with both quasiclassical trajectory results andcollinear close-coupling results. They also compared the full adiabaticinternal states calculation to those keeping either the initial or finalinternal state unperturbed. They found quite good agreement in finalstate distribution between experiment (148)--the quasiclassical, the fulladiabatic DWBA, and the "static product" DWBA--but poor agree-ment when the reactant internal state was not allowed to distort adia-batically in the DWBA calculation. The extent of agreement in this caseis truly surprising because the H + F2 reaction is highly exothermic andone would expect very significant vibrational coupling in the exitchannel. Given the developments toward more general applicability ofthe DWBA method and the encouraging results to date, it is likely to betested, developed, and used more widely in the next few years.

FRANCK-CONDON APPROACHES We consider here a number of differentapproaches, most of which utilize a Born approximation at some point.

www.annualreviews.org/aronlineAnnual Reviews

Page 17: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 417

They are bound together by the common expression of the T-matrixelement in the form

where X~+) and X~-) are (approximations to) scattering wavefunctionslocalized in reactant and product channels and the matrix element ofthe complicated operator, ~’~, is usually further reduced to an overlapcalculation (X(o-)IX(a+)) of the nuclear wavefunctions.

The formal approach was introduced by Halavee & Shapiro (149) andshortly followed by Schatz & Ross (150-152). Although this work wascovered in Wyatt’s review (19), there has been intense activity sincethen. Concise reviews of the assumptions and development of the theoryare given in two recent articles (153, 154), together with extensions andevaluations. The basic assumptions common to Franck-Condon (FC)theories are simple to state. The exact T-matrix element for electroni-cally adiabatic scattering on a single electronic surface (Eq. 2) considered (and is, in fact) a Born-Oppenheimer approximation to thesolution of the full Schrrdinger equation (including electronic coordi-nates) in the adiabatic approximation. Thus, as Halavee & Shapiro (149)pointed out, one can formally use a diabatic (or quasi-adiabatic) (155)electronic representation from which two nuclear surfaces result, corre-sponding to reactant and product nonreactive surfaces with a remainingelectronic perturbation term [which would produce the adiabatic (lower)surface when diagonalized]. The T-matrix element then takes the form(Eq. 4). Franck-Condon type theories result when these two surfaces areused to define the reactant and product nuclear scattering and when theremaining perturbation is taken to first order only. If the electronicmatrix element is slowly varying with nuclear coordinates, a pure FCoverlap approximation results (153):

where Met is formally the electronic matrix element between the quasi-adiabatic electronic states yielding the nonreactive surfaces for reactantsand products.

Several approximations are made in deriving the Franck-Condontheory as represented by Eq. 5. They are (153): 1. neglect of all higherelectronic states and transitions to them, 2. neglect of nuclear electroniccoupling and nuclear kinetic energy coupling terms, and 3. the assump-tion that the electronic matrix element is slowly varying over the regionwhere the wavefunctions X~+) and X~-~ overlap. In addition it appears(154) that 4. a weak coupling approximation between the two electronicsurfaces is made (in 2 above), because the T-matrix element (Eq. 5) only first order in the coupling potential.

www.annualreviews.org/aronlineAnnual Reviews

Page 18: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

418 WALKER & LIGI-|T

From the Franck-Condon theory outlined above a number of Franck-Condon models result from further approximations (149-154, 156). earlier versions, very crude approximations to evaluating the channelwavefunctions X(~+) and X~-) were coupled with approximate evalua-tions of the multidimensional overlap integrals. Simple analytic ap-proximations could then be obtained, particularly for the relative inter-nal energy distributions of products. Two recent studies on collinearreactions report on the adequacy of these further approximations.

Vila et al (153) constructed the quasiadiabatic channel potentialsurfaces from LEPS and anti-LEPS surfaces and a coupling functionhighly localized near the col. They then evaluated exactly the (two-dimensional) channel functions X~+) and X~-) on these surfaces bysolving the close-coupling equations numerically and evaluated theoverlap integral in Eq. 5 numerically. They compared, then, three sets ofproduct vibrational distribution results--exact quantum close-coupling,"exact" Franck-Condon theory, and a simple Franck-Condon model(150-152). For the three reactions [H2(v=0)+F, D2(v=0)+F,H+C12(v=0)], they found very substantial agreement. The most probablevibrational state was almost always correct and the general shape of thedistributions was correct. Although the "exact" FC results were some-what better than the simple FC model results, even the latter are ingood qualitative agreement. The results also appear relatively insensitiveto the only free parameter in the model--the range of the couplingfunction.

Fung & Freed (154) generalized the Franck-Condon theory to poly-atomic molecules (provided a single reaction coordinate is well defined).They reduced the multidimensional overlap integral to a "rapidly con-vergent series of one-dimensional bound-continuum integrals," per-formed the adiabatic to quasi-adiabatic surfaces decomposition, and(again) applied the results to a model collinear system for which exactquantum calculations could be performed (D + IH). The channel wave-functions were approximated using a forced oscillator model for thehalf-collisions, as well as various simpler (Airy function, delta function)approximations. They also found qualitatively good agreement betweenthe exact quantum and their best FC model results for the relativeproduct vibrational distribution. More interesting, perhaps, is that theyreported values of the absolute reaction probabilities. [Vila et al (153)did not, even though the data were easily available from their work.]Fung & Freed (154) found that the absolute transition probabilitieswere extremely sensitive to the electronic coupling functions assumed,varying over several orders of magnitude (~106), even though therelative distribution was rather insensitive. This sensitivity is presumably

www.annualreviews.org/aronlineAnnual Reviews

Page 19: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 419

attributable to assumption 4 or 2 in the original derivation in whichweak coupling is assumed between the quasiadiabatic surfaces. Becausethese calculations were performed at low energies (for which tunnelingis important), a high degree of sensitivity to the treatment of the colre#on is not surprising. These results, however, do indicate that the FCtheory is currently not capable of reliably determining the.~magnitudesof reaction cross sections, irrespective of further model assumptions.

Other aspects of Franck-Condon models have also been scrutinizedrecently. Vila et al (156) investigated the angular distributions of reac-tion products in 2D and 3D models for.:several potential and kinematiccoupling regimes. They used only simplified models, however, wffhangular wavefunctions given either by free rotor or (adiabatic) hinderedrotor functions. Again the calculated angular distributions (hinderedrotor basis) qualitatively agree with exact quantum and classical trajec-tory calculations where available, and again absolute magnitudes werenot presented.

Very recently, Wong & Brumer (157) derived a Franck-Condon typetheory from a Faddeev decomposition of the scattering wavefunction. Apotential advantage of this approach is that three-channel A + BC~AB+ C, AC+ B reactions can easily be incorporated, whereas the quasi-adiabatic decomposition of the potentials in the "standard" FC theoryis quite arbitrary. Wong & Brumer reduce their equations to a simpleform, the key to which is the solution of linear algebraic equations inmomentum space. Comparison with exact quantum collinear results(158) for H+C12 shows excellent agreement in the predicted relativevibrational products distribution, and again no absolute reaction proba-bilities were given.

The Franck-Condon approach to reactive scattering has been shownin the last few years to be qualitatively correct in predicting relativeinternal and angular distributions for simple re.active systems. Even inits simplest forms, it provides useful results. However, even in the mostaccurate forms, it appears to be completely unreliable in predictingabsolute transition magnitudes and therefore absolute cross sections.This may be a defect basic to the theory in the sense that localizing theoverlap of the quasiadiabatic scattering solutions requires large interac-tion potentials between the quasiadiabatic surfaces and thereby invali-dates perturbation theory. Extension of the interaction region, whilereducing the perturbations, will degrade the accuracy of the overlapapproximation. Thus the approximations inherent in the FC theory maylimit its accuracy. The optimal choice of surfaces has been consideredby Mukamel & Ross (155) and a first order K-matrix approximation include higher orders in the reactive perturbation was given. Its use by

www.annualreviews.org/aronlineAnnual Reviews

Page 20: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

420 WALKER ~ LIGHT

Fung & Freed (154), however, did not appear to help the absoluteaccuracy. Until this can be remedied, the FC theory must be consideredvaluable as a guide to internal energy distributions only.

REACTIVE DECOUPLING APPROXIMATIONS Because very many close-coupled equations are required for the exact description of most atom-diatom systems, much attention has been given recently to approxima-tions that reduce the dimensionality of the coupled equations and stillpermit the evaluation of state-to-state cross sections. Two of theseapproximations, the J,-conserving approximation and the infinite ordersudden approximation (IOS), have recently been applied to reactivescattering.

The only reviews to date that consider applications of Jz-conservingapproximations to reactive collisions is again that of Wyatt (18, 19). TheJz-conserving work of Elkowitz & Wyatt (159) and Kuppermann et (160) on H+H2, and of Redmon & Wyatt (161) on F+H2, is reviewedthere.

The centrifugal sudden (CS) approximation was introduced by Mc-Guire & Kouri (162) and Pack (163) for inelastic scattering (164). inelastic scattering the CS approximations can be made directly byreplacing the orbital angular momentum operator in a space-fixedcoordinate system by an average eigenvalue, [([+ 1)h2. Cross sectionsare then obtained by using "/-labeled" S-matrix elements and recouplingwith appropriate sums over Clebsch-Gordan coefficients (165-168).Because body-fixed coordinates are used in reactive scattering, only theprimary assumption of the CS approximation (the neglect of Coriolis-type coupling) was used by Wyatt et al and Kuppermalm et al. Theydecouple the original complete set of coupled equations into indepen-dent sets, one for each value of the projection of the total angularmomentum onto the body-fixed axis (165). Additional approximationsmay then be made to simplify the equations further (166-168).

Generalizing this approximation to reactive scattering is made dif-ficult because the body-fixed axes of reactants and products must differ.Thus, the approximation, which imposes a conservation condition in thebody-fixed frame, is not the same for reactions and products, and leadsto continuity problems. Elkowitz & Wyatt (159) and Kuppermann et (160) resolved this problem in different ways. The former, using complete set of natural collision coordinates, define a "body-fixed"frame that rotates smoothly with the scattering coordinate from that forreactants to that for products. In this way, the CS approximation isimposed uniformly throughout the reaction, albeit at the cost of ne-glecting some comple.x effective potential terms.

www.annualreviews.org/aronlineAnnual Reviews

Page 21: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 421

Kuppermann et al (160) essentially switch coordinates on a surface,matching wavefunction and derivatives there. Continuity can onlybe preserved by coupling the set of solutions (for various values of K,the "’J,-conserved’" quantity) for reactants with the set of solutions forthe products. Both procedures appear to work, but greatly increase thecomplexity of the Jz-conserving approximation over that required forinelastic collisions. Nevertheless, this approximation greatly reduces thenumber of coupled equations to be solved, and permits the quantumcalculation of reaction probabilities to be extended to heavier systemsfor which the number of open channels involved would prohibit exactclose-coupling calculations at the present time [e.g. F+ H2; see Refer-ence (161)].

The infinite order sudden (IOS) approximation, also reviewed Kouri (164) for inelastic collisions, has recently been developed (169-171) for and applied (171-173) to reactive scattering. In the simplestterms (174), the lOS approximation for inelastic scattering consists replacing the eigenvalues of the rotational angular momentum operator(in the CS equations) by a single effective energy; by utilizing closure the rotational functions, we obtain a set of coupled equations forvibrational states containing the internal (A-BC) angle as a parameter.Thus the rotational manifold of states disappears completely from thecoupled equations, being replaced by the necessity to solve this verymuch easier problem at a set of angles. The angle-dependent T-matrixelements are then projected onto the asymptotic rotor states to de-termine transition probabilities. This approximation does not conserveenergy, even for inelastic collisions.

The difficulties of extending the IOS to reactive scattering are evengreater than for the J~-conserving approximation alone, because theapproximation is more extreme in each chemical channel and, althoughthe wavefunction matching on a channel-matching surface appearsstraightforward, the momentum (or derivative) matching appears moredifficult to justify (169-171, 175). Barg & Drolshagen (170) confinetheir discussion to a two channel light-heavy-light atom (L-H-L)exchange reaction where they find that a one-to-one angle matching onthe surface produces complete orbital-to-rotational angular momentumconversion (and vice versa) on the matching surface. Their preliminaryresults (176) on D + HC1 "show agreement with quasi-calculations...asgood as a factor of one to three for the reactive integral rotational crosssections."

The work of Bowman & Lee (171, 172) appears to be more general,allowing for asymmetric three-channel reactions. They, however, as-sume that l~=la and J~=Jl~ on the matching surface (although the

www.annualreviews.org/aronlineAnnual Reviews

Page 22: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

422 WALKER & LIGHT

opposite result of Barg & Drolshagen could be used as well), thus alsofinding a one-to-one angular matching procedure. In applications toH + H2 they find (171) that their low energy cross sections overestimatethe true cross sections by as much as a factor of ten, but better(although not quantitative) agreement is found at higher energies (E = to .7 eV). As might be expected in a theory for which energy is notconserved and rotational transitions occur well into closed channels, theproduct rotational distributions are in poor qualitative agreement withexact results.

The matching equations derived by Khare et al (169) are moregeneral than those above, but similar simplifications are suggested [cfEqs. 81 and 91-102 of Reference (169)]. As in Bowman & Lee’s work,Baer et al (173) find the IOS cross section is too large at threshold, butimproves somewhat with increasing energy.

In summary, the search for an adequate lOS description of reactivecollisions has been very active, and with good reason. The IOS ap-proximation permits one to replace a very large intractable set ofequations by numerous very small sets (involving vibrational statesonly) for describing quantum reactive scattering. The lOS approxima-tions, however, are quite strong and it remains to be seen whether thisapproach will produce scattering results more accurate than classicaltrajectory results. The preliminary results to date offer some hope, butalso indicate that rather extensive tests (and probably modifications)will be required before the lOS approach can be used with confidence.Other inelastic scattering decoupling approximations (e.g. /-dominant)have not yet been developed for reactive scattering.

OTHER QUANTUM APPROXIMATIONS Several recent models of reactivesystems have been proposed in which a portion of the dynamics is donequantum mechanically, the remainder being treated classically, statisti-cally, or by a crude quantum model. We shall discuss these briefly.

Fischer et al (177) proposed a "statistical dynamic" model of ex-change reactions in which the vibrational-translational degrees of free-dom are treated in the FC manner (see above), but with a statisticaltreatment of rotational degrees of freedom. By assuming the factoriza-tion of T-matrix elements into independent elements for the differentdegrees of freedom and utilizing the concept of an internal temperaturefor the "statistical" degrees of freedom, they obtain approximate ana-lytic results for product energy distributions. Comparison of thesedistributions with experiment for H(D) + F2(C12) and C1 + H(D)I quite encouraging agreement (178). Correlations in the distributions for

www.annualreviews.org/aronlineAnnual Reviews

Page 23: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 423

the various degrees of freedom are suppressed and absolute crosssections presumably will not be obtained accurately.

A related approach by Baer (179) attempts to take distributionsknown from exact quantum calculations on reduced (1D or 2D) modelsand, using statistical arguments, generate full quantum distributions.Relatively good agreement with classical trajectory results on H + Br2was obtained from 1D quantum vibrational distributions using thismethod.

Basilevsky & gyaboy, after deriving a collinear three atom reactivescattering method (117), make quasiclassical approximations on thecollinear dynamics (180, 181) and finally obtain a very crude analyticalsolution, which, however, may give some qualitative insight into thecauses of vibrational population inversions.

Finally, Wyatt &.~ Walker (182) and Wyatt (183) used full quantumclose-coupling methods over a portion of the lEVI- and 3Y~u+ surfaces ofthe F+H2 system to study quenching of the F(2PI/2) in the reactivesystem. Because the ground state reaction is very exothermic and cannotbe handled exactly by current close-coupling methods, he included theeffects of reaction by applying the detailed quantum transition statetheory (DQTST) (184, 185) at a surface just beyond the

OTHER TOPICS

Transition State and Statistical TheoriesAlthough the transition state theory (TST) of chemical reactions hasbeen in textbooks for many years, it has been examined, reformulated,and tested extensively in the past few years. Pechukas (24) has given excellent (and entertaining) review of statistical approximations (prim-arily TST) including work up to 1976. Chesnavich & Bowers (186) 1979 review both TST and other statistical theories and their applicationto ion-molecule reactions. Garrett & Truhlar (23, 187) give very exten-sive references.

It has been known for many years that classical transition statetheory, in either the canonical or microcanonical form, provides anupper bound on classical reaction rates. Pechukas & Pollak (188-190)have, in the past few years, defined the mathematical conditions re-quired for this bound to be exact. Building on previous work (188, 189)they (190) recently showed that for a collinear atom-diatom system,microcanonical TST would be exact for all systems and energies forwhich a unique transition state exists. (In their definition, the transition

www.annualreviews.org/aronlineAnnual Reviews

Page 24: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

424 WALKER ~ LIGHT

state is defined by a periodic vibration in the interaction region. Morethan one such periodic trajectory in the interaction region may define acollision complex.) Although extension of these rigorous classical resultsto 3D and the examination of real potentials to determine their applica-bility may be difficult, at least the mathematical framework for thisexamination has been constructed.

Garrett & Truhlar (187, 191) considered the relationship of varia-tional microcanonical TST (which should yield the least upper bound tothe rate constant at fixed E) with other more approximate (but moreeasily applied) formulations such as the minimum density of statescriteria (192, 193) and Miller’s unified statistical theory (194). recently, they applied their generalized theory (canonical variationalTST) to a number of 3D reactions, showing it superior to conventionalTST (195, 196). In comparing the results of all these theories with theexact classical canonical rate constants determined from classical trajec-tory calculations for nine collinear triatomic systems, Garrett & Truhlar(187, 191) show that the microcanonical variational and unified statisti-cal theories are superior to conventional canonical and canonical varia-tional theories, although the differences are not great. On the otherhand, quantized versions of these approaches show that the generalizedvariational TST can give significant improvements over conventionalTST (197).

Truhlar has also considered the effects of the classical nature of mostTST calculations by comparing them with quantum calculations (23).The accuracy of quantum tunneling corrections was investigated byGarrett & Truhlar for the several collinear systems for which accuratequantum calculations are available (198). Most recently they applied thegeneralized TST theories with quantum corrections to the H + D2 andD + H2 reactions using the accurate a priori surface of Liu & Siegbahn(33, 34) and found excellent agreement for both rate constants andkinetic isotope effects (199) with the best available experimental data.From this last work it appears that TST, when carefully applied, shouldproduce rate constants of quite useful accuracy (within a factor of 2) least for reactions with a single activation energy barrier and withreasonable skewing angles. It should, however, be tested further onqualitatively different potential energy surfaces and at higher energies.

We shall not detail all the work using TST, but only mention thesignificant new work for ion-molecule reactions presented by Chesna-vich & Bowers (186), and the construction of a TST for three-bodyion-molecule recombination reactions considered by Herbst (200). Fi-nally, Chesnavich et al (201) also used microcanonical variational TST

www.annualreviews.org/aronlineAnnual Reviews

Page 25: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 425

and classical trajectories to obtain an upper bound for capture crosssections in ion-dipole collisions (long a problem because of the long-range anisotropic potential). They find good agreement between thebound, trajectory, and experimental results.

There has been some further development of statistical theories inwhich a strong coupling collision complex is formed. In addition to therelevant work mentioned above (23, 196, 197), Brumer et al (46, 202)have investigated, via classical trajectory studies, the dependence ofcomplex formation on geometrical and mass factors, as well as energyand potential energy characteristics. They characterize the subset oftrajectories that form complexes by passage through a region of thepotential surface in which substantial exponential divergence takesplace. In comparing the product distributions of this subset with thatobtained from the Wagner & Parks formulation (203) of statisticaltheories in terms of a strong coupling region, they find that the criterionof substantial exponential divergence indeed produces a statistical dis-tribution in qualitative agreement with phase space theory. A moredetailed analysis of the subset of trajectories showed clearly the statisti-cal nature of the distributions independent of the specific assumptionsof phase space theory.

Pollak & Levine (204) have recently shown how the definition of strong coupling region and assumption of maximal entropy of the finaldistributions can lead to a variety of statistical theories depending onconstraints imposed on the distributions. They show that Miller’s uni-fied theory, TST, and phase space theory can be related by constraintson flux across appropriate boundary surfaces. In application to collinearH + H2, they find that addition of a second constraint [average numberof (re)crossings of a surface] produces excellent agreement with theprobability of reaction determined by trajectory calculations. However,trajectory calculations are required to determine the constraint, which isthen used in the statistical theory.

Pollak & Pechukas (205) also re-derived Miller’s unified statisticaltheory, applied it to collinear H+ HE, and derived a lower bound aswell as an upper bound on the reaction probability. It proved to beremarkably accurate.

We have considered information theory approaches to reaction dy-namics beyond the scope of this review. We take this position notbecause the work is not relevant--it is--but because it has beenreviewed with some frequency and space limitations preclude its inclu-sion here. The interested reader should see reviews by Levine andco-workers (11, 12).

www.annualreviews.org/aronlineAnnual Reviews

Page 26: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

426 WALKER & LIGHT

Collision-Induced Dissociation

Whenever the relative A+ BC collision energy exceeds the BC bondenergy, we open up the product reaction channel A+B+C. Suchprocesses are called collision-induced dissociation (CID) and have beeninvestigated using (extensions of) methods developed for rearrangementcollisions. Both classical and quantum methods have been used to studyCID; for detailed recent reviews, see the discussions of Kuntz (25) andDiestler (26).

Most recent work on CID uses the quasiclassical trajectory approachand focuses attention on the dependence of the dissociation rate (orcross section) on the initial internal molecular energy (206-211). Molec-ular rotational energy plays an important part in the dissociationdynamics (206, 208, 209, 211), suggesting that a purely vibrationalladder-climbing model oversimplifies the actual process. The ladder-climbing model allows only a single vibrational quantum transition percollision, with dissociation occurring only from the highest boundvibrational state. The actual process is better characterized by a com-bined vibration-rotation ladder climbing (206) in which the dissociationstep occurs from molecules whose total energy is near the dissociationlimit, but with a significant spread of internal energy between vibrationand rotation (208, 209, 211). Blais & Truhlar (209) and Lehr & Birks(211) present excellent discussions of the importance of rotations CID and should be consulted for references to earlier literature.

Semielassieal (212, 213) and quantum mechanical (214-218) calcula-tions of CID have been reported for collinear ABC systems. Semiclassi-cal and time independent quantum methods (212-216) require a discre-tization of the continuum molecular states; dissociation is then definedas a sum over transitions to the continuum. Time-dependent wave-packet approaches (217, 218) avoid this difficulty; the dissociationprobability is determined by projecting out the bound state componentsof the wavefunction at large times. These calculations demonstrate thatdissociation occurs from vibrational states well below the continuumand that the ladder-climbing model oversimplifies the dynamics evenwithout rotations.

Considering the importance of rotations demonstrated by the 3Dclassical calculations discussed previously, it is doubtful that collinearcalculations will be quantitatively reliable for CID. Extending quantumCID methods to higher dimensions (to include rotations) will cause substantial increase in complexity. In the near future, detailed CIDdynamics will be best treated by quasiclassical techniques. These calcu-lations will be accurate to the extent that good potential energy surfaces

www.annualreviews.org/aronlineAnnual Reviews

Page 27: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 427

are available and to the extent that quantum effects (presumably nearthe threshold for dissociation) are minimal. Studying quantum effects inthree dimensional CID will require substantial future developments inthe methodology of the problem.

Laser-Collision Reactive Processes

Although lasers have been used for many years to modify the initialstates of reactants, to analyze the products of reactions, and to cause(collisionless) photodissociation, it is only in the last five years or so thatthe effects of strong laser fields on collision dynamics have beenstudied. Because the field of multiphoton dissociation has been thesubject of recent conference proceedings (219) and because lasers usedfor reactant preparation and product detection are primarily experimen-tal tools for studying state-to-state chemical reactions, we shall com-ment here only on recent developments in laser-collision dynamics,particularly reaction dynamics. To date the preponderance of work inthis area (meager though it is) has been theoretical and the onlylaser-collision-induced reactive processes observed have been laser-induced associative ionization (220) (A +B+hv-->AB+ +e-) and, veryrecently, a laser-induced exchange reaction (221) (K + HgBr2 +hv-->KBr+ HgBr*).

The former process was studied formally (qualitatively) by Bellum George (222), using the discretization techniques for the continuumpreviously developed for field free associative ionization (223) and the"electronic-field" representation of the scattering process in the laserfield, also advocated by George et al (224). In this representation, thezero order-time independent states are taken to be those determined bythe fixed nuclei electrons plus field Hamiltonian, and scattering (with orwithout non-Born-Oppenheimer coupling) occurs on these surfaces.This leads to coupled-time independent equations for the scattering.Using this representation George and co-workers have also investigatedlaser-induced unimolecular dissociation (225) and electronically in-elastic scattering of F on H2 (226). Yuan & George (227) investigatedthe reactive scattering of F on H2 in the presence of a laser field usingsemiclassical methods.

Light & Altenberger-Siczek (228) also used this framework to con-sider the effects of near resonant laser radiation on atom-diatom ex-change reactions in collinear, two electronic surface models. By assum-ing near resonant fields and electronically allowed transition couplings,they showed that significant alterations in transition probabilities toexcited states could be obtained, although "true" laser catalysis would

www.annualreviews.org/aronlineAnnual Reviews

Page 28: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

428 WALKER & LIGHT

be unlikely. The recent experiments of Hering et al (221) appear confirm that such processes can occur at modest (< 107W/cm2) laserpowers, but the cross sections appear to be small (~10 -17 cm2).

An alternative approach to these processes has been taken by Cope-land (229) and by Orel & Miller (230, 231). Copeland derives, but not apply, semiclassical dynamical equations in a time-dependent repre-sentation not utilizing the "electronic-field" states. Orel & Miller treateverything classically, with the laser field modifying the classical dy-namics of the motion of the nuclei on a single electronic energy surfacedue to a time dependent perturbation (the laser field). They show thatthe field will affect the dynamics even for infrared inactive species dueto the lack of symmetry of the motion along the reaction coordinate andthat the effective threshold for reaction will be lowered. In a qualitativeanalytic model, however, they also show that laser powers on the orderof 101°- 1012W/cm2 would be required to produce significant effects.

Overall, the research in this area has produced mixed results. Processesare reasonably well understood theoretically--the framework of bothtime-independent multiple surface approaches and time-dependentsingle surface approaches has been established. However, the paucity ofexperimental data confirms the general trend of theoretical predictionthat to make laser-collision-induced processes occur with reasonableprobability, very strong (~> 109W/cm2) fields are required. In addition,the difficulty of obtaining experimental data suggests that the utility ofthese processes as probes of potential energy surfaces of reactioncomplexes is severely limited and that the prognosis for such processesas general synthetic tools is poor.

ACKNOWLEDGMENTS

This work was performed in part under the auspices of the US Depart-ment of Energy. Partial support by the National Science Foundationunder NSF grant number CHE79-06896 is gratefully acknowledged.

Literature Cited

1. Levine, R. D., Bernstein, R. B. 1974.Molecular Reaction Dynamics. NewYork: Oxford Univ. Press, 250 pp.

2. Child, M. S. 1974. Molecular Colli-sion Theory. London & New York:Academic. 300 pp.

3. Bernstein, R. B., ed. 1979. Atom-Molecule Collision Theory, A Guidefor the Experimentalist. New York &London: Plenum. 779 pp.

4. Miller, W. H., ed. 1976. Dynamics ofMolecular Collisions, Pt. A. NewYork & London: Plenum. 318 pp.

5. Miller, W. H., ed. 1976. Dynamics ofMolecular Collisions, Pt. B. NewYork & London: Plenum. 380 pp.

6. Bernstein, R. B. 1979. See Ref. 3, pp.1-43

7. Levy, M. R. 1979. Prog. React. Kinet.10:1-252

8. Schaefer, H. F. 1II. 1979. See Ref. 3,pp. 45-78

9. Kuntz, P. J. 1979. See Ref. 3, pp.79-110

10. Kuntz, P. J. 1976. See Ref. 5, pp.53-120

www.annualreviews.org/aronlineAnnual Reviews

Page 29: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 429

11. Levine, R. D., Kinsey, J. L. 1979.See Ref. 3, pp. 693-750

12. Levine, R. D. 1978 Ann. Re~. Phys.Chem. 29:59-92

13. Pattengill, M. D. 1979. See Ref. 3,pp. 359-75

14. Truhlar, D. G., Muckerman, J. T.1979. See Ref. 3, pp. 505-66

15. Porter, R. N., Raft, L. M. 1976. SeeRef. 5, pp. 1-52

16. Secrest, D. 1979. See Ref. 3, pp.265-99

17. Light, J. C. 1979. See Ref. 3, pp.467-76

18. Wyatt, R. E. 1979. See Ref. 3, pp.567-94

19. Wyatt, R. E. 1979. See Ref. 3, pp.477-503

20. Connor, J. N. L. 1979. Comput. Phys.Commun. 17:117-44

21. Truhlar, D. G., Wyatt, R. E. 1976.Ann. Re~. Phys. Chem. 27:1-43

22. Micha, D. A. 1975. Adt~. Chem. Phys.30:7-75

23. Truhlar, D. G. 1979. J. Phys. Chem.83:188-99

24. Pechukas, P. 1976. See Ref. 5, pp.269-322

25. Kuntz, P. J. 1979. See Ref. 3, pp.669-92

26. Diestler, D. J. 1979. See Ref. 3, pp.655-67

27. Porter, R. N. 1974. Ann. Re~. Phys.Chem. 25:317-55

28. Karplus, M., Porter, R. N., Sharma,R. D. 1965. J. Chem. Phys. 43:3259-87

29. Osherov, V. I., Ushakov, V. G.,Lomakin, L. A. 1978. Chem. Phys.Lett. 55:513-14

30. Mayne, H. R., Toennies, J. P. 1979.J. Chem. Phys. 70:5314-15

31. Mayne, H. R. 1979. Chem. Phys. Lett.66:487-92

32. Truhlar, D. G., Horowitz, C. J. 1978.J. Chem. Phys. 68:2466-76

33. Liu, B. 1973. J. Chem. Phys. 58:1925-37

34. Siegbahn, P., Lith B. 1978. J. Chem.Phys. 68:2457-65

35. Zuhrt, C., Schneider, F., Havemann,U., Ziilicke, L., Herman, Z. 1979.Chem. Phys. 38:205-l0

36. Gittins, M. A., Hirst, D. M. 1979.Chem. Phys. Lett. 65:507 10

37. Schinke, R., Lester, W. A. Jr. 1979.J. Chem. Phys. 70:4893-4902

38. Luntz, A. C., Schink¢, R., Lester, W.A. Jr., Giinthard, H, H. 1979. J.Chem. Phys. 70:5908-9

39. Herbst, E., Payne, L. G., Champion,R. L., Doverspik¢, L. D. 1979. Chem.Phys. 42:413-21

40. Preston, R. K., Cohen, J. S. 1976. J.Chem. Phys. 65:1589-90

41. Jaffe, R. L. 1979. Chem. Phys.40:185-206

42. Persky, A. 1979. J. Chem. Phys.70:3910-19

43. Polanyi, J. C., Schreiber, J. L. 1978.Chem. Phys. 31:113-36

44. Polanyi, J. C., Sathyamurthy, N.1978. Chem. Phys. 33:287-303

45. Polanyi, J. C., Sathyamurthy, N.1979. Chem. Phys. 37:259-64

46. Duff, J. W., Brumer, P. 1979. J.Chem. Phys. 71:2693-702

47. Brown, J. C., Bass, H. E., Thomp-son, D. L. 1979. J. Chem. Phys.70:2326-34

48. Hase, W. L., Wolf, R. J., Sloane, C.S. 1979. J. Chem. Phys. 71:2911-28

49. Grant, E. R., Bunker, D. L. 1978. J.Chem. Phys. 68:628-36

50. Gray, J. C., Truhlar, D. G., Clemens,L., Duff, J. W., Chapman, F. M. Jr.,Morrell, G. O., Hayes, E. F. 1978. J.Chem. Phys. 69:240-52

51. Gray, J. C., Garrett, B. C., Truhlar,D. G. 1979. J. Chem. Phys. 70:5921-22

52. Truhlar, D. G. 1979. J. Phys. Chem.83:188-99

53. Gray, J. C., Truhlar, D. G., Baer, M.1979. J. Phys. Chem. 83:i045-52

54. Connor, J. N. L., Jakubetz, W.,Lagan~, A. 1979. J. Phys. Chem.83:73-78

55. Connor, J. N. L., Lagan~, A. 1979.Comput. Phys. Commun. 17:145-48

56. Connor, J. N. L., Lagan~, A. 1979.Mol. Phys. 38:657-67

57. Hutchinson, J. S., Wyatt, R. E. 1979.J. Chem. Phys. 70:3509-23

58. Duff, J. W., Brumer, P. 1979. J.Chem. Phys. 71:3895-96

59. Deleted in proof60. Faist, M. B., Muckerman, J. T.,

Schubert, F. E. 1978. J. Chem. Phys.69:4087-96

61. Muekerman, J. T., Faist, M. B. 1979.J. Phys. Chem. 83:79-87

62. Suzukawa, H. H. Jr., Wolfsberg, M.1978. J. Chem. Phys. 68:1423-25

63. Billing, G. D., Poulsen, L. L. 1979.Chem. Phys. Lett. 66:177-28

64. Brumer, P. 1973. J. Comput. Phys.14:391-419

65. Bruraer, P., Dtfff, J. W. 1976. J.Chem. Phys. 65:3566-74

www.annualreviews.org/aronlineAnnual Reviews

Page 30: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

430 WALKER & LIGHT

66. Duff, J. W., Brumer, P. 1977. J.Chem. Phys. 67:4898-4911

67. Ford, J. 1973. Adv. Chem. Phys.24:155-85

68. Percival, I. C. 1977.Adv. Chem. Phys.36:1-61

69. Schatz, G. C. 1979. Chem. Phys. Lett.67:248-51

70. Stratt, R. M., Handy, N. C., Miller,W. H. 1979. J. Chem. Phys. 71:3311-22

71. Cerjan, C., Reinhardt, W. P. 1979. J.Chem. Phys. 71:1819-31

72. Heller, E. J. 1980. J. Chem. Phys.72:1337-47

73. Miller, W. H. 1974. Adv. Chem. Phys.30:77-136

74. Child, M. S. 1974. See Ref. 2, pp.180-209

75. Child, M. S. 1976. See Ref. 5, pp.171-215

76. Gordon, R. G., McGinnis, R. P.1971. J. Chem. Phys. 55:~898-906

77. Brumer, P. 1974. Chem. Phys. Lett.28:345-51

78. Blais, N. C., Truhlar, D. G. 1976. J.Chem. Phys. 65:5335--56

79. Truhlar, D. G. 1976. Int. J. QuantumChem. @rap. 10:239-50

80. Raff, L. M. 1974. J. Chem. Phys.60:2220-44

81. Louck, J. D., Galbraith, H. W. 1976.Rev. Mod. Phys. 48:69-106

82. Biedenharn, L. C., Louek, J. D. 1980.Angular Momentum in QuantumPhysics, Chap. 7, Sect. 11. Reading,Mass.: Addison-Wesley. In press

¯ 83. Suzukawa, H. H. Jr., Wolfsberg, M.,Thompson, D. L. 1978. J. Chem.Phys. 68:455-72

84. Sehatz, G. C., Moser, M. D. 1978.Chem. Phys. 35:239-51

85. Sehatz, G. C., Moser, M. D. 1978. J.Chem. Phys. 68:1992-93

86. Schatz, G. C., Mulloney, T. 1979. J.Phys. Chem. 83:989-99

87. Schatz, G. C., Mulloney, T. 1979. J.Chem. Phys. 71:5257-67

88. Eastes, W., Marcus, R. A. 1974. J.Chem. Phys. 61:4301-6

89. Noid, D. W., Marcus, R. A. 1975. J.Chem. Phys. 62:2119-24

90. Noid, D. W., Marcus, R. A. 1977. J.Chem. Phys. 67:559-67

91. Noid, D, W., Koszykowski, M. L.,Marcus, R. A. 1979. J. Chem. Phys.71:2864-73

92. Chapman, S., Garrett, B. C., Miller,W. H. 1976. J. Chem. Phys. 64:502-9

93. Handy, N. C., Colwell, S. M., Miller,W. H. 1977. Faraday Discuss. Chem.Soc. 62:29-39

94. Miller, W. H. 1977. Faraday Discuss.Chem. Soc. 62:40-46

95. Sorbie, K. S. 1976. MoL Phys.32:1577-90

96. Sorbic, K. S., Handy, N. C. 1976.MoL Phys. 32:1322-47

97. Colwell, S. M., Handy, N. C. 1978.Mol. Phys. 35:1183-90

98. Jaffe, C., Rehahardt, W. P. 1979. J.Chem. Phys. 71:1862-69

99. Wyatt, R. E. 1977. In State to StateChemistry, ACS Syrup. Ser. No. 56,exl. P. R. Brooks, E. F. Hayes, pp.185-205. Washington DC: Am.Chem. Soc. 259 pp.

100, McNutt, J. F., Wyatt, R. E. 1979. J.Chem. Phys. 70:5307-12

101. Witriol, N. M., Stettler, J. D., Rather,M. A., Sabin, J. R., Trickey, S. B.1977. J. Chem. Phys. 66:1141-59

102. Garrett, B. C., Miller, W. H. 1978. J.Chem. Phys. 68:4051-55

103. Adams, J. E., Miller, W. H. 1979. J.Phys. Chem. 83:1505-7

104. Rabitz, H., Askar, A., Cakmak, A. S.1978. Chem. Phys. 29:61-65

105. Askar, A., Cakmak, A. S., Rabitz, H.1978. Chem. Phys. 33:267-86

106. Baker, A. J., Soliman, M. O. 1979. J.Comput. Phys. 32:289-324

107. Askar, A., Cakmak, A. S., Rabitz, H.1980. J. Chem. Phys. 72:5287-89

108. Light, J. C., Walker, R. B. 1976. J.Chem. Phys. 65:4272-82

109. Stechel, E. B., Walker, R. B., Light,J. C. 1978. J. Chem. Phys. 69:3518-31

110. Light, J. C., Walker, R. B., Steehel,E. B., Schmalz, T. G. 1979. Comput.Phys. Commun. 17:89-97

I 11. Walker, R. B. 1978. Quantum Chem.Program Exch., Program No. 352.Dep. Chem., Indiana Univ.,Bloomington

112. Gordon, R. G. 1969. J. Chem. Phys.51:14-25

113. Adams, J. T., Smith, R. L., Hayes, E.F. 1974. J. Chem. Phys. 61:2193-99

114. Manz, J. 1974. Mol. Phys. 28:399-413

115. Manz, J. 1975. Mol. Phys. 30:899-910

116. Connor, J. N. L., Jakubetz, W.,Manz, J. 1975. Mol. Phys. 29:347-55

117. Basilevsky, M. V., Ryaboy, V. M.1979. Chem. Phys. 41:461-76

www.annualreviews.org/aronlineAnnual Reviews

Page 31: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 431

118. Light, J. C., Walker, R. B. 1976. J.Chem. Phys. 65:1598-99

119. Johnson, B. R. 1978. J. Chem. Phys.69:4678-88

120. Johnson, B. R. 1973. J. Comput. Phys.13:445-49

121. Johnson, B. R. 1979. Proc. WorkshopAlgorithms Comput. Codes At. andMol. Quantum Scattering Theory,1:86-104. Lawrence Berkeley Lab.Rep. No. LBL-9501(1979). 402 pp.

122. Thomas, L. D. 1979. J. Chem. Phys.70:2979-85

123. Stechel, E. B., Schmalz, T. G., Light,J. C. 1979. J. Chem. Phys. 70:5640-59

124. Schmalz, T. G., Stechel, E. B., Light,J. C. 1979. J. Chem. Phys. 70:5660-71

125. Leasure, S., Bowman, J. M. 1978. J.Chem. Phys. 68:2825-30

126. Deleted in proof127. Schatz, G. C., Kuppermann, A. 1976.

J. Chem. Phys. 65:4642-67128. Walker, R. B., Light, J. C., Alten-

berger-Siczek, A. 1976. J. Chem.Phys. 64:1166-81

129. Steehel, E. B., Walker, R. B., Light,J. C. 1977. Rep. Workshop React.Collisions, ed. C. E. Moser, pp. 1-49.Orsay France: CECAM 342 pp.

130. Schatz, G. C., Kuppermann, A. 1975.J. Chem. Phys. 62:2502-3

131. Elkowitz, A. B., Wyatt, R. E. 1975.J. Chem. Phys. 62:2504-5

132. Elkowitz, A. B., Wyatt, R. E. 1975.J. Chem. Phys. 63:702-21

133. Schatz, G. C., Kuppermann, A. 1976.J. Chem. Phys. 65:4668-92

134. Walker, R. B., Steehel, E. B., Light,J. C. 1978. J. Chem. Phys. 69:2922-23

135. Redmon, M. J., Wyatt, R. E. 1979.Chem. Phys. Lett. 63:209-12

136. Wyatt, R. E. 1980. Proc. 3rd Conf.Quantum Chem., Kyoto, Japan.Dordreeht, Neth.: Reidel. In press

137. Sparks, R. K., Hayden, C. C.,Shobatake, K., Neumark, D. M., Lee,Y. T. 1980. See Ref. 136. In press

138. Redmon, M. J. 1979. Int. J. QuantumChem. Syrup. 13:559-68

139. Porter, R. N., Karplus, M. 1964. J.Chem. Phys. 40:1105-15

140. Clary, D. C., Nesbet, R. K. 1978.Chem. Phys. Lett. 59:437-42

141. Clary, D. C., Nesbet, R. K. 1979. J.Chem. Phys. 71:1101-9

142. Choi, B. H., Tang, K. T. 1974. J.

Chem. Phys. 61:5147-57143. Choi, B. H., Tang, K. T. 1976. J.

Chem. Phys. 65:5161-80144. Tang, K. T., Choi, B. H. 1975. J.

Chem. Phys. 62:3642-58145. Shan, Y., Choi, B. H., Poe, R. T.,

Tang, K. T. 1978. Chem. Phys. Lett.57:379- 84

146. Schafer, T. P., Siska, P. E., Parson, J.M., Tully, F. P., Wong, Y. C., Lee,Y. T. 1970. J. Chem. Phys. 53:3385-87

147. Clary, D. C., Connor, J. N. L. 1979.Chem. Phys. Lett. 66:493-97

148. Brandt, D., Polanyl, J. C. 1978.Chem. Phys. 35:23-34

149. Halavee, U., Shapiro, M. 1976. J.Chem. Phys. 64:2826-39

150. Schatz, G. C., Ross, J. 1977. J. Chem.Phys. 66:1021-36

151. Schatz, G. C., Ross, J. 1977. J. Chem.Phys. 66:1037 53

152. Schatz, G. C., Ross, J. 1977. J. Chem.Phys. 66:2943-58

153. Vila, C. L., Kinsey, J. L., Ross, J.,Schatz, G. C. 1979. J. Chem. Phys.70:2414-24

154. Fung, K. H., Freed, K. F. 1978.Chem. Phys. 30:249-67

155. Mukamel, S., Ross, J. 1977. J. Chem.Phys. 66:3759-66

156. Vila, C. L., Zvijac, D. J., Ross, J.1979. J. Chem. Phys. 70:5362-75

157. Wong, J. K. C., Brumer, P. 1979.Chem. Phys. Lett. 68:517-22

158. Baer, M. 1974. J. Chem. Phys.60:1057-63

159. Elkowitz, A. B., Wyatt, R. E. 1976.Mol. Phys. 31:189-201

160. Kuppermann, A., Schatz, G. C.,Dwyer, J. 1977. Chem. Phys. Left.45:71-73

161. Redmon, M. J., Wyatt, R. E. 1977.Int. J. Quantum Chem. Symp. 11:343-51

162. McGuire, P., Kouri, D. J. 1974. J.Chem. Phys. 60:2488-99

163. Pack, R. T 1974. J. Chem. Phys.60:633-39

164. Kouri, D. J. 1979. See Ref. 3, pp.301-58

165. Coomb¢, D. A., Snider, R. F. 1979.J. Chem. Phys. 71:4284-96

166. Parker, G. A., Pack, R. T 1977. J.Chem. Phys. 66:2850-53

167. Parker, G. A., Pack, R. T 1978. J.Chem. Phys. 68:1585-1601

168. Khare, V., Kouri, D. J., Pack, R. T1978. J. Chem. Phys. 69:4419-30

www.annualreviews.org/aronlineAnnual Reviews

Page 32: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

432 WALteR & LIOHT

169. Khare, V., Kouri, D. J., Baer, M.1979. J. Chem. Phys. 71:1188-1205

170. Baxg, G., Drolshagen, G. 1980.Chem. Phys. 47:209-34

171. Bowman, J. M., Lee, K. T. 1980. J.Chem. Phys. 72:5071-88

172. Bowman, J. M., Lee, K. T. 1978. J.Chem. Phys. 68:3940-41

173. Baer, M., Khare, V., Kouri, D. J.1979. Chem. Phys. Left. 68:378-81

174. Secrest, D. 1975. J. Chem. Phys.62:710-19

175. Khaxe, V., Kouri, D. J. 1978. J.Chem. Phys. 69:4916-20

176. Drolshagen, G. 1979. PhD thesis.Bericht, Max-Planck Institut f~rStr6mungsforschung, G~ttingen,Fed. Repub. Germany

177. Fischer, S., Venzl, G., Robin, J.,Ratner, M. A. 1978. Chem. Phys.27:251-61

178. Venzl, G., Fischer, S. 1978. Chem.Phys. 33:305-12

179. Baer, M. 1978. Chem. Phys. Lett.57:316-20

180. Basilevsky, M. V., Ryaboy, V. M.1979. Chem. Phys. 41:477-88

181. Basilevsky, M. V., Ryaboy, V. M.1979. Chem. Phys. 41:489-94

182. Wyatt, R. E., Walker, R. B. 1979 J.Chem. Phys. 70:1501-10

183. Wyatt, R. E. 1979. Chem. Phys. Lett.63:503-8

184. Light, J. C., Altenberger-Siczek, A.1975. Chem. Phys. Left. 30:195-99

185. Light, J. C., Altenberger-Siczek, A.1976. J. Chem. Phys. 64:1907-13

186. Chesnavich, W. J., Bowers, M. 1979.Ga~ Pha~e lon Chemistry, ed. M.Bowers, i:119-51. New York:Academic

187. Garrett, B. C., Truhlar, D. G. 1979.J. Phys. Chem. 83:1052-79; erratum,p. 3058

188. Pollak, E., Pechukas, P. 1978. J.Chem. Phys. 69:1218-26

189. Pechukas, P., Pollak, E. 1977. J.Chem. Phys. 67:5976-77

190. Pechukas, P., Pollak, E. 1979. J.Chem. Phys. 71:2062-67

191. Garrett, B. C., Truhlar, D. (3. 1979.J. Chem. Phys. 70:1593-98

192. Bunker, D. L., Pattengill, M. 1968. J.Chem. Phys. 48:772-76

193. Wong, W. H., Marcus, R. A. 1971. J.Chem. Phys. 55:5625-29

194. Miller, W. H. 1976. J. Chem. Phys.65:2216-23

195. Garfett, B. C., Truhlar, D. G. 1979.J. Am. Chem. Soc. 101:4534-48

196. Garrett, B. C., Truhlar, D. G. 1979.J. Am. Chem. Soc. 101:5207-17

197. Garrett, B. C., Truhlar, D. G. 1979.J. Phys. Chem. 83:1079-1112

198. Garrett, B. C., Truhlar, D. G. 1979.J. Phys. Chem. 83:200-3; erratum,p. 3058

199. Garrett, B. C., Truhlar, D. G. 1980.J. Chem. Phys. 72:3460-71

200. Herbst, E. 1979. J. Chem. Phys.70:2201-4

201. Chesnavich, W. J., Su, T., Bowers,M. 1980. J. Chem. Phys. 72:2641-55

202. Fitz, D. E., Brumer, P. 1979. J. Chem.Phys. 70:5527-33

203. Wagner, A., Parks, E. 1976. J. Chem.Phys. 65:4343-61

204. Pollak, E., Levine, R. D. 1980. J.Chem. Phys. 72:2990-97

205. Pollak, E., Pechukas, P. 1979. J.Chem. Phys. 70:325-33

206. Dove, J. E., Raynor, S. 1978. Chem.Phys. 28:113-24

207. Havemann, U., Pacak, V., Herman,Z., Schneider, F., Zuhrt, C., Z/~licke,L. 1978. Chem. Phys. 28:147-54

208. Stace, A. J. 1978. Chem. Phys. Lett.55:77-79

209. Blais, N. C., Truhlar, D. G. 1979. J.Chem. Phys. 70:2962-78

210. Kuntz, P. J., Kendrick, J., Whitton,W. N. 1979. Chem. Phys. 38:147-60

211. Lehr, T., Birks, J. W. 1979. J. Chem.Phys. 70:4843-48

212. Rusinek, I., Roberts, R. E. 1976. J.Chem. Phys. 65:872-80

213. Rusinek, I., Roberts, R. E. 1978. J.Chem. Phys. 68:!147-52

214. Wolken, G. Jr. 1975. J. Chem. Phys.63:528-33

215. Knapp, E.-W., Diesfler, D. J., Lin,Y.-W. 1977. Chem. Phys. Lett.49:378-83

216. Knapp, E.-W., Diestler, D. J. 1977.J. Chem. Phys. 67:4969-75

217. Ford, L. W., Diestler, D. J., Wagner,A. F. 1975. J. Chem. Phys. 63:2019-34

218. Kulander, K. C. 1978.J. Chem. Phys.69:5064-72

219. Eberly, J., Lambropoulus, P., eds.1978. Multiphoton Processes. Proc.Int. Conf. Univ. Rochester, June 6-9,1977. New York: Wiley

220. Hellfeld, A. V., Caddick, J., Weiner,J. 1978. Phys. Rev. Lett. 40:1369-73

www.annualreviews.org/aronlineAnnual Reviews

Page 33: REACTIVE MOLECULAR COLLISIONSlight-group.uchicago.edu/jcl74.pdf · 2003-08-16 · ble from reactive molecular collision theory. These include the vast body of sophisticated and detailed

REACTIVE MOLECULAR COLLISIONS 433

221. Hering, P., Brooks, P., Cud, R. 1980.Phys. Re~. Lett. 44:687-90

222. Bellum, J. C., George, T. F. 1978. J.Chem. Phys. 68:134-44

223. Bellum, J. C., Micha, D. A. 1977.Chem. Phys. 20:121-27

224. George, T. F., Yuan, J.-M., Zimmer-man, I. H., Laing, J. R. 1977. Fara-day Discuss. Chem. Soc. 62:246-54

225. Yuan, J.-M., George, T. F. 1978. J.Chem. Phys. 68:3040-52

226. DeVries, P. L., George, T. F. 1979. J.Chem. Phys. 71:1543-49

227. Yuan, J.-M., George, T. F. 1979. J.Chem. Phys. 70:990-94

228. Light, J. C., Altenberger-Siczek, A,1979. J. Chem. Phys. 70:4108-16

229. Copeland, D. A. 1978. J. Chem. Phys.68:3005-15

230. Orel, A. E., Miller, W. H. 1978.Chem. Phys. Lett. 57:362-63

231. Orel, A. E., Miller, W. H. 1979. J.Chem. Phys. 70:4393-99

www.annualreviews.org/aronlineAnnual Reviews