Download reactive molecular collisions

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Quantum group wikipedia , lookup

Quantum computing wikipedia , lookup

Symmetry in quantum mechanics wikipedia , lookup

Particle in a box wikipedia , lookup

Cross section (physics) wikipedia , lookup

Hydrogen atom wikipedia , lookup

Quantum field theory wikipedia , lookup

Orchestrated objective reduction wikipedia , lookup

Quantum electrodynamics wikipedia , lookup

Quantum teleportation wikipedia , lookup

Wave–particle duality wikipedia , lookup

Path integral formulation wikipedia , lookup

EPR paradox wikipedia , lookup

Topological quantum field theory wikipedia , lookup

Franck–Condon principle wikipedia , lookup

Quantum machine learning wikipedia , lookup

Instanton wikipedia , lookup

Quantum key distribution wikipedia , lookup

Quantum state wikipedia , lookup

Interpretations of quantum mechanics wikipedia , lookup

Renormalization group wikipedia , lookup

Molecular Hamiltonian wikipedia , lookup

Rotational–vibrational spectroscopy wikipedia , lookup

Electron scattering wikipedia , lookup

Rutherford backscattering spectrometry wikipedia , lookup

Scalar field theory wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Renormalization wikipedia , lookup

History of quantum field theory wikipedia , lookup

Canonical quantization wikipedia , lookup

Hidden variable theory wikipedia , lookup

T-symmetry wikipedia , lookup

Transcript
Annual Reviews
www.annualreviews.org/aronline
Ann.Re~.Phys.Chern,1980.31:401-33
REACTIVE MOLECULAR
COLLISIONS
2712
Robert B. Walker
Theoretical Division, University of California,
tory, Los Alamos, New Mexico 87545
Los Alamos Scientific
Labora-
John C. Light
Department of Chemistry and the James Franck Institute,
Chicago, Illinois 60637
University of Chicago,
INTRODUCTION
Thereactive molecular collision is the fundamentalmicroscopicevent.
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 comprehensive 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 experimentalist,
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].
1TheUSGovernment
has the fight to retain a nonexclusive,royalty-free license in and
to any copyrightcoveringthis paper.
401
Annual Reviews
www.annualreviews.org/aronline
402
WALKER & LI(;HT
Our review necessarily does not cover several topics that are inseparable from reactive molecular collision theory. These include the vast
body of sophisticated and detailed experimental data, upon which the
theory must be built. Experiments are discussed in many reviews,
including Bernstein’s AMCT
chapter (6). The most recent experimental
review is given by Levy (7), and includes a truly exhaustive compilation
of data. for every atom-diatom system studied in recent times.
Wealso cannot do justice to describing the efforts of the quantum
chemist to generate the potential energy surfaces upon which all dynamical theories critically
depend. In his AMCT
review, Schaefer (8)
describes considerable progress with ab initio methods. The dynamicist’s
need to evaluate points on potential surfaces often and cheaply prompts
our interest in semiempirical methods, such as the method of diatomicsin-molecules (DIM). A detailed description of the DIM method
presented by Kuntz (9) in an AMCTreview, and the earlier DMCB
review by Kuntz (10) describes how reaction dynamics is affected
specific properties of potential energy surfaces.
Information theoretic concepts now permeate almost every aspect of
molecular collision dynamics, including reaction. This is an extremely
vigorous field, and reviews appear on a yearly basis. The most recent is
again an AMCT
review, a chapter by Levine & Kinsey (11). Levine (12)
has also given a review in the 1978 volume of the Annual Review of
Physical Chemistry series. Weconsider other features of reaction dynamics here.
After describing what we do not cover in this review, we nowturn our
attention to the topics we do plan to discuss. There are three major
sections in this review: the first describes the classical dynamical approach to reactions; the second concentrates on quantum dynamics; the
third section considers somespecial topics.
Most dynamical treatments of reactive collisions use classical mechanics-assembling
the global dynamics by calculating
ensembles of
trajectories evolving on the appropriate potential surface(s). Wewould
do the calculations quantum mechanically, except that at present the
only practical quantum technology for reaction dynamics that retains
the full mathematical dimensionality of the problem is applicable only
to the simplest systems. By assuming that nuclear motion can be treated
classically, we opt instead for a practical (but at least well-defined)
dynamical approximation. The classical
approach is documented in
many reviews: The most recent are the AMCTreviews 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 methodologyapplied
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR COLLISIONS
403
atom-diatom systems. This review describes in detail how to select
initial conditions for trajectories;
the Truhlar & Muckermanreview
gives a more detailed discussion for analyzing final conditions.
Althoughclassical mechanicsis itself well defined, its applications to
molecular reactions and the interpretation of its results evolve with
experience. Here we briefly consider the theory as it is most often
applied (to atom-diatom systems), and we then discuss the problems
expect to encounter as these methods are extended to more complicated
systems.
Exact quantumscattering calculations are limited by basis-set considerations to calculations in one physical dimension. There is only a
handful of converged three-dimensional calculations. Several reviews of
quantum collision
methods are contained in Bernstein’s
AMCT
handbook. The review by Secrest (16) compares several numerical
methods cormnonlyused to solve the coupled equations characteristic of
quantumcollision 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) also
presents a separate discussion of various approximate quantum approaches to reactive scattering. Morerecently, Connor(20) has reviewed
reactive molecular collision calculations up to early 1979. This review
contains extensive literature references. Earlier reviews of quantum
reactive scattering include the 1976 review by Truhlar & Wyatt (21),
concentrating on the H+H: reaction, and the 1975 review by Micha
(22), covering quantumreactive scattering.
Exact, fully three-dimensional quantumreactive calculations are still
sufficiently difficult that very little new work has been reported in the
last year. Here we concentrate on approximation methods, because
developmentof reliable approximate techniques is essential if we are to
treat reactions quantum mechanically.
The final section briefly describes progress in several allied fields.
There is considerable activity in statistical theories of reaction, including
transition state theories and phase space theories. Truhlar (23) has
recently reviewed transition state theories, and Pechukas’ DMCB
review
(24) presents a general and very readable discussion of statistical reaction theories. At higher collision energies, collision-induced dissociation
(CID) becomesa significant alternative channel for reaction. A review
of 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 becomeincreasingly interested in whether or
not intense electromagnetic fields can affect reaction dynamics. We
conclude this review by discussing recent developments in this area.
Annual Reviews
www.annualreviews.org/aronline
404
WALKER
& LIGHT
CLASSICAL
TRAJECTORY
CALCULATIONS
Far more calculations of dynamical processes employa classical trajectory approach than any other. Classical trajectory methods have been
applied extensively to molecular systems ever since modemelectronic
computers have been available. The ease with which most problems can
be adapted to the modem minicomputer environment ensures the
viability of the method for the indefinite future. Almost any review
article describing theoretical treatments of molecular dynamics emphasizes the contribution of classical mechanics. The classical approach
itself has been reviewed often; in 1974 Porter (27) reviewed the application of classical methods from their inception up to the early 1970s. In
the introduction, we mentioned several of the most recent reviews of
classical methods (13-15). Classical calculations were also reviewed
Connor (20).
The trajectory approach to dynamics earns its popularity because it is
the only method currently available that is general enough to treat the
wide range of problems facing the dynamicist. Trajectory calculations of
atom-diatom reactions now appear routinely; the treatment today is
basically that originally described by Karplus et al (28) in 1965.
In the last year alone, manytrajectory calculations have appeared.
Because of its basic importance to the dynamics community, calculations for the H+H
2 exchange reaction continue to receive frequent
attention. Osherov et al (29) provide a sketchy account of calculations
for reaction from the initial v = 1 excited state of H2. Mayne& Toennies (30) and Mayne (31) also report calculations
for the 2
reaction, using the most recent and accurate HHHpotential surface of
Truhlar &Horowitz (32). This surface is an analytic fit to the ab initio
calculations of Liu (33) and Siegbahn & Liu (34).
Trajectory calculations are frequently used to compare theoretical
dynamics 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) consider
the effectiveness of reactant translational, vibrational, and rotational
energy in promoting reaction. Gittins & Hirst (36) report calculations
for a single surface model of the N+ + H2 reaction, in which the surface
has an extremely deep well for C2~ geometries. Schinke & Lester (37)
report calculations for the O(3P)+2 reaction, u sing t heir o wn ab i nitio
potential surface. Luntz ei al (38) report a study of the O(~D)+H2
reaction, in which both trajectory results and experimental product
rotational state distributions differ significantly from phase space theory
predictions. Herbst et al (39) studied the ion-molecule reaction O- + 2
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR COLLISIONS
405
at relatively high energies, using the multisurface "trajectory-surfaceleaking" model of Preston & Cohen(40). Jaffe (41) reports theoretical
rate constants for the atmospherically important CIO+ O reaction, and
Persky (42) reports rate constants, isotope effects, and energy disposal
features for the CI + HDreaction.
Trajectory calculations are also used to investigate the relationship
between scattering dynamicsand potential surface effects or statistical
theories. Polanyi &Schreiber (43) use trajectory results to test information theoretic predictions of branching ratios for the F + HDreaction.
Polanyi & Sathyamurthy(44, 45) use trajectories to study the relationship between features of the potential surface and dynamics for model
systems. Duff & Brumer(46) use trajectories as a tool for investigating
statistical behavior in two reactions involving long-lived complexes, the
K + NaC1and H + IC1 reactions.
In the future we will undoubtedly see an increasing number of
calculations involving more than three nuclei. Recent examples are the
four-center reactions of HBr+ C12 and HBr+ BrC1, studied by Brownet
al (47). In this calculation, both reactants and products are diatomic
molecules, but calculations are more difficult to interpret when triatomics and higher polyatomics are considered. We discuss some of
these additional difficulties later, but a flavor for the problemsinvolved
can be detected in the unimolecular decomposition reaction studies of
Hase et al (48) for C2H
5 and of Grant & Bunker (49) for ethane.
Classical
vs Quantum Studies
Apart from the practical problems that must be faced when implementing a classical trajectory calculation, there is always the conceptual
question: Howreliable is classical mechanics for treating nuclear motion in molecular systems? Truhlar & Muckerman(14) present a discussion of this point in their review. At the most detailed dynamicallevel, a
failure of classical mechanics can usually be attributed to circumstances
in which the important dynamics is associated with motion characterized by a large deBroglie wavelength. For this reason, classical
mechanics does poorly at predicting reaction thresholds and misses
resonance features altogether. On the other hand, when the dynamics of
interest involves less detail (more averaging), classical mechanics tends
to do a better job.
Of course, the best way to see howwell a classical calculation works
is to solve the problem both classically
and quantum mechanically.
Classical vs quantumstudies are very popular in the literature. Because
exact quantumreactive calculations are primarily restricted to collinear
Annual Reviews
www.annualreviews.org/aronline
406
WALKER ~ LIGHT
geometries, virtually all collinear classical calculations are performed
nowadays purely to provide a data base for classical
vs quantum
comparisons. Recent examples of this type of work may be found in
papers by Truhlar and co-workers (50-53) for the collinear I+H
2,
H+C12, H+H2, F+H2, and CI+H2 systems, and in the papers of
Connor 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 and
classical vs statistical behavior. In connection with statistical behavior in
this system, see also the related paper by Duff & Brumer (58).
Practical Considerations
Having decided to use classical trajectories to study a particular dynamical system, we must still face several practical considerations. For
the basic single surface A+BCreaction, the methodology (13-15)
known, and the problems discussed below are generally not serious. But
for polyatomic systems, many of these problems must be reconsidered
[see References (48, 49) and references cited therein].
Weconsider classical trajectory calculations in three stages.
1.
2.
3.
Wemust decide howto specify initial conditions for each trajectory.
The system is allowed to evolve according to the classical equations
of motion.
Wedecide 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. We
nowconsider some of these problems, as they occur for A + BCreactive
systems.
1NITIALCONDITIONS
The conventional wisdom in molecular scattering
calculations (whether reactive or not) is to prepare the molecular species
so that its rotational and vibrational action variables correspond to
integral quantum numbers (times/~) for the associated quantum system.
Calculations done this way are termed quasiclassical trajectory calculations; this is the methodalmost universally used in molecular scattering
systems. It is not difficult to prepare initial conditions this way for
diatomic molecules; however, in treating triatornic and higher polyatomic species, this part of the problem becomesincreasingly difficult
because nonseparability of the internal molecular degrees of freedom
becomes important. At higher energies it may no longer be possible to
find good classical action variables for such systems.
A second problem with initial conditions is statistics. After quantizing
the classical action variables, the uncertainty principle requires all
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
407
values of the conjugate angle variables to be equally likely. The commonpractice is to select these variables randomly in a multidimensional
Monte Carlo integration
procedure. The Monte Carlo approach is
normally 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 systematically
improved by running more trajectories,
thus accumulating all the resuits. (If insufficient accuracy has been obtained in the nonrandom
selection method and a greater number of trajectories
has to be integrated, then the trajectories already computedare discarded.)
Statistical
problems arise when the desired dynamical event occurs
with a low probability. In such cases, many trajectories are required
before enoughof interest are found (for example, reactive trajectories at
low collision energies). Several strategies can then be employed to
improve the initial condition sampling problem. These strategies include
importance sampling procedures, stratified
sampling procedures, and
the nonrandom initial condition selection method. These methods are
discussed in detail in recent reviews (13-15); discussion and use of the
importance sampling method may be found in several papers (43,
60-61). Furthermore, Suzukawa & Wolfsberg (62) recently presented
evidence for the improved convergence properties of the nonrandom
initial condition selection method.
INTEGRATING
TRAJECTORIES
Once an initial
condition algorithm has
been selected, integrating the equations of motion is usually straightforward. In terms of computer time, however, it is by far the most
time-consuming part of the study. Many numerical methods for integrating the usual differential equations are discussed in earlier reviews
(14, 15).
Collisions typically begin and end at separations large enough so that
the interaction between unboundparticles is negligible. Unfortunately,
the interesting dynamics (reaction and energy transfer) occurs during
the relatively short time that the particles are close together. Therefore,
a significant part of the trajectory effort is expended in the nearly
noninteracting region. Billing & Poulsen (63) propose an asymptotic
perturbation theory to reduce the expense of bringing the collision
partners close together. Their approach uses action angle variables and
is therefore more difficult to apply to polyatomic molecules.
A problem associated with trajectory integration arises whenever the
interesting dynamicsoccurs slowly on a time scale defined by the fastest
molecular motions followed. In such cases, a very large number of
trajectory integration steps becomenecessary. A standard procedure to
verify the accuracy of (selected) individual trajectories is to run them
Annual Reviews
www.annualreviews.org/aronline
408
WALKER & LIGHT
backwardsfrom their final conditions to see howwell the initial conditions are reproduced. Duff & Brumer (46, 58, 64-66) point out that
manyclassical systems exhibit instabilities (exponential divergence between trajectories initially close together in phase space) that preclude
reliable integration for extremely long times--times so long that trajectories cannot be back-integrated successfully. Related to this problem
are questions of ergodic and stochastic behavior in molecular systems
(64, 65, 67-72), which are beyond the scope of this review. These
questions are nowentering the reactive collision literature (46, 57, 58)
and more work in this area is sure to follow in the 1980s.
Although
it is straightforward to analyze classical
trajectory data and obtain estimates of averaged quantities such as rate
constants and vibrational energies, there are no reliable methods for
obtaining accurate quantum state-to-state
cross sections from these
data. Classical mechanics samples a continuous distribution of final
action variables, whereas quantum mechanics produces only discrete
distributions of action variables. To express the classical data in a
quantum mechanical language, we must somehow "’discretize’"
the
classical distribution. The rare trajectory with quantized final conditions
is of central importanceto 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 becomesinsurmountably difficult.
Several methods are commonlyused to quantize the final conditions
obtained from trajectories.
One is the histogram method, which takes
the final quantum state as the state with quantum numbers closest to
the actual nonintegral action variables obtained classically. A second
approach finds the quantum final state distribution which reproduces
(some of) the moments of the distribution obtained classically.
Both
methods have been reviewed in detail (13, 14). A third approach, called
smooth sampling, assigns trajectories to more than one quantum bin in
proportion to the difference between the actual final actions and the
nearest integral actions (55, 76-79). Howwell these procedures actually
work in practice is usually the central objective of the classical vs
quantumcomparison studies referred to earlier (50-57).
FINAL CONDITIONS
Calculations
Invoh~ing
Polyatomics
Calculations involving triatomic and higher polyatomic species are of
increasing interest. In these larger molecules the distribution of energy
among the molecular vibrational degrees of freedom may affect dynamical processes such as reaction and energy transfer. The problem of
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR
COLLISIONS
409
quantizing final states (and also of preparing initial states) becomes
much more difficult when molecules more complicated than diatomics
are treated. Unfortunately, the quantity we most want to study in
polyatomics is the energy distribution within the molecule.
In order to minimize the computational expense of trajectories,
we
want to use a coordinate system in which the equations of motion are
efficiently formulated. It has been recognized for sometime (27, 80) that
the most efficient coordinate system is the space-fixed (SF) Cartesian
reference frame. The task of preparing initial conditions and analyzing
final conditions requires the canonical transformation between SF
Cartesian coordinates and coordinates which directly express the degrees of freedom of the polyatomic in more separable terms. First, we
want to separate the overall dynamicsinto translational, rotational, and
vibrational components. Separating overall translational motion is easy.
However, Coriolis interactions make it impossible to separate cleanly
the rotational and vibrational motions of a polyatomic molecule. By
viewing the motion of the polyatomic from a body-fixed (BF) reference
frame satisfying the Eckart conditions (81-83), we minimize the Coriolis coupling between overall rotation (carried completely by the BF
frame) and vibration. Before and after collision, the total energy of the
polyatomic is constant in time, and so it is always possible to define the
total energy transfer in a collision. However,the Coriolis interaction
complicates apportioning this energy transfer between rotational and
vibrational motions, because the rotational and vibrational energies of
the polyatomic vary with time. By time averaging (83) the rotational and
vibrational energies over several (of the lowest frequency) vibrational
periods, we can define rotational and vibrational energy transfers.
At this point, we can now turn our attention to distributing the
internal energy amongthe 3N-6 vibrational degrees of freedom. Thinking quantum mechanically about the states of a polyatomic, we are
inclined to speak of internal excitations in terms of the system normal
modes. However, the idea of normal modes to the practitioner
of
quantummechanics differs significantly from the classical definition of
normal modes. Classically, normal modesrefer to the coordinate representation in which there are no off-diagonal quadratic terms in the
multidimensional Taylor series expansion of the molecular force field
about the equilibrium conformation of the molecule. The practitioner of
quantum mechanics, in referring loosely to normal modes, really means
those molecular degrees of freedom in which excitations are maintained
indefinitely in the absence of perturbations (classical constants of motion). These degrees of freedom are classical normal modes only in the
limit of infinitesimally small excitations. As soon as finite excitations are
Annual Reviews
www.annualreviews.org/aronline
410
WALTZER
~ LIGHT
allowed classically (even excitations so small that they correspond to
quantum zero point energies), then normal modesare no longer classical
constants of motion. Instead, the true constants of motion are expressed
classically in terms of the "gc~od" actions and action angle variables.
Finding these variables is a subject of intense activity beyond the scope
of this review. For recent work in this area, the reader should consult
the 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 workers
have concentrated on triatomics; at present the methods appear practical enoughthat it shoul.d be possible to perform quasiclassical trajectory
calculations for systems in which the most complicated polyatomic is a
triatomic. Howthese methods will perform for higher polyatomics is
presently unknown. Without a definition of good action variables,
trajectory calculations will be limited in their ability to describe the role
of energy distribution within a polyatomic molecule.
QUANTUM
Exact
REACTIVE
CALCULATIONS
Methods
The development of practical, methods capable of providing a dynamically accurate (quantum) description of molecular reactive scattering
processes has been a field of considerable activity within the last 15
years. At first, accurate quantumcalculations were restricted to A + BC
reactions in the one-dimensional collinear (1 D) world, whereas there are
now several accurate, fully three-dimensional (3D) quantum calculations. Even in the last four years, progress in quantumreactive scattering
2 has been reviewed often. Because of the central role the H+H
exchange reaction has played in this area, Truhlar & Wyatt reviewed
both theoretical and experimental work on this reaction in 1976 (21).
Further progress and some results
for the 3D H+H2 and
F+H
2
reactions were again presented by Wyatt (99) in 1977. The 1979 review
of reactive calculations by Connor(20) considers primarily the results
quantumcollinear calculations, and Light’s 1979 review (17) presents
formal introduction to the quantum reactive problem and discusses
several formal considerations appropriate to the quantumreactive problem. By far the most comprehensive review of theoretical
methods
available to date is that of Wyatt (18), who presents a comparative
discussion of similarities and differences in current 3Dquantumscattering theories. After a brief introduction, we consider in this section both
exact collinear and exact three-dimensional quantum approaches and
point out newly proposed methods that may extend our capabilities in
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR
COLLISIONS
411
the future. Only very limited work has been done on quantum systems
more complicated than the single surface A + BC reaction, and so here
we consider only this system.
Although other methods have been tried, the most productive current
approach for accurate quantum calculations of chemical reactions consists of a coupled-channel solution of the Schrrdinger equation, written
in natural collision coordinates (NCC). In the coupled-channel formalism, all degrees of freedom except one are expanded in a target basis
set of square integrable functions. The coupled-channel equations are
then solved numerically for motion in the final degree of freedom,
referred to variously as either the reaction, propagation, scattering, or
translational coordinate. Algorithms for solving the coupled equations
typically require computer times that go asymptotically as N3, where N
3is the number of channels in the wavefunction expansion. This N
dependence on basis-set size defines one primary goal of any close
coupling theory of quantum reactive scattering--to
devise a method
that keeps the basis set as small as possible.
Of course, any quantumclose-coupling study, reactive or not, requires
a large basis expansion if particles with large masses, strong coupling
and/or high energies are involved. Quantumdynamics is mostly concerned with light-mass low-energy systems anyway, because these presumably show the most pronounced quantum effects.
There are, however, problems peculiar to reactions that require special strategies for minimizing basis-set size requirements. One such
strategy is the choice of natural collision coordinates. Because of the
difference in asymptotic motions for reactants and products, we must
use different coordinates for them. Natural coordinates transform (at
least some of) these coordinates smoothly. Their use also minimizes the
target basis-set size because they can be chosen to represent the "natural" 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 can
choose a small internal basis to represent bound-like motion.
The second commonstrategy is prompted by the following consideration. For reactants, we want an internal basis that efficiently represents
the molecular BC vibrational and rotational motions, whereas for products, the internal basis should be equally efficient in representing the
(different)
vibrations
and rotations
of the (AB or AC) product
molecule(s). Consequently, it is commonpractice to deform the internal
basis, either continuously or by segments, along the reaction coordinate.
This procedure is unnecessary in nonreactive systems, where the deformation of the target induced by the projectile is muchless severe.
Annual Reviews
www.annualreviews.org/aronline
412
WALKER
~
LIGHT
Dynamicalapproximations for reactive scattering are interesting in so
far as they effect a reduction in the numberof coupled equations to be
solved simultaneously. Approaches such as these, including reactive
versions of the coupled states (CS) or infinite order sudden (IOS)
approximations, are discussed later.
COLLINEAR
CALCULATIONS
The most drastic (and most effective)
way
to reduce the size of the reactive basis set is to restrict the nuclei to
move along a commonline. Today, collinear quantum calculations for
the A + BC reaction are commonplaceand quite inexpensive. So long as
the mass combinations and energies of interest do not require too large
a basis set, the quantum calculation is often cheaper than the corresponding quasiclassical trajectory calculation would be. For certain
mass combinations
(H-L-H, L-H-I-L and H-H-H, where L=light,
H=heavy), large basis expansions are necessary in an NCCclosecoupling approach. The H-L-H mass combination is especially
troublesome because it generates an extreme (mass dependent) skewing
of the collinear configuration space. The NCCin such a skewed space
do not represent effectively the natural motions of the triatomic in the
strong interaction region; large basis expansions result from the coordinate defect. The L-H-H and H-H-H mass combinations are difficult
because of the preponderance of heavy nuclei.
Collinear calculations, in spite of the severity of the dynamical
approximation, are important for a variety of reasons. Newcoordinate
systems and new theoretical approaches are most easily explored for
collinear systems. For example, McNutt & Wyatt (100) discuss a generalized hyperbolic coordinate system, an extension of coordinates suggested by Witriol et al (101), which are not multivalued for ABC
configurations typical of A+B + C. Coordinates such as these may be
useful for coupled-channel treatments of collision-induced dissociation.
A variety of alternative theoretical treatments of quantum reactive
scattering have been introduced for collinear systems. Garrett & Miller
(102) and Adams & Miller (103) have proposed an exchange kernel
approach that avoids the use of natural collision coordinates at the
expense of solving coupled integro-differential
equations instead of
ordinary coupled equations. Rabitz et al (104) and Askar et al (105)
treat the collinear reactive problem using the finite-element method.
This methodis widely used in other fields (106), and also avoids using
natural collision coordinates. Askar et al (107) also suggest that
combination of the finite-element method (in the strong interaction
region) with the coupled-channel method (in the asymptotic region)
may be computationally advantageous.
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
413
A variety of numerical methods have been proposed to solve reactive
coupled-channel equations. We are naturally inclined to prefer the
R-matrix propagation method (108-111). It is a fast and general
numerical method, easy to program, stable to the inclusion of closed
channels, and automatically produces unitary S-matrices with arbitrarily
small basis expansions. There are, of course, several alternative methods, and the one actually used is mostly a matter of convenience. Those
commonlyused include Gordon’s (112) propagation method, the integral equation method proposed by Adamset al (113), and the statepath-sum method of Manz (114, 115) and Connor et al (116).
state-path-sum method and the R-matrix propagation method are similar in that they are both invariant imbedding solutions (16) to the
Schr/Sdinger equation; both methods obtain the global dynamics by
assembling solutions of many boundary value problems, each localized
in small regions (sectors) of configuration space. Recently, Basilevsky
Ryaboy (117) have also presented a sector-oriented
method for the
collinear reactive problem; to us, their approach seems very similar to
our own. Because they do not factor the scattering wavefunction optimally (118), their coupled equations are complicated by first derivative
operators. They then introduce several approximations, finally obtaining
a tractable method. With comparable numerical effort, the R-matrix
methodgives accurate solutions.
Weshould also note that several new approaches for solving coupledchannel equations have been proposed. These have been proposed for
nonreactive systems, and it would be interesting to investigate them
within a reactive framework. Johnson (119) presented a renormalized
Numerovalgorithm that shares many of the computational advantages
of 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 the
order of 200 channds) are necessary. His method obtains each colunm
of the S-matrix independently. Stechel et al (123) and Schmalz et
(124) proposed an R-matrix method using nonorthogonal coordinates
and a nonorthogonal basis to study the charge exchange problem. This
problem is interesting for quantumreactions because of its similarity to
the H-L-H mass combination. Finally, Leasure & Bowman(125) have
described a method based upon solving the Laplace transform of the
Schrrdinger equation.
Because collinear quantum calculations are feasible, much work in
this area is devoted to comparing exact quantum dynamics to dynamics
obtained by approximate methods. We have already mentioned the
popularity of these calculations for estimating the reliability of qua-
Annual Reviews
www.annualreviews.org/aronline
414
WALKER ~:
LIGHT
siclassical trajectory calculations (50-57). In addition, exact collinear
dynamics is useful in evaluating agreement with information theory
methods, transition state theories, and phase space theories. Muchof
this work has been described in Cormor’s review article (20) and
discussed later in this review.
fully converged coupled-channel
calculation of the 3D A+ BC reaction still represents the major challenge to the dynamicist. Theoretical approaches to the problem have
been described by Wyatt and co-workers [see References (18, 99) and
the references cited therein], Schatz & Kuppermann(127), and Light
and co-workers (128, 129). These methods have in turn each produced
converged calculations only for the low energy H+ H2 reaction (130134); it is gratifying that all calculations for this reaction are in substantial agreement with each other, especially at threshold energies.
Only a few accurate 3D studies have been reported since Wyatt’s
reviews (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 broadened
in the 3D reaction. Wyatt (136) observed 3D F+H2 resonance manifestations at collision energies consistent with recent experimental observations of Sparks et al (137) for the reactive differential cross section.
Both the theoretical and experimental observations represent state-ofthe-art achievements, and the similarities of these observations is remarkable.
In addition to these F+H2 results,
Redmon(138) presented preliminary (unconverged) results for the extremely difficult H +02 reaction. The H+H
2 calculation of Walker et al (134) compared results
calculations on the Porter & Karplus (139) 3 potential s urface with t he
Siegbahn & Liu (34) and Truhlar & Horowitz (32) surfaces. Lowenergy
total reactive cross sections for this surface were also given. Clary &
Nesbet (140, 141) have used these programs to investigate the evolution
of entropy in the product internal state distribution along the reaction
coordinate, with emphasis on applying information theory to 3D reactions.
There will not be, in the near future, many converged 3D quantum
reactive calculations. Basis-set size limitations preclude the accurate
study of reactions involving nuclei other than hydrogen and its isotopes.
These few accurate 3D calculations will serve as reference points upon
which to investigate approximations capable of providing dynamics of
satisfactory accuracy at computationally feasible expense.
THREE-DIMENSIONAL CALCULATIONS A
Annual Reviews
www.annualreviews.org/aronline
REACTIVEMOLECULAR
COLLISIONS
Approximate
Quantum Theories
of Reactive
415
Scattering
Due to the formidable formal and computational problems associated
¯ ~with exact quantumtreatments of reactive scattering, it is not surprising
that there is considerable current research devoted to finding approximate quantum treatments that are relatively simple and accurate. Wyatt
(19) has written an excellent review of the status of these efforts as
late 1978. Althoughwe shall concentrate here on results of the past year,
because of the importance of the area, we shall review some of the
earlier work as well. Most of these efforts fall into three basic categories: (a) distorted wave Born approximations (DWBA)
or adiabatic
perturbation 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 sudden
approximations that utilize exact quantum methods on decoupled Hamiltonians. In addition, we discuss a few other approaches that do not
readily fit into the above categories.
DISTORTEDWAVEBORNAPPROXIMATION
Although the DWBAwas developed for reactive scattering and used earlier, Choi & Tang (142)
1974 showed that the integrals over the Euler angles in the matrix
elements could be done analytically, thus greatly reducing the effort
required to apply the approximation. Webriefly review their formulation here before commentingon recent results. The DWBA
is based on
the exact equation for the differential cross section in terms of the
T-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 incoming
boundary conditions) with the approximate Hamiltonian
3.
H/s= H- V
~
and ~k~(+) is the solution (with outgoing boundary conditions) of the
complete Hamiltonian, H. To be exact, %(+) must contain all accessible
internal states, although X~-) may, with a judicious choice Of V~, contain
only a single internal state. The DWBA
consists, then, of replacing the
Annual Reviews
www.annualreviews.org/aronline
416
WALKER
~ LIGHT
exact scattering wavefunction, ~2+~, by a simpler approximate wavefunction satisfying the sameboundaryconditions. In practice ff~+~ has
been replaced by X~+~, a wavefunctioncorrespondingto a single internal state only. Thusthe DWBA
contains the coupling potential only to
first order, although both X~÷~ and Xg-~ maycorrespond to waveswith
full potential scattering and adiabatically distorted internal states. With
the Euler angle integrations performedanalytically, Choi & Tang(142)
reduced the numerical problem from an extremely lengthy fivedimensional integral to a much more manageable sum of threedimensionalintegrals. Thecalculations are nevertheless nontrivial, becausefor eachstate-to-state cross section at a given scattering angle, (a)
the adiabatic internal states must be determined at each value of the
scattering coordinate, (b) the elastic scattering wavefunctionmust
determinedfor this state, and (c) the three-dimensional integrals must
be done for all angular momentum
componentsof the initial and final
orbital angular momenta.
Remarkably,this approximationworksvery well for the H +H2 (143)
and D+H~(144) reactions at low energies. It has been applied to the
F+H
2 reaction (145) using both unperturbed and adiabatically perturbed internal states, but with H2 initially in the ground(v=0, j=0)
state. Theyfound that the adiabatically perturbed internal states producedmuchlarger reactive cross sections, but that the product internal
state distribution was largely unaffected. Althoughthe internal energy
distribution for the related F+D
2 reaction appears to be in good
agreementwith the experimental results (146), no quantitative measure
of accuracyis available.
Clary & Connor (147) used the DWBA
on the H+F2(t~=0, j=0)
reaction and comparedit with both quasiclassical trajectory results and
collinear close-coupling results. Theyalso comparedthe full adiabatic
internal states calculation to those keepingeither the initial or final
internal state unperturbed. They found quite good agreementin final
state distribution betweenexperiment(148)--the quasiclassical, the full
adiabatic DWBA,
and the "static product" DWBA--butpoor agreementwhenthe reactant internal state was not allowed to distort adiabatically in the DWBA
calculation. Theextent of agreementin this case
is truly surprising becausethe H + F2 reaction is highly exothermicand
one would expect very significant vibrational coupling in the exit
channel. Giventhe developmentstoward more general applicability of
the DWBA
methodand the encouragingresults to date, it is likely to be
tested, developed,and used morewidely in the next few years.
FRANCK-CONDON
APPROACHES
Weconsider here a number of different
approaches, most of whichutilize a Born approximationat somepoint.
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR
COLLISIONS
They are bound together
element in the form
by the commonexpression
417
of the T-matrix
where X~+) and X~-) are (approximations to) scattering wavefunctions
localized in reactant and product channels and the matrix element of
the complicated operator, ~’~, is usually further reduced to an overlap
calculation (X(o-)IX(a +)) of the nuclear wavefunctions.
The formal approach was introduced by Halavee & Shapiro (149) and
shortly followed by Schatz & Ross (150-152). Although this work was
covered in Wyatt’s review (19), there has been intense activity since
then. Concise reviews of the assumptions and development of the theory
are given in two recent articles (153, 154), together with extensions and
evaluations.
The basic assumptions commonto Franck-Condon (FC)
theories are simple to state. The exact T-matrix element for electronically adiabatic scattering on a single electronic surface (Eq. 2)
considered (and is, in fact) a Born-Oppenheimerapproximation to the
solution of the full Schrrdinger equation (including electronic coordinates) 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, corresponding to reactant and product nonreactive surfaces with a remaining
electronic perturbation term [which would produce the adiabatic (lower)
surface when diagonalized]. The T-matrix element then takes the form
(Eq. 4). Franck-Condontype theories result whenthese two surfaces are
used to define the reactant and product nuclear scattering and whenthe
remaining perturbation is taken to first order only. If the electronic
matrix element is slowly varying with nuclear coordinates, a pure FC
overlap approximation results (153):
where Met is formally the electronic matrix element between the quasiadiabatic electronic states yielding the nonreactive surfaces for reactants
and products.
Several approximations are made in deriving the Franck-Condon
theory as represented by Eq. 5. They are (153): 1. neglect of all higher
electronic states and transitions to them, 2. neglect of nuclear electronic
coupling and nuclear kinetic energy coupling terms, and 3. the assumption that the electronic matrix element is slowly varying over the region
where the wavefunctions X~+) and X~-~ overlap. In addition it appears
(154) that 4. a weak coupling approximation between the two electronic
surfaces is made (in 2 above), because the T-matrix element (Eq. 5)
only first order in the coupling potential.
Annual Reviews
www.annualreviews.org/aronline
418
WALKER
& LIGI-|T
From the Franck-Condon theory outlined above a number of FranckCondon models result from further approximations (149-154, 156).
earlier versions, very crude approximations to evaluating the channel
wavefunctions X(~+) and X~-) were coupled with approximate evaluations of the multidimensional overlap integrals. Simple analytic approximations could then be obtained, particularly for the relative internal energy distributions of products. Tworecent studies on collinear
reactions report on the adequacy of these further approximations.
Vila et al (153) constructed the quasiadiabatic channel potential
surfaces from LEPS and anti-LEPS surfaces and a coupling function
highly localized near the col. They then evaluated exactly the (twodimensional) channel functions X~+) and X~-) on these surfaces by
solving the close-coupling equations numerically and evaluated the
overlap integral in Eq. 5 numerically. They compared, then, three sets of
product vibrational distribution results--exact quantumclose-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 probable
vibrational state was almost always correct and the general shape of the
distributions was correct. Although the "exact" FC results were somewhat better than the simple FC model results, even the latter are in
good qualitative agreement. The results also appear relatively insensitive
to the only free parameter in the model--the range of the coupling
function.
Fung & Freed (154) generalized the Franck-Condon theory to polyatomic molecules (provided a single reaction coordinate is well defined).
They reduced the multidimensional overlap integral to a "rapidly convergent series of one-dimensional bound-continuum integrals,"
performed the adiabatic to quasi-adiabatic surfaces decomposition, and
(again) applied the results to a model collinear system for which exact
quantum calculations could be performed (D + IH). The channel wavefunctions were approximated using a forced oscillator model for the
half-collisions, as well as various simpler (Airy function, delta function)
approximations. They also found qualitatively
good agreement between
the exact quantum and their best FC model results for the relative
product vibrational distribution. Moreinteresting, perhaps, is that they
reported 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 probabilities
were extremely sensitive to the electronic coupling functions assumed,
varying over several orders of magnitude (~106),
even though the
relative distribution was rather insensitive. This sensitivity is presumably
Annual Reviews
www.annualreviews.org/aronline
REACTIVE
MOLECULAR
COLLISIONS 419
attributable to assumption 4 or 2 in the original derivation in which
weak coupling is assumed between the quasiadiabatic surfaces. Because
these calculations were performed at low energies (for which tunneling
is important), a high degree of sensitivity to the treatment of the col
re#on is not surprising. These results, however, do indicate that the FC
theory is currently not capable of reliably determining the.~magnitudes
of reaction cross sections, irrespective of further modelassumptions.
Other aspects of Franck-Condon models have also been scrutinized
recently. Vila et al (156) investigated the angular distributions of reaction products in 2D and 3D models for.:several potential and kinematic
coupling regimes. They used only simplified models, however, wffh
angular wavefunctions given either by free rotor or (adiabatic) hindered
rotor functions. Again the calculated angular distributions (hindered
rotor basis) qualitatively agree with exact quantumand classical trajectory calculations where available, and again absolute magnitudes were
not presented.
Very recently, Wong& Brumer (157) derived a Franck-Condon type
theory from a Faddeev decomposition of the scattering wavefunction. A
potential advantage of this approach is that three-channel A + BC~AB
+ C, AC+B reactions can easily be incorporated, whereas the quasiadiabatic decomposition of the potentials in the "standard" FC theory
is quite arbitrary. Wong& Brumer reduce their equations to a simple
form, the key to which is the solution of linear algebraic equations in
momentumspace. Comparison with exact quantum collinear results
(158) for H+C12shows excellent agreement in the predicted relative
vibrational products distribution, and again no absolute reaction probabilities were given.
The Franck-Condon approach to reactive scattering has been shown
in the last few years to be qualitatively correct in predicting relative
internal and angular distributions for simple re.active systems. Even in
its simplest forms, it provides useful results. However,even in the most
accurate forms, it appears to be completely unreliable in predicting
absolute transition magnitudes and therefore absolute cross sections.
This maybe a defect basic to the theory in the sense that localizing the
overlap of the quasiadiabatic scattering solutions requires large interaction potentials between the quasiadiabatic surfaces and thereby invalidates perturbation theory. Extension of the interaction region, while
reducing the perturbations, will degrade the accuracy of the overlap
approximation. Thus the approximations inherent in the FC theory may
limit its accuracy. The optimal choice of surfaces has been considered
by Mukamel& Ross (155) and a first order K-matrix approximation
include higher orders in the reactive perturbation was given. Its use by
Annual Reviews
www.annualreviews.org/aronline
420
WALKER ~ LIGHT
Fung & Freed (154), however, did not appear to help the absolute
accuracy. Until this can be remedied, the FC theory must be considered
valuable as a guide to internal energy distributions only.
DECOUPLING APPROXIMATIONS Because very many closecoupled equations are required for the exact description of most atomdiatom systems, much attention has been given recently to approximations that reduce the dimensionality of the coupled equations and still
permit the evaluation of state-to-state
cross sections. Twoof these
approximations, the J,-conserving approximation and the infinite order
sudden approximation (IOS), have recently been applied to reactive
scattering.
The only reviews to date that consider applications of Jz-conserving
approximations to reactive collisions is again that of Wyatt (18, 19). The
Jz-conserving work of Elkowitz & Wyatt (159) and Kuppermann et
(160) on H+H
2, and of Redmon & Wyatt (161) on F+H2, is reviewed
there.
The centrifugal sudden (CS) approximation was introduced by McGuire &Kouri (162) and Pack (163) for inelastic scattering (164).
inelastic scattering the CS approximations can be made directly by
replacing the orbital angular momentumoperator in a space-fixed
coordinate system by an average eigenvalue, [([+ 1)h2. Cross sections
are then obtained by using "/-labeled" S-matrix elements and recoupling
with appropriate sums over Clebsch-Gordan coefficients
(165-168).
Because body-fixed coordinates are used in reactive scattering, only the
primary assumption of the CS approximation (the neglect of Coriolistype coupling) was used by Wyatt et al and Kuppermalmet al. They
decouple the original complete set of coupled equations into independent sets, one for each value of the projection of the total angular
momentumonto the body-fixed axis (165). Additional approximations
may then be made to simplify the equations further (166-168).
Generalizing this approximation to reactive scattering is made difficult because the body-fixed axes of reactants and products must differ.
Thus, the approximation, which imposes a conservation condition in the
body-fixed frame, is not the same for reactions and products, and leads
to continuity problems. Elkowitz & Wyatt (159) and Kuppermannet
(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 for
reactants to that for products. In this way, the CS approximation is
imposed uniformly throughout the reaction, albeit at the cost of neglecting somecomple.x effective potential terms.
REACTIVE
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
421
Kuppermannet al (160) essentially switch coordinates on a surface,
matching wavefunction and derivatives
there. Continuity can only
be preserved by coupling the set of solutions (for various values of K,
the "’J,-conserved’" quantity) for reactants with the set of solutions for
the products. Both procedures appear to work, but greatly increase the
complexity of the Jz-conserving approximation over that required for
inelastic collisions. Nevertheless, this approximationgreatly reduces the
number of coupled equations to be solved, and permits the quantum
calculation of reaction probabilities to be extended to heavier systems
for which the number of open channels involved would prohibit exact
close-coupling calculations at the present time [e.g. F+ H2; see Reference (161)].
The infinite order sudden (IOS) approximation, also reviewed
Kouri (164) for inelastic collisions, has recently been developed (169171) for and applied (171-173) to reactive scattering. In the simplest
terms (174), the lOS approximation for inelastic scattering consists
replacing the eigenvalues of the rotational angular momentum
operator
(in the CSequations) by a single effective energy; by utilizing closure
the rotational functions, we obtain a set of coupled equations for
vibrational states containing the internal (A-BC)angle as a parameter.
Thus the rotational manifold of states disappears completely from the
coupled equations, being replaced by the necessity to solve this very
much easier problem at a set of angles. The angle-dependent T-matrix
elements are then projected onto the asymptotic rotor states to determine transition probabilities. This approximation does not conserve
energy, even for inelastic collisions.
The difficulties of extending the IOS to reactive scattering are even
greater than for the J~-conserving approximation alone, because the
approximation is more extreme in each chemical channel and, although
the wavefunction matching on a channel-matching surface appears
straightforward, the momentum(or derivative) matching appears more
difficult to justify (169-171, 175). Barg & Drolshagen (170) confine
their discussion to a two channel light-heavy-light
atom (L-H-L)
exchange reaction where they find that a one-to-one angle matching on
the surface produces complete orbital-to-rotational
angular momentum
conversion (and vice versa) on the matching surface. Their preliminary
results (176) on D + HC1"show agreement with quasi-calculations...as
goodas a factor of one to three for the reactive integral rotational cross
sections."
The work of Bowman& Lee (171, 172) appears to be more general,
allowing for asymmetric three-channel reactions. They, however, assume that l~=la and J~=Jl~ on the matching surface (although the
Annual Reviews
www.annualreviews.org/aronline
422
WALKER& LIGHT
opposite result of Barg &Drolshagen could be used as well), thus also
finding a one-to-one angular matching procedure. In applications to
H + H2 they find (171) that their low energy cross sections overestimate
the 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 not
conserved and rotational transitions occur well into closed channels, the
product rotational distributions are in poor qualitative agreement with
exact results.
The matching equations derived by Khare et al (169) are more
general than those above, but similar simplifications are suggested [cf
Eqs. 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, but
improves somewhat with increasing energy.
In summary, the search for an adequate lOS description of reactive
collisions has been very active, and with good reason. The IOS approximation permits one to replace a very large intractable set of
equations by numerous very small sets (involving vibrational states
only) for describing quantum reactive scattering. The lOS approximations, however, are quite strong and it remains to be seen whether this
approach will produce scattering results more accurate than classical
trajectory results. The preliminary results to date offer somehope, but
also 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.
OTHERQUANTUM
APPROXIMATIONS
Several recent models of reactive
systems have been proposed in which a portion of the dynamics is done
quantummechanically, the remainder being treated classically, statistically, or by a crude quantummodel. Weshall discuss these briefly.
Fischer et al (177) proposed a "statistical
dynamic" model of exchange reactions in which the vibrational-translational
degrees of freedomare treated in the FC manner (see above), but with a statistical
treatment of rotational degrees of freedom. By assuming the factorization of T-matrix elements into independent elements for the different
degrees of freedom and utilizing the concept of an internal temperature
for the "statistical"
degrees of freedom, they obtain approximate analytic results for product energy distributions.
Comparison of these
distributions with experiment for H(D) + F2(C12) and C1 + H(D)I
quite encouraging agreement (178). Correlations in the distributions for
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
423
the various degrees of freedom are suppressed and absolute cross
sections presumably will not be obtained accurately.
A related approach by Baer (179) attempts to take distributions
known from exact quantum calculations on reduced (1D or 2D) models
and, using statistical
arguments, generate full quantumdistributions.
Relatively
good
agreement
with classical trajectory results on H + Br
2
was obtained from 1D quantum vibrational distributions
using this
method.
Basilevsky & gyaboy, after deriving a collinear three atom reactive
scattering method (117), make quasiclassical
approximations on the
collinear dynamics (180, 181) and finally obtain a very crude analytical
solution, which, however, may give some qualitative insight into the
causes of vibrational population inversions.
Finally, Wyatt &.~ Walker (182) and Wyatt (183) used full quantum
close-coupling methodsover a portion of the lEVI- and 3Y~u+surfaces of
the F+H2 system to study quenching of the F(2PI/2) in the reactive
system. Because the ground state reaction is very exothermic and cannot
be handled exactly by current close-coupling methods, he included the
effects of reaction by applying the detailed quantum transition state
theory (DQTST)(184, 185) at a surface just beyond the
OTHERTOPICS
TransitionState andStatistical Theories
Although the transition state theory (TST) of chemical reactions has
been 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 (primarily TST) including work up to 1976. Chesnavich & Bowers (186)
1979 review both TSTand other statistical theories and their application
to ion-molecule reactions. Garrett & Truhlar (23, 187) give very extensive references.
It has been known for many years that classical transition state
theory, in either the canonical or microcanonical form, provides an
upper bound on classical reaction rates. Pechukas & Pollak (188-190)
have, in the past few years, defined the mathematical conditions required 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 for
which a unique transition state exists. (In their definition, the transition
Annual Reviews
www.annualreviews.org/aronline
424
WALKER
~
LIGHT
state is defined by a periodic vibration in the interaction region. More
than one such periodic trajectory in the interaction region maydefine a
collision complex.) Althoughextension of these rigorous classical results
to 3D and the examination of real potentials to determine their applicability may be difficult,
at least the mathematical framework for this
examination has been constructed.
Garrett & Truhlar (187, 191) considered the relationship of variational microcanonical TST(which should yield the least upper bound to
the rate constant at fixed E) with other more approximate (but more
easily applied) formulations such as the minimumdensity of states
criteria (192, 193) and Miller’s unified statistical
theory (194).
recently, they applied their generalized theory (canonical variational
TST) to a number of 3D reactions, showing it superior to conventional
TST(195, 196). In comparing the results of all these theories with the
exact classical canonical rate constants determined from classical trajectory calculations for nine collinear triatomic systems, Garrett & Truhlar
(187, 191) show that the microcanonical variational and unified statistical theories are superior to conventional canonical and canonical variational theories, although the differences are not great. On the other
hand, quantized versions of these approaches show that the generalized
variational TST can give significant improvements over conventional
TST(197).
Truhlar has also considered the effects of the classical nature of most
TST calculations by comparing them with quantum calculations (23).
The accuracy of quantum tunneling corrections was investigated by
Garrett & Truhlar for the several collinear systems for which accurate
quantumcalculations are available (198). Most recently they applied the
generalized TST theories with quantum corrections to the H + D2 and
D + H2 reactions using the accurate a priori surface of Liu & Siegbahn
(33, 34) and found excellent agreement for both rate constants and
kinetic isotope effects (199) with the best available experimental data.
From this last work it appears that TST, when carefully applied, should
produce rate constants of quite useful accuracy (within a factor of 2)
least for reactions with a single activation energy barrier and with
reasonable skewing angles. It should, however, be tested further on
qualitatively different potential energy surfaces and at higher energies.
Weshall not detail all the work using TST, but only mention the
significant new work for ion-molecule reactions presented by Chesnavich & Bowers (186), and the construction of a TST for three-body
ion-molecule recombination reactions considered by Herbst (200). Finally, Chesnavich et al (201) also used microcanonical variational TST
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
425
and classical trajectories to obtain an upper bound for capture cross
sections in ion-dipole collisions (long a problem because of the longrange anisotropic potential).
They find good agreement between the
bound, trajectory, and experimental results.
There has been some further development of statistical
theories in
which a strong coupling collision complex is formed. In addition to the
relevant work mentioned above (23, 196, 197), Brumer et al (46, 202)
have investigated, via classical trajectory studies, the dependence of
complex formation on geometrical and mass factors, as well as energy
and potential energy characteristics.
They characterize the subset of
trajectories
that form complexes by passage through a region of the
potential surface in which substantial exponential divergence takes
place. In comparing the product distributions of this subset with that
obtained from the Wagner & Parks formulation (203) of statistical
theories in terms of a strong coupling region, they find that the criterion
of substantial exponential divergence indeed produces a statistical distribution in qualitative agreement with phase space theory. A more
detailed analysis of the subset of trajectories showedclearly the statistical nature of the distributions independent of the specific assumptions
of phase space theory.
Pollak & Levine (204) have recently shown how the definition of
strong coupling region and assumption of maximalentropy of the final
distributions can lead to a variety of statistical theories depending on
constraints imposed on the distributions. They show that Miller’s unified theory, TST, and phase space theory can be related by constraints
on flux across appropriate boundarysurfaces. In application to collinear
H + H2, they find that addition of a second constraint [average number
of (re)crossings of a surface] produces excellent agreement with the
probability of reaction determined by trajectory calculations. However,
trajectory calculations are required to determine the constraint, which is
then used in the statistical theory.
Pollak & Pechukas (205) also re-derived Miller’s unified statistical
theory, applied it to collinear H+ HE, and derived a lower bound as
well as an upper bound on the reaction probability. It proved to be
remarkably accurate.
Wehave considered information theory approaches to reaction dynamics beyond the scope of this review. Wetake this position not
because the work is not relevant--it
is--but because it has been
reviewed with some frequency and space limitations preclude its inclusion here. The interested reader should see reviews by Levine and
co-workers (11, 12).
Annual Reviews
www.annualreviews.org/aronline
426
WALKER& LIGHT
Collision-Induced Dissociation
Wheneverthe relative A+BCcollision energy exceeds the BCbond
energy, we open up the product reaction channel A+B+C.Such
processes are called collision-induced dissociation (CID)and have been
investigated using (extensions of) methodsdevelopedfor rearrangement
collisions. Both classical and quantummethodshave been used to study
CID;for detailed recent reviews, see the discussions of Kuntz(25) and
Diestler (26).
Most recent work on CIDuses the quasiclassical trajectory approach
and focuses attention on the dependenceof the dissociation rate (or
cross section) on the initial internal molecularenergy(206-211).Molecular rotational energy plays an important part in the dissociation
dynamics(206, 208, 209, 211), suggesting that a purely vibrational
ladder-climbing model oversimplifies the actual process. The ladderclimbing modelallows only a single vibrational quantumtransition per
collision, with dissociation occurring only from the highest bound
vibrational state. The actual process is better characterized by a combined vibration-rotation
ladder climbing (206) in which the dissociation
step occurs from molecules whosetotal energy is near the dissociation
limit, but with a significant spread of internal energybetweenvibration
and rotation (208, 209, 211). Blais & Truhlar (209) and Lehr & Birks
(211) present excellent discussions of the importance of rotations
CIDand should be consulted for references to earlier literature.
Semielassieal (212, 213) and quantummechanical(214-218) calculations of CIDhave been reported for collinear ABCsystems. Semiclassical and time independent quantummethods(212-216) require a discretization of the continuummolecularstates; dissociation is then defined
as a sum over transitions to the continuum. Time-dependent wavepacket approaches (217, 218) avoid this difficulty; the dissociation
probability is determinedby projecting out the boundstate components
of the wavefunctionat large times. These calculations demonstratethat
dissociation occurs from vibrational states well below the continuum
and that the ladder-climbing model oversimplifies the dynamicseven
without rotations.
Considering the importance of rotations demonstrated by the 3D
classical calculations discussedpreviously, it is doubtful that collinear
calculations will be quantitatively reliable for CID.Extendingquantum
CIDmethodsto higher dimensions(to include rotations) will cause
substantial increase in complexity. In the near future, detailed CID
dynamicswill be best treated by quasiclassical techniques. Thesecalculations will be accurate to the extent that goodpotential energysurfaces
Annual Reviews
www.annualreviews.org/aronline
REACTIVEMOLECULAR
COLLISIONS 427
are available and to the extent that quantum effects (presumably near
the threshold for dissociation) are minimal. Studying quantumeffects in
three dimensional CID will require substantial future developments in
the methodology of the problem.
Laser-Collision
Reactive Processes
Although lasers have been used for many years to modify the initial
states of reactants, to analyze the products of reactions, and to cause
(collisionless) photodissociation, it is only in the last five years or so that
the effects of strong laser fields on collision dynamics have been
studied. Because the field of multiphoton dissociation has been the
subject of recent conference proceedings (219) and because lasers used
for reactant preparation and product detection are primarily experimental tools for studying state-to-state chemical reactions, we shall comment here only on recent developments in laser-collision
dynamics,
particularly reaction dynamics. To date the preponderance of work in
this area (meager though it is) has been theoretical
and the only
laser-collision-induced
reactive processes observed have been laser+ +e-) and, very
induced associative ionization (220) (A +B+hv-->AB
recently, 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 continuum
previously developed for field free associative ionization (223) and the
"electronic-field" representation of the scattering process in the laser
field, also advocated by George et al (224). In this representation, the
zero order-time independent states are taken to be those determined by
the fixed nuclei electrons plus field Hamiltonian, and scattering (with or
without 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 investigated
laser-induced unimolecular dissociation (225) and electronically inelastic scattering of F on H2 (226). Yuan & George (227) investigated
the reactive scattering of F on H2 in the presence of a laser field using
semiclassical methods.
Light & Altenberger-Siczek (228) also used this framework to consider the effects of near resonant laser radiation on atom-diatom exchange reactions in collinear, two electronic surface models. By assuming near resonant fields and electronically allowed transition couplings,
they showed that significant alterations in transition probabilities to
excited states could be obtained, although "true" laser catalysis would
Annual Reviews
www.annualreviews.org/aronline
428
WALKER & LIGHT
be unlikely. The recent experiments of Hering et al (221) appear
2) laser
confirm that such processes can occur at modest (< 107W/cm
-17
powers, but the cross sections appear to be small (~10
cm2).
An alternative approach to these processes has been taken by Copeland (229) and by Orel & Miller (230, 231). Copelandderives, but
not apply, semiclassical dynamicalequations in a time-dependentrepresentation not utilizing the "electronic-field" states. Orel &Miller treat
everything classically, with the laser field modifyingthe classical dynamicsof the motion of the nuclei on a single electronic energy surface
due to a time dependent perturbation (the laser field). They show that
the field will affect the dynamicseven for infrared inactive species due
to the lack of symmetryof the motion along the reaction coordinate and
that the effective threshold for reaction will be lowered. In a qualitative
analytic model, however, they also show that laser powers on the order
2 would be required to produce significant effects.
of 101°- 1012W/cm
Overall, the research in this area has producedmixedresults. Processes
are reasonably well understood theoretically--the
frameworkof both
time-independent multiple surface approaches and time-dependent
single surface approacheshas been established. However,the paucity of
experimental data confirms the general trend of theoretical prediction
that to makelaser-collision-induced processes occur with reasonable
probability, very strong (~> 109W/cm2) fields are required. In addition,
the difficulty of obtaining experimentaldata suggests that the utility of
these processes as probes of potential energy surfaces of reaction
complexesis severely limited and that the prognosis for such processes
as general synthetic tools is poor.
ACKNOWLEDGMENTS
This work was performed in part under the auspices of the US Department of Energy. Partial support by the National Science Foundation
under NSFgrant numberCHE79-06896is gratefully acknowledged.
Literature Cited
1.
2.
3.
4.
Levine, R. D., Bernstein, R. B. 1974.
Molecular Reaction Dynamics. New
York: Oxford Univ. Press, 250 pp.
Child, M. S. 1974. Molecular Collision Theory. London & New York:
Academic. 300 pp.
Bernstein, R. B., ed. 1979. AtomMolecule Collision Theory, A Guide
for the Experimentalist. New York&
London: Plenum. 779 pp.
Miller, W. H., ed. 1976. Dynamicsof
Molecular Collisions,
Pt. A. New
York & London: Plenum. 318 pp.
5.
Miller, W. H., ed. 1976. Dynamicsof
Molecular Collisions,
Pt. B. New
York & 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
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
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. See
Ref. 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. 43:325987
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.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
429
Herbst, E., Payne, L. G., Champion,
R. L., Doverspik¢, L. D. 1979. Chem.
Phys. 42:413-21
Preston, R. K., Cohen, J. S. 1976. J.
Chem. Phys. 65:1589-90
Jaffe,
R. L. 1979. Chem. Phys.
40:185-206
Persky, A. 1979. J. Chem. Phys.
70:3910-19
Polanyi, J. C., Schreiber, J. L. 1978.
Chem. Phys. 31:113-36
Polanyi, J. C., Sathyamurthy, N.
1978. Chem. Phys. 33:287-303
Polanyi, J. C., Sathyamurthy, N.
1979. Chem. Phys. 37:259-64
Duff, J. W., Brumer, P. 1979. J.
Chem. Phys. 71:2693-702
Brown, J. C., Bass, H. E., Thompson, D. L. 1979. J. Chem. Phys.
70:2326-34
Hase, W. L., Wolf, R. J., Sloane, C.
S. 1979. J. Chem. Phys. 71:2911-28
Grant, E. R., Bunker, D. L. 1978. J.
Chem. Phys. 68:628-36
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
Gray, J. C., Garrett, B. C., Truhlar,
D. G. 1979. J. Chem. Phys. 70:592122
Truhlar, D. G. 1979. J. Phys. Chem.
83:188-99
Gray, J. C., Truhlar, D. G., Baer, M.
1979. J. Phys. Chem. 83:i045-52
Connor, J. N. L., Jakubetz, W.,
Lagan~, A. 1979. J. Phys. Chem.
83:73-78
Connor, J. N. L., Lagan~, A. 1979.
Comput. Phys. Commun. 17:145-48
Connor, J. N. L., Lagan~, A. 1979.
Mol. Phys. 38:657-67
Hutchinson, J. S., Wyatt, R. E. 1979.
J. Chem. Phys. 70:3509-23
Duff, J. W., Brumer, P. 1979. J.
Chem. Phys. 71:3895-96
Deleted in proof
Faist, M. B., Muckerman, J. T.,
Schubert, F. E. 1978. J. Chem. Phys.
69:4087-96
Muekerman,J. T., Faist, M. B. 1979.
J. Phys. Chem. 83:79-87
Suzukawa, H. H. Jr., Wolfsberg, M.
1978. J. Chem. Phys. 68:1423-25
Billing, G. D., Poulsen, L. L. 1979.
Chem. Phys. Lett. 66:177-28
Brumer, P. 1973. J. Comput. Phys.
14:391-419
Bruraer, P., Dtfff, J. W. 1976. J.
Chem. Phys. 65:3566-74
Annual Reviews
www.annualreviews.org/aronline
430
66.
WALKER & LIGHT
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. Quantum
Chem. @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 Quantum
Physics, 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.
94.
95.
96.
97.
98.
99.
100,
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
I 11.
112.
113.
114.
115.
116.
117.
Handy, N. C., Colwell, S. M., Miller,
W. H. 1977. Faraday Discuss. Chem.
Soc. 62:29-39
Miller, W. H. 1977. Faraday Discuss.
Chem. Soc. 62:40-46
Sorbie,
K. S. 1976. MoL Phys.
32:1577-90
Sorbic, K. S., Handy, N. C. 1976.
MoL Phys. 32:1322-47
Colwell, S. M., Handy, N. C. 1978.
Mol. Phys. 35:1183-90
Jaffe, C., Rehahardt, W. P. 1979. J.
Chem. Phys. 71:1862-69
Wyatt, R. E. 1977. In State to State
Chemistry, ACSSyrup. Ser. No. 56,
exl. P. R. Brooks, E. F. Hayes, pp.
185-205.
Washington
DC: Am.
Chem. Soc. 259 pp.
McNutt, J. F., Wyatt, R. E. 1979. J.
Chem. Phys. 70:5307-12
Witriol, N. M., Stettler, J. D., Rather,
M. A., Sabin, J. R., Trickey, S. B.
1977. J. Chem. Phys. 66:1141-59
Garrett, B. C., Miller, W. H. 1978. J.
Chem. Phys. 68:4051-55
Adams,J. E., Miller, W. H. 1979. J.
Phys. Chem. 83:1505-7
Rabitz, H., Askar, A., Cakmak, A. S.
1978. Chem. Phys. 29:61-65
Askar, A., Cakmak, A. S., Rabitz, H.
1978. Chem. Phys. 33:267-86
Baker, A. J., Soliman, M. O. 1979. J.
Comput. Phys. 32:289-324
Askar, A., Cakmak, A. S., Rabitz, H.
1980. J. Chem. Phys. 72:5287-89
Light, J. C., Walker, R. B. 1976. J.
Chem. Phys. 65:4272-82
Stechel, E. B., Walker, R. B., Light,
J. C. 1978. J. Chem. Phys. 69:351831
Light, J. C., Walker, R. B., Steehel,
E. B., Schmalz, T. G. 1979. Comput.
Phys. Commun. 17:89-97
Walker, R. B. 1978. Quantum Chem.
Program Exch., Program No. 352.
Dep. Chem., Indiana
Univ.,
Bloomington
Gordon, R. G. 1969. J. Chem. Phys.
51:14-25
Adams,J. T., Smith, R. L., Hayes, E.
F. 1974. J. Chem. Phys. 61:2193-99
Manz, J. 1974. Mol. Phys. 28:399413
Manz, J. 1975. Mol. Phys. 30:899910
Connor, J. N. L., Jakubetz, W.,
Manz, J. 1975. Mol. Phys. 29:347-55
Basilevsky, M. V., Ryaboy, V. M.
1979. Chem. Phys. 41:461-76
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
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. Workshop
Algorithms Comput. Codes At. and
Mol. 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. 70:564059
124. Schmalz, T. G., Stechel, E. B., Light,
J. C. 1979. J. Chem. Phys. 70:566071
125. Leasure, S., Bowman,J. M. 1978. J.
Chem. Phys. 68:2825-30
126. Deleted in proof
127. Schatz, G. C., Kuppermann,A. 1976.
J. Chem. Phys. 65:4642-67
128. Walker, R. B., Light, J. C., Altenberger-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. 69:292223
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. Quantum
Chem. 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.
431
Chem. Phys. 61:5147-57
143. Choi, B. H., Tang, K. T. 1976. J.
Chem. Phys. 65:5161-80
144. Tang, K. T., Choi, B. H. 1975. J.
Chem. Phys. 62:3642-58
145. 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. 53:338587
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. T
1978. J. Chem. Phys. 69:4419-30
Annual Reviews
www.annualreviews.org/aronline
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~r
Str6mungsforschung,
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.,Soc.
Truhlar,
D. G. 1979.
J. Am. Chem.
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. 63:201934
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. NewYork: Wiley
220. Hellfeld, A. V., Caddick, J., Weiner,
J. 1978. Phys. Rev. Lett. 40:1369-73
Annual Reviews
www.annualreviews.org/aronline
REACTIVE MOLECULARCOLLISIONS
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., Zimmerman, I. H., Laing, J. R. 1977. Faraday Discuss.
Chem. Soc. 62:
246-54
225. Yuan, J.-M., George, T. F. 1978. J.
Chem. Phys. 68:3040-52
221.
226.
227.
228.
229.
230.
231.
433
DeVries, P. L., George, T. F. 1979. J.
Chem. Phys. 71:1543-49
Yuan, J.-M., George, T. F. 1979. J.
Chem. Phys. 70:990-94
Light, J. C., Altenberger-Siczek, A,
1979. J. Chem. Phys. 70:4108-16
Copeland, D. A. 1978. J. Chem. Phys.
68:3005-15
Orel, A. E., Miller, W. H. 1978.
Chem. Phys. Lett. 57:362-63
Orel, A. E., Miller, W. H. 1979. J.
Chem. Phys. 70:4393-99