Annual Reviews
www.annualreviews.org/aronline
Ann. Rev. Phys. Chem.1984.35: 159~89
Copyright©1984by AnnualReviewsInc. All rights reserved
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION
STATE THEORY
Donald G. Truhlar
Departmentof Chemistry, University of Minnesota,Minneapolis,
Minnesota 55455
Bruce C. Garrett
ChemicalDynamicsCorporation, 1550 W. Henderson Road,
Columbus,Ohio 43220
INTRODUCTION
1 is the mostwidelyusedtheory for calculating
Transitionstate theory(TST)
rates of bimolecularreactions occurringin the gas phaseand in condensed
phases. TSTis also incorporated into the widely used RRKM
theory for
unimolecular reactions. The popularity of TSTis largely due to its
simplicityandits usefulnessfor correlatingtrendsin reactionrates in terms
of easily interpretedquantities. Severalformsof variationaltransition state
theory (VTST)have been proposed, one as early as 1937; however,until
recently, most applications of TSThave beenlimited to the conventional
(nonvariational) formulation. In recent years there has been renewed
interest in VTST
for providinginsights into the factors controllingchemical
t Analphabetical list of the abbreviations used in this article is as follows : CVT,canonical
variational transition state theory; DA,dynamical-pathvibrational-average tunneling
approximation;
GTST,generalized
transition
state
theory; ICVT, improved canonical
variational transition state theory; LA,least-action tunnelingapproximation;LAG,leastaction ground-statetransmissioncoefficient; LC,large-curvaturetunneling approximation;
LCG,large-curvature ground-state transmission coefficient; MCP,Marcus-Coltrin path; PA,
phase-average tunneling approximation; RRKM,Rice-Ramsperger-Kassel-Marcus; SAG,
semiclassical vibrationally(-rotationally) adiabatic ground-state transmission coefficient ; SC,
small-curvature tunneling approximation ; SO, second-order tunneling approximation ; SOP,
semiclassical optical potential tunneling approximation;TST,transition state theory; US,
unified statistical model;VA,vibrational-average tunneling approximation;VTST,variational transition state theory; #VT,microcanonical variational transition state theory.
159
0066-426X/84/1101-0159
$02.00
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
160
TRUHLAR & GARRETT
reaction rates, and VTSThas been developed into a practical quantitative
tool. The present review is concerned with the most recent developments,
and we shall aim to complementrather than duplicate several other recent
reviews, which we now summarize.
Chesnavich & Bowers (30, 31) reviewed applications of statistical
methods in gas-phase ion chemistry, including detailed discussions of
transition-state switching models and applications of VTSTto ion-dipole
capture. Walker & Light (188) reviewed the theory of reactive collisions,
including progress to date on VTST.Truhlar & Garrett (173) provided
introduction to VTSTand an overview of their early work on the subject.
Twolater articles, by Garrett et al (67) and Truhlar et al (181), provided
partial reviews of selected aspects of further work of this group; these
articles are partly review and partly new material, and they are discussed
further below. Pechukas (124, 125) reviewed recent developments
transition state theory, including VTST,and we especially wish to single out
his discussions of quantal and semiclassical approaches and of periodicorbit dividing surfaces. Pollak (138) has also reviewed periodic-orbit
dividing surfaces and related topics. Laidler & King (95) reviewed the
historical origins of transitions state theory, excluding VTST, with
coverageup to about 1938; I-Iirschfelder (81) provided additional historical
comments.Hase (79) reviewed the history and use of variational concepts
unimolcular rate theory. Truhlar et al (177) reviewed the current status
transition state theory, including VTST,with special emphasis on’ the
validity of the equilibrium and dynamical bottleneck assumptions, on
localized states in unimolecularlydecayingsystems, and on frictional effects
in solution-phase reactions. Truhlar et al (180) wrote a handbook-type
chapter and concentrated on the practical aspects of VTSTcalculations,
with emphasis on reactions of polyatomics, anharmonicity, tunneling, and
other corrections.
See (200) on TST and VTST, emphasizing organic
applications.
The present review of VTSTconcentrates on work reported since the
previous review in this series, by Pechukas(124). Readers are referred
Garrett & Truhlar (56) and to the reviews mentioned above for more
extensive references to earlier work. Wealso restrict the present review to
gas-phase reactions.
In addition to variational transition state theory, this chapter briefly
considers relevant recent developmentsin selected aspects of several related
subjects: related dynamical theories, the role of tunneling in chemical
reactions, the calculation of dynamical bottlenecks and rate constants for
state-selected processes, the role of resonances in chemical reactions, and
vibrational bonding.
Annual Reviews
www.annualreviews.org/aronline
VARIATIONAL
TRANSITION
STATE THt/ORY
161
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
BASIC CONCEPTS
Since VTSTis introduced in several of the papers above, we give here only a
brief review as backgroundto the later sections.
Transition state theory is a statistical mechanical theory of chemical
reaction rates that may be derived from two fundamental assumptions.
First one defines a reaction coordinate s leading from reactants (negative s)
to products (positive s) and a (generalized) transition state as a system
way between reactants and products with a fixed value for s (thus the
transition state is a system with one less degree of freedom than the
reactants). The first assumptionis that transition state species that originate
as reactants are in local equilibrium with reactants. The second assumption
is that any system passing through the transition state does so only once
(before the next collision or before it is stabilized or thermalized as
reactant or product). These assumptionsmaybe called the local-equilibrium
and no-recrossing assumptions. Early workers were aware that the validity
of the no-recrossing assumption dependson the location of the transition
state, and that the transition state maybe variationally defined as the
phase-space hypersurface with the least one-wayflux through it (82, 192)
as the hypersurfacethat yields the smallest free energyof activation (48). But
in all conventional formulations of transition state theory, the transition
state passes through a saddlepoint on a potential energy surface and the
omitted coordinate s is taken as the unbound saddlepoint normal mode
(49). Variational transition state theory (VTST)is the namewe apply
theories that use the minimum-fluxor maximum-free-energy-of-activation
criteria to choose the transition state. VTSTdoes not provide exact
expressions for rate constants becauseit still involves the local-equilibrium
assumption and because additional approximations are required to
translate the flux-through-a-hypersurface argumentinto practical terms in
a quantummechanical world. Furthermore the variational search for the
best transition state is usually carried out with constraints (e.g, onedimensional search in coordinate space) for practical reasons.
Althoughthe early statements of the classical variational principle by
Wigner(192) and Horiuti (82) are correct and clear, the reader should
warned that there has also been some confusion about VTST.Thus Evans
(48) incorrectly implied that the maximumfree energy of activation
corresponds to a minimum-probability rather than a minimum-flux
condition and that the minimum-free-energy transition state passes
through the saddlep.oint. Later workers sometimes confused the minimumflux condition with a minimum-density-of-states condition. Someworkers
attempted to discuss variational ideas in terms of free energy surfaces as
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
162
TRUHLAR & GARRETT
functions of more than one coordinate, whereas the free energy to be
minimizedis a function of the location of a transition state surface, not a
function of all the coordinates of the system. For another example, Swarc
(169) provided a correct description of howto carry out a maximum-freeenergy-of-activation calculation with classical reaction-coordinate motion,
but Eyring (50) commented
incorrectly that it is not possible to define the
free energy of activation if the transition-state hypersurface does not pass
through a stationary point of the potential energy, i.e. the saddlepoint.
Swarc himself was unsure how to compare calculations with different
reaction coordinates, but correct appreciation of the variational criterion
showsthat it applies to arbitrary variations in the transition state, not just
to different choices for a given reaction coordinate.
Modern appreciation of VTSTconcepts includes the discussion of
Eliason &Hirschfelder (47) of the relationship of a variational criterion for
the transition-state-theory
rate constant to collision theory and the
applications by Keck(89, see also 90) of variational theory to atom-atom
recombination in the presence of a third body. Keck (91) also presented
variational theory in a more general context. Bunker & Pattengill (23),
Marcus (102, 103), and Wong& Marcus (196) proposed related schemes
(not VTST)for unimolecular and bimolecular reactions (see also 55,
171). Tweedale & Laidler (185) provided an example of a free-energy-ofactivation curve as a function of reaction coordinate for a collinear atomdiatom reaction, and Quack & Troe (145-149) used VTSTand related
dynamical schemes for a series of calculations on unimolecular decompositions of triatomics. In the last .five years or so there has been
considerable activity in elucidating the classical mechanicsof variational
transition states; general techniques have been proposed for calculating
free energy of activation profiles from potential energy surfaces and for
performing VTSTcalculations for arbitrary systems; and variational
transition state theory including important quantization and tunneling
effects has begunto receive extensive testing as a general practical tool for
for the calculation of bimolecular rate constants.
Somecommentson notation: Transition state theory (TST) refers
conventional, generalized, or variational transition state theory. Whenit is
necessary to makea distinction, conventional transition state theory refers
to placing the transition state at a saddlepoint on the potential energy
surface, generalized transition state theory (GTST)refers to arbitrary
locations of the transition state, and variational transition state theory
refers to GTSTwhen the location of the transition state is determined
variationally. The optimumtransition states for microcanonical or canonical ensembles correspond to a minimumsum of states or a maximum
free
energy of activation, respectively. Microcanonicalvariational theory (#VT)
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
163
and canonical variational theory (CVT)denote the results obtained making
the transition-state
theory assumption at the globally best dynamical
bottleneck for a microcanonical or canonical ensemble (55-57). Improved
canonical variational theory (ICVT) refers to using #VTbelow the/~VT
energy threshold and optimizing the variational transition states for the
non-zero contributions based on a canonical ensemble truncated from
belowat the threshold energy (61, 70).
WhenTSTis comparedto gas-phase experimental results, one tests both
fundamental assumptions as well as the potential energy surface. When
classical TSTis tested against accurate classical dynamics(trajectories), one
makes the same local-equilibrium
approximation and uses the same
potential energy surface for both calculations ; hence only the no-recrossing
assumption is tested. Whenquantized TSTis tested against accurate (i.e.
converged) quantal dynamical calculations one again uses the localequilibrium assumption and the same potential energy surface in both
cases, but nowone tests not only the implicit translation of the classical norecrossing assumption to a quantum mechanical world but also the
accuracy of the incorporation of quantal effects such as tunneling into the
TSTcalculation.
CLASSICAL
THEORY
VARIATIONAL
TRANSITION
STATE
The fundamental TSTdynamical assumption of no recrossing is inherently
a classical approximation, and classical TSTcan be formulated invoking
the fundamental equilibrium and dynamical assumptions without any
ambiguity or further approximations. In classical mechanics, TSTprovides
an upper boundon the cross section or the local-equilibrium rate constant,
and this boundis the basis for classical VTST,in whichthe calculated cross
section or rate constant is minimizedwith respect to the location of the
transition state. Physically one interprets the generalized transition states
as tentative dynamical bottlenecks to the phase-space flow of trajectories
from reactants to products. The variational transition state is the best
dynamic bottleneck for an equilibrium ensemble.
For collinear atom-diatom reactions, the classical microcanonical
variational transition states are periodic trajectories that vibrate between
two equipotentials in the interaction region (123, 124, 138). Such trajectories (called pods) may be found numerically. Pechukas (123, 124)
Pollak (128, 135) have discussed the problems with generalizing the pods
treatment to reactions in three dimensions. More approximate but more
general techniques for variational!y optimizing transition states, straightforwardly applicable in any number of dimensions, involve modeling
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
164
TRUHLAR
& GARRETT
the vibrational and rotational state sums(classical phase space volumes)
generalized transition states by the usual techniques of bound-state theory
and searching numerically for the optimumtransition states (56, 109).
Before 1979, most tests of the accuracy of the TST no-recrossing
assumption were carried out for collinear H + H2 with the conventional
transition state location at the saddlepoint (26, 27, 78, 127, 166). These
studies showedthat this assumptionis exact for this system up to about 0.2
eV above the barrier for collinear reactions and in 3D the agreement is
better than 10~o up to 1 eV above the barrier. More recently, both
conventional TSTand VTSThave been compared to accurate classical
calculations for a variety of collinear atom-diatomreactions. These studies
show that VTSToften provides significant improvements in accuracy as
comparedto conventional TST(142, 167, 56, 61).
These tests of classical VTSTagainst accurate classical rate constants
have been for bimolecular reactions with a single saddlepoint. Although
more than one dynamical bottleneck can occur for single-saddlepoint
reactions because of the decrease in the bound vibrational frequency in
going from reactants toward the saddlepoint (an entropic effect), the
presence of the two dynamical bottlenecks is an energetic effect for twosaddlepoint surfaces. Garrett et al (68) tested VTSTfor a potential energy
surface with two identical saddlepoints. The second saddlepoint makes
the no-recrossing assumption less valid at the first saddlepoint. In fact,
TSTand #VToverestimate the exact classical rate constant by a factor
of two at total energies infinitesimally above the saddlepoint energy.
However,the calculations show that the #VTand conventional TSTresults
overestimate the exact classical one by only 20~ at an energy 0.1 kcal/mol
above the saddlepoint. Over a temperature range from 100 to 10,000 K,
conventional TSTrates agree with #VTand CVTones to within 10~. For
the system studied, the worst agreement between any form of TSTand the
exact classical results is for conventional TSTat high temperature; for
exampleat 2400 K conventional TSTis too high by 21~o and at 10,000 K it
is too high by 47%.
Wolf & Hase (194) applied minimum-state-density criteria, which are
similar to VTST,to find critical configurations for RRKM
calculations on
the dissociation of H-C-C model systems. The variational RRKM
rate
constants were larger than those computedfrom trajectories, typically by a
factor of two for the tighter transition states and by factors of 5-50 for the
looser cases. The largeness of the latter is probably due to the use of harmonic, separable approximation for the classical density of states and to
an oversimplified treatment of the hindered-rotor degrees of freedom.
J. Miller (111, see also 112) has applied classical generalized TSTto
reaction with no intrinsic barrier: H+O2.In these calculations, the
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION, STATE THEORY
165
harmonic-oscillator, rigid-rotor approximation is used to evaluate the sum
of states at the generalized transition state and the density of states of
reactants. The vibrational frequencies and momentsof inertia are obtained
from an ab initio potential energy surface. Althoughreaction cross sections
were computedfor several locations of the dividing surface, the location of
the dividing surface that gives the minimum
reaction cross section for each
energy was not found. Using the dividing surface that gives the smallest
cross sections, the generalized TSTresults severely underestimate the
quasiclassical trajectory results at all total energies below about 33
kcal/mol. However, we emphasize that these comparisons are for purely
classical generalized TSTversus trajectories with quantized initial conditions; hence no definite ordering of the resulting cross-sections should be
expected.
Martin &Raft (109) have suggested a general procedure for classical
variational transition state theory calculations in atom-diatomreactions in
three dimensions. The dividing surface is expressed as a linear combination
of internal coordinates and the coefficients in this expansionare variationally optimized to minimize the thermal rate constant. Calculations were
performed for the H + H2 and H + 12 reactions. For the H + H2 system the
variational TSTresults are within 22~ of the exact classical ones over a
temperature range 300 to 1100 K. The agreement is not as good for the
H + 12 system, in which the VTSTresult overestimates the classical trajectory rate by a factor of 2.3 at 600 K. By carrying out combined-phasespace-trajectory calculations (4, 85) at their best dividing surface, a factor
18 reduction in computer time with a decrease of a factor of four in
statistical uncertainty was realized for the H + H2system, as comparedto a
trajectory calculation with sampling in the reactants’ region.
Classical variational transition state theory has also been applied to the
calculations of capture rate constants in collisions of ions with polar
molecules. Su & Chesnavich (165) have extended earlier calculations
(32, 33) to reduce the numerical error. The systems studied corresponded
H- and H~ reacting with a variety of polar molecules. For these systems
the/~VTrate constants agree very well with classical trajectory ones.
Swamy& Hase (168) have carried out similar studies for alkali ions
recombining with H20. In these studies the agreement between classical
trajectory rate constants and #VT ones is not as favorable: for the
Li ÷ + H20system errors of 2.3 were found at 300 and 1000 K, and for the
K+ +H20 system errors of 2.9 and 6 were found at 300 and 1000 K,
respectively. The errors are the result of trajectories that form short-lived
collision complexesthat are not stabilized by a third-body collision, thus
leading to recrossing of the dividing surface.
Onedifficulty in calculating reliable thermal rate constants is the lack of
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
166
TRUHLAR & GARRETT
potential energy surface information. In VTSTthe necessary information is
the potential energy in a region about the minimum-energypath, whereas
in conventional TST only information about the potential near the
saddlepoint and in the reactant region is necessary. Truhlar et al (182) have
developed methods of interpolating parameters in the reaction-path
Hamiltonian between reactants, the saddlepoint, and products. They
compared/~VTvalculations based on interpolation to conventional TST,
to btVTusing the actual potential energy surface information, and to exact
classical rate constants for a symmetric, a nearly symmetric, and two
asymmetric collinear atom-diatom reactions. In all cases the/~VT results
computedusing the interpolated potential energy surface information are
in good agreement with the/~VTresults obtained using the actual potential.
The interpolation schemes provide useful means of obtaining improved
estimates of the rate constants for systems with limited potential energy
surface information.
VARIATIONAL TRANSITION STATE THEORY IN
THE REAL, QUANTIZED WORLD
VTSTcalculations in the quantummechanical world have been carried out
using the ansatz that if quantum effects on reaction-coordinate motion,
which is responsible for movementfrom the reactants’ region of phase space
or state space to the products, are temporarily neglected, it still makes
physical sense to minimize the rate constant (57). The intermediate-step
quantity involved in this step, a rate constant corresponding to classical
reaction-coordinate motion but a quantum mechanical treatment of all
other degrees of freedom, has been called the hybrid rate constant.
Minimizing the hybrid rate constant with respect to dividing-surface
location is called quantized VTST.(In a quantized calculation there is
usually not a large difference between the results of/~VT, CVT,and ICVT
calculations; in such cases we can just say VTST.) Quantal effects on
reaction-coordinate motion and to some extent even quantal nonseparability of the reaction coordinate can be included, if desired, by multiplying
the hybrid rate constant by a transmission coefficient. This~ generally
includes both classically forbidden barrier penetration and nonclassical
barrier reflection, but since the former usually dominatesit is convenientto
call this a tunneling correction.
Microcanonical variational theory for the hybrid rate constant is
equivalent to makingan adiabatic approximation for all degrees of freedom
with respect to the reaction coordinate (56, 57). Thus there is a strong
connection between VTSTand adiabatic collision theories. By use of the
adiabatic analogy or diabatie generalizations, the dynamicbottlenecks of
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
167
TSTcan also sometimesbe interpreted as dynamical bottlenecks for stateselected reactions or for the decay of quasibound collisional resonance
states.
The most important quantummechanical effect on reaction-coordinate
motion is tunneling. Thus the ability to estimate tunneling probabilities
accurately is essential to the accurate use of transition state theory for many
reactions. In general the tunneling contribution may be estimated by any
semielassieal or quantal method; in some eases the adiabatic approximation mentioned above in conjunction with the classical-reactioncoordinate motion part of the calculation is also useful for the tunneling
calculations, and the adiabatic derivation of TSTmakes it clear howto
include tunneling consistently (183, 70). Quantized VTSTwith semiclassical adiabatic transmission coefficients based on the #round-state
s-wave reaction probability is abbreviated VTST/SAG.Two kinds of
nonadiabatic transmission coefficients have also been applied as corrections to quantized VTST;these have been called the large-curvature
#round-state and /east-action #round-state methods, and they lead to
results abbreviated VTST/LCGand VTST/LAG.
Although it is not a necessary part of VTST,in our ownwork we have
always considered one-dimensional sequences of generalized-transitionstate dividing surfaces orthogonal to a gradient-following-path in massscaled coordinates. This choice of dividing surfaces is convenient; it
eliminates potential coupling between the reaction coordinate and the
other degrees of freedom through quadratic terms, and it promotes the
dynamicseparability of the reaction coordinate, thus tending to minimize
local recrossing effects. Furthermorethe use of a gradient-following path in
mass-scaled coordinates facilitates the inclusion of internal centrifugal
effects in tunneling calculations. An excellent discussion of gradientfollowing paths and the structure of the Hamiltonianin coordinate systems
¯ built on such paths has been given by Natanson(120).
Practical VTSTcalculations for a quantized world have so far been based
on variationally
optimizing the hybrid rate constant and adding a
tunneling correction (70, 181) because more rigorous extensions of VTST
to a quantum mechanical world do not provide a useful bound [see the
discussions in (123, 172, 43)-I. Pollak (131) has presented a newtransition
state expression with boundingproperties and discussed its expansion in a
power series in h. The h expansion is knownto be slowly convergent for
tunneling contributions. It would be interesting to see whether practical
and accurate bounds for real chemical reactions could be obtained from
this formulation or whether the formalism provides a practically advantageous way to choose variational dividing surfaces.
Garrett et al (71) tested VTST/SAG
calculations for model collinear
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
168
TRUHLAR & GARRETT
reactions with mass combinations H + FH and D + FD and with a lowbarrier, twin-saddlepoint potential energy surface by comparison against
accurate quantal dynamical calculations. The comparisons showed that
VTST/SAG
predicts accurate local-equilibrium rate constants within a
factor of 1.57 over the 200-7000 K range for the H + FHmass combination
and within a factor of 1.42 over the 20(~2400K range for D + FD, with the
largest errors at the lowest temperature in each case. Employment
of fully
quantal vibrationally adiabatic tunneling calculations rather than semiclassical ones improved the accuracy. In another series of tests of
VTST/SAG
calculations against accurate quantal collinear rates, Bondi
et al (16) considered the reactions H+H2, Mu+H2, and Mu+D2
200-2400 K, using two different potential energy surfaces for two 0f the
mass combinations, for a total of five cases. For H + H2 and Mu+ H2 the
accuracy was 38~ or better over the whole temperature range, but for
Mud- 02 the errors were in this range only for T ~ 500 K or 800 K, depending on the surface. The accuracy of VTST/SAG
calculations for the seven
systems discussed in this paragraphis actually slightly worsethan typical of
that found in 18 previous test cases of VTST/SAG
calculations against
accurate quantal equilibrium rate constants for collinear systems, as
reviewedpreviously (61, 181, 177). In fact, for 300 K, for the full set of
cases the ratio to the accurate quantal equilibrium rate constant of the rate
constant calculated by VTST/SAG
calculations employing CVTfor the
VTSTpart and the small-curvature-tunneling approximation (163) for the
SAGpart, is in the range 0.49 to 1.54 in 22 cases and in the range 0.62 to 1.30
in 17 cases. In contrast, conventional transition state theory calculations
often show large errors in these 25 cases: the ratio of conventional
transition state theory rate constants to accurate quantal equilibrium rate
constants is in the above two ranges in only seven and five cases,
respectively, and even extending the range to 0.4 to 2.5 increases the number
of cases to only 13.
For Mu+ D2 the VTST/SAG
calculation on the most accurate surface
decreases the er.rors from factors of 59, 23, and 1.8 in conventional TSTat
200 K, 300 K, and 2400K to 0.42, 0.49, and 0.92, respectively. Theseresults,
like all VTST/SAG
calculations mentionedso far, are based on straight-line
GTSTdividing surfaces with anharmonicity treated by a Morse approximation [the MorseI approximation in the notation of Garrett & Truhlar (56,
57)]. Pollak (138) has shownthat more accurate results can be obtained
the Mu+D
2 mass combination by semiclassical quantization of pods.
Straight-line dividing surfaces for collinear reactions have more straightforward three-dimensional and polyatomic analogs than do pods, so we
prefer this treatment to one based on pods ; but it is not too impractical to
go beyond the Morse approximation for anharmonicity. If the anharmonic
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
169
energy levels of the straight-line generalized transition states are calculated
numerically by the WKBapproximation, without the Morse approximation, VTST/SAG
agrees with quantal equilibrium rate constants for
Mu+ D2 on the most accurate surface within 8~ over the 200-2400 K
temperature range (Garrett &Truhlar 1984, J. Chem.Phys. In press); it is
encouraging that it is not necessary to use the curved pods as dividing
surfaces to achieve this accuracy. Further work (Garrett &Truhlar 1984, J.
Chem. Phys. In press) shows that although the Morse approximation
usually leads to reasonably good agreement with the WKB
approximation
for the zero point energy at the variational transition state (quantitative
differences are largest for reactions with high zero point energies), using the
WKB
method for stretching vibrations does provide systematic improvement over that achieved in previously reported tests of VTST/SAG
calculations against accurate quantal dynamics.The rest of the VTST-plustunneling results discussed in the present section were all obtained by the
more convenient Morse approximation for stretches and by a mixed
quadratic-quartic approximation (62, 84) for bends that have no cubic
anharmonicity.
Bondi et al (16) also predicted three-dimensional rate constants for the
Mu+ Hz reaction based on the most accurate available potential energy
surface, the so-called LSTH[Liu (!00), Siegbahn & Liu (157), Truhlar
Horowitz (178)] surface. At 600 K, the ratio of the VTSTrate to the TST
one is 0.11, the ratio of the VTST/SAG
rate to one calculated (15) from full
quasiclassical trajectory calculations is 0.065, and the kinetic isotope effect
(ratio of the rate for Mu+ H2 to that for H + H2) is 0.017. The small values
for all three ratios are direct consequencesof the large zero point effects for
this system and the large dependenceof the ground-state stretching energy
level of the generalized transition state on the value of the reaction
coordinate. Despite the size of this effect, the predicted rate constants at
608 K and 875 K are in good agreement with later experimental values
[D. M. Garner and D. G. Fleming 1982, unpublished; cited in (15)].
These calculations involve no empirical elements or adjustable parameters
and they are believed to be the first totally ab initio reliable, quantitative
predictions of chemical reaction rates. With our present confidence in the
reliability of our dynamical calculations for a given potential energy
surface, the difficulty of making such predictions for other reactions
depends more on future advances in electronic-structure calculations of
potential energy surfaces than on further advancesin treating the dynamics.
Blais et al (13-15), again using the LSTHsurface, compared VTST,
VTST/SAG,
and quasiclassical trajectory calculations of rate constants
and activation energies for H + Hz, D + Hz, and Mu+ Hz to each other at
444-2400 K and to experiment at 444-875 K. The VTST/SAG
calculations
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
170
TRUHLAR & GARRETT
are in best agreement with experiment at the lower temperatures because of
the importance of tunneling. The most interesting aspect of these calculations is the temperature dependenceof the activation energy. For example
the
calculations for D + H
2 activation energy predicted by the VTST/SAG
rises from 6.8 kcal/mol at 300 K to 7.5, 9.0, and 14.0 kcal/mol at 444, 875,
and 2400 K. The latter two values, at temperatures at which tunneling
effects are less important, are in goodagreement with trajectory values of
8.7 and 13.9 kcal/mol. Since recrossing errors increase with temperature in
classical tests for this kind of system, the agreementof variational transition
state theory with full trajectory calculations for the slope of the rate
constant at 2400 K is encouraging, especially for using the simpler theory
for the important practical problemof extrapolating rate constants to high
temperature for combustion applications.
Clary (38) tested VTST/SAGcalculations (72) against presumably
accurate quantal calculations (35, 37, 38, see also 40) for t.he threedimensional D + C1Hexchange reaction and three isotopic .analogs, with
the samepotential energy surface used for both sets of calculations so that
the comparison provided a test of the dynamical methods. The quantal
calculations were performed by a method (35, 36) combining the energy
sudden and centrifugal sudden approximations in a way particularly
appropriate for the transfer of a heavy particle betweentwo light ones. The
VTST/SAG
and quantal rate constants for 295 K differed by only 15~o, 5~,
25~o, and 12~ for the four cases studied. Good(but not as good) agreement,
an error of 38~ for T>~300 K, had also been obtained for the only previous
test (70, 181), for H + H2, of VTST/SAG
calculations against presumably
accurate three-dimensional rate constants (156) for a given potential energy
surface. Note that for H + H2 no sudden approximations were made in the
quantal calculations; the accurate calculations are possible in this case
because of.the lightness of all three atoms, yielding a relatively small
number of channels. In the H + H2 case, as for the H + FHcase discussed
above, the agreement is congiderably improved if fully quantal onedimensional tunneling calculations (60; see also 124, 173) are substituted
for the semiclassical tunneling calculations. Using such fully quantal
tunneling calculations, VTST-plus-tunneling results have recently been
reported (176) for eight tritium-substituted analogs of the H + 2 reaction,
using the accurate LSTHpotential energy surface that was also used for the
Mu+ H2 calculations discussed above.
Clary et al (41) tested VTST/SAG
calculations against accurate quantal
calculations
for collinear
H+BrHand D+BrHand against energysudden-approximation, centrifugal-sudden-approximation
calculations
for the same reactions in three dimensions. The VTST/SAG
and quantal
rate constants showed good agreement in all four cases; e.g. for three-
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITION
STATE
THEORY
171
dimensional H + BrH they agree to 20~o or better for 150-500 K, even
though the SAGtransmission coefficient is 1050 at 150 K.
Garrett et al (72, 73) applied VTST/SAGcalculations
to threedimensional kinetic isotope effects in the reactions C1 + H2, D2, T2, HD,
DH, HT, and TH. They considered eleven different potential energy
surfaces with a goal of finding a surface that was consistent with experiment.
They found large differences from conventional TST in many cases,
especially for intramolecular HD/DH
kinetic isotope effects. None of the
VTST/SAG
calculations was in completely satisfactory agreement with
experiment, perhaps because of errors in all the surfaces but also perhaps
because of remaining uncertainties in the tunneling calculations or other
errors in the dynamicscalculations, such as different amountsof recrossing
for the different isotopic combinations.
Isaacson & Truhlar (84) extended the VTSTformulation to general
nonlinear polyatomic reactions and Skodje e~ al (162) similarly extended
the SAGtransmission-coefficient
approximation. Both extensions make
use of the polyatomic reaction-path Hamiltonian of Miller et al (118) and
assume independent generalized normal modes. Extensions of these
methodshave also been presented (180) for three-dimensional polyatomic
reactions with linear generalized transition states. Isaacson &Truhlar (84)
and Truhlar et al (181) applied the general polyatomic formalism to the
reaction OH+ Hz ~ H20 + H, as well as to reactions of OHwith D2, HD,
and DH,using Schatz &Elgersma’s (155) fit of Walch& Dunning’s(186)
initio potential energy surface. They found that variational optimization of
the generalized transition state loweredthe calculated rate by a factor of 1.9
at 298 K, and quantal effects on reaction-coordinate motion increased it by
a factory of 17 at the sametemperature; both the optimization effect and the
tunneling correction are decreasing functions of temperature. The final
results agree with the recommendedexperimental rate constants of Cohen
& Westberg (42) within a factor of 1.7 over the 298-2400 K temperature
range, over which the rate constant varies by a factor of 2 x 103. The
calculated results are a factor of 1.6-1.7 higher than experiment(151, 42)
298 K, a factor of 0.6-0.8 lower at 600 K, and moreaccurate at 2400K. Thus
they slightly underestimate the low-temperature activation energy and
slightly overestimate the high-temperature activation energy. Nevertheless
the agreement with experiment is remarkably good. In general one hopes
that an ab initio surface is useful for force constants for boundgeneralized
normal modesbut one expects to have to adjust the ad initio barrier height
to obtain such good agreement with experiment; in this case good
agreement was obtained without adjustment. More important is the insight
that the VTST/SAG
calculation gives into the detailed dynamics. The
variational transition states occur 0.07~0.10 ao earlier along the reaction
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
172
TRUHLAR & GARRETT
path (in scaled coordinates with a reduced mass of 1.8 amu) than the
saddlepoint, curvature of the reaction patl~ increases the SAGtransmission
coefficient by a factor of 4.3 at 298 K, and the results are sensitive to
anharmonicity. The HE/D2 kinetic, isotope effect is larger than the
experimental one (151) at 298-600 K; we mayspeculate that this and the
too-low activation energy at these temperatures occur because the barrier
on the potential energy surface is slightly too thin and allows a little too
muchtunneling, although several other explanatibns could also be given.
Truhlar et al (174) used a combination of VTST/SAG
and trajectory
calculations to adjust a newpotential energy surface for F + H2to a variety
of
2. experimental data for the reactions F + H2 and F + D
Garrett et al (67) presented detailed studies of the trends in variational
transition state locations for atom-diatomreactions with a variety of mass
combinations and potential energy surfaces, complementingearlier systematic studies (58, 59) of this subject that were limited to rotated-Morse.
bond-energy-bond-order surfaces. The main conclusions of these studies
are as follows. The ratio k~/k vTsT of rate constants calculated by the
conventional and variational theories is largest for symmetric or nearly
symmetric reactions in which a light particle is transferred between two
heavier ones. In these cases the saddlepoint tends to be symmetricor nearly
symmetric and small changes in geometry can cause large changes in the
zero point energy requirement for a boundstretching coordinate. (There is
only one such coordinate for atom-diatom collisions; for polyatomic
reactions the analogous stretching coordinate is the one involving atoms
participating in the bondchanging.) The effect on k~/kws’r van be very large
(up to several orders of magnitude) and decreases with temperature.
second important case is very asymmetric reactions with saddlepoints
located well into the reactant or product channel. In these cases the bending
effects becomedominant. Since potential energy varies more slowly with
distance along the reaction coordinate, in these cases the variational
transition states may be muchfarther removedfrom the saddlepoint and
more temperature dependent. The ratio k~/k ws~ increases with temperature in these cases but is usually only a factor of two to three.
In order to provide further evidence that the large k~/kv’rsT; ratio for
symmetric heavy-light-heavy systems are not artifacts of the potential
energy surfaces considered, Garrett et al (74) performed VTSTcalculations
for an ab initio potential energy surface for the reactions 37C1-I-H35C1 and
37C1 d- D35C1..For three-dimensional 37C1 -I- H35C1
they obtained v’rsT
k:~/k
values of 110, 28, and 9 at 200 K, 300 K, and 600 K, confirming the large
effect. The saddlepoint on the ab initio surface is symmetricwith nearestneighbor distances at 1.47 A, potential energy 6.3 kcal/mol, and a bound
stretching frequency of 337 cm- 1. At the 300 K twin asymmetricvariational
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
173
transition states, these values are 1.60 A, 1.35 A, 5.8 kcal/mol, and 1682
cm1. Tunnelingcalculations based on vibrational adiabaticity along the
minimum-energy
path are not valid for these mass combinationsbecause
the minimum-energy
reaction path has large curvature in mass-scaled
coordinates. Instead, the authors used a newlarge-curvature tunneling
method,leading to the LCGtransmission coefficient mentionedabove. In
order to compareto experiment(92, 93), the surface wasscaled along the
minimum-energy
path (but not orthogonal to it, see the discussion above
for the OH÷ H2reaction) so that the VTST/LCG
rate constant for 37C1
÷D35C1 agreed with experiment at 368.2 K; this yielded a collinear
saddlepoint potential energyof 9.0 kcal/mol and a noncollinear saddlepoint potential energyabout1.5 kcal/mollower.Thekinetic isotope effects
calculated for the scaled surface, as calculated by the conventionalTST,
VTST,and VTST/LCG
methods, are comparedto experiment (92) in Table
1. Althoughboth the conventionalTSTresults and the VTST/LCG
results
are in qualitative agreementwith experiment,the physical factor controlling the kinetic isotopeeffect is entirely differentin the twocalculations.In
conventional TSTthis kind of isotope effect is determined by the
saddlepoint stretching force constants, as in the widelyused MelanderWestheimer
model(110, 190). In the VTSTcalculations without tunneling,
the kinetic isotopeeffect at the temperatures
of Table1 is less than 1.03, so
essentially the entire effect in the VTST/LCG
calculations is dueto quantal
effects on reaction coordinate motion.Furthermore,in the LCGmodelthe
tunneling for these reactions occursby rapid light-atom motionsat fixed
C1-C1distance, andmostof it occursfor C1-C1distances muchlarger than
the C1-C1distance at the saddlepoint or evenat the outer turning-point
distance of the C1-C1symmetric-stretchzero-point motionof the conventional transition state. Theseresults cast strongdoubtsonthe validity of the
common
pra.ctice in physical organic chemistryof interpreting this kind of
isotope effect for H or H+ transfer in terms of transition state force
constants. In a moregeneral context, our VTSTcalculations call for a
critical reexamination
of conventionalTSTinterpretations (110) of kinetic
isotope effects evenin cases whentunnelingeffects on the kinetic isotope
Table 1 H/D kinetic isotope effects
scaled ab initio surface
for 35C1+H37C1on
T(K)
*
CVT
CVT/LCG
Experimental
368
423
2.8
2.5
1.0
1.0
4.2
3.4
5.0±0.7
4.1+0.4
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
174
TRUHLAR & GARRETT
effect are small because we find in manycases that variational transition
states for different isotopic versions of a given reaction are different whereas
the basic principle of the conventional analysis is that they are not.
Bondi et al (17) tested the VTST/LCG
method against accurate quantal
equilibrium rate constants for collinear CI+HC1, CI+DCI, and C1
+ MuC1,using not the ab initio or scaled ab initio surface but a similar
semiempirical surface with a collinear saddlepoint potential energy of 8.55
kcal/mol. The comparison for the C1 + MuC1case is given in Table 2. Both
the variational effect, as measuredby k*/k vvsr, and the quantal effect on
wsT, are very large,
reaction-coordinate motion, as measuredby kVTSWrCO/k
but the final results are accurate within 39~o over a factor of five in
temperature. The success of the VTST/LCGmethod in this case is a
consequenceof the success of the tunneling calculations. To verify that the
VTSTpart of the calculation is also meaningful, Truhlar, Garrett, Hipes &
Kuppermann(1984, J. Chem. Phys. In press) tested the same methods
against accurate quantal equilibrium rate constants for the reaction I + HI
on a low-barrier surface for which tunneling effects are negligible. The
results are shownin Table 2, and they verify that the VTSTand VTST/LCG
methodsare reliable for heavy-light-heavy reactions in the low-barrier, notunneling limit as well as the high-barrier, tunneling-dominatedlimit.
So far in this article we have considered primarily tight transition states
in which two bonds are simultaneously appreciably partially broken or
newlymade. Variational transition state theory is also applicable to loose
and nearly loose transition states, and we nowconsider recent papers on
that subject.
Cates et al (24) considered the reactions ÷ +H2 ~ HC1+ +Hand
HC1÷ +H2 --, H2C1+ +H. Both reactions are exoergic but the authors
found a positive temperature dependence for the former and a negative
Table 2 Ratio of approximate rate constants to accurate
quantal equilibrium ones for eollinear reactions on semiempirical
surfaces
Reaction
T (K)
:~
ICVT
ICVT/LCG
C1 + MuC1
200
400
1000
100
200
400
1000
92300
581
44
17500
214
96
19
0.003
0.081
0.56
0.77
0.99
1.1
1.4
0.68
1.02
1.39
0.77
0.99
1.1
1.4
I+HI
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
175
temperature dependencefor the latter. Theyinterpreted this in terms of an
early barrier for the former and, following Farneth & Brauman (51),
Olmstead & Brauman(122), Asubioj o &Brauman(5), Jasinski & Brauman
(88), and Pellerite & Brauman(129), in terms of a tight variational
transition state for the latter. It wouldbe interesting to see whether the
latter interpretation could be supported by actual VTSTcalculations on a
full potential energy surface.
Troe (170) provided a simplified version of the statistical adiabatic
channel model (145), which is similar to VTST, for unimolecular bond
fission reactions and the reverse radical association schemes. This work
addresses the difficult question of the correlation of vibrational, rotational,
and orbital (centrifugal) energies betweenthe two limits of tight and loose
generalized transition states. The properties of the potential energy surface
are interpolated by a schemesimilar to that originally applied by Quack&
Troe (145).
In the section on classical VTSTwe discussed the work of J. Miller (111)
on H+O2--, HO+O.This is an endoergic reaction whose dynamical
bottleneck lies in the exit channel ; it is equivalent but morestraightforward
to consider the early generalized transition states of the reaction O + OH.
Rai & Truhlar (150) applied quantized VTSTto this case using the ab initio
potential energy surface of Melius & Blint (201), and the semiempirical
reaction-path correlation schemeof Quack&Troe (148). For the Melius-Blint
surface the variational transition state for 300 K occurs at an O-to-OH
distance of 5.4 A, which is muchlarger than the range of O-to-OHdistances
for which most of the electronic structure calculations were performed. One
advantage of VTSTcalculations over collision theory calculations is that
this kind of information about critical geometries is available and mayserve
as a guide to future electronic structure calculations so that they maybe
carried out at the dynamically most important geometries. For 300 K the
VTSTrate constants, as well as the trajectory calculations of Miller (111),
are larger than the experimental rate constants (42), presumably because
the ab initio surface is too attractive. It is not clear whetherthis is a fault of
the electronic structure calculations or the extrapolation to large O-to-OH
distances. The calculations based on the Quack-Troe scheme were more
successful, but it is not knownwhether this is fortuitous or meaningful,
especially since the calculations are sensitive to howthe rotational-orbitalmotioncorrelations are treated, and this is quite uncertain.
Clearly further progress on the transition state theory of systems with
loose and nearly loose transition states will require better knowledgeof
potential energy surfaces for such systems. Duchovicet al (45) have recently
performed state-of-the-art electronic structure calculations for the poten-
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
176
TRUHLAR & GARRETT
tial energy along the dissociation coordinate in CH,~--, CH3 q- H. Further
work along this line is sorely needed.
VTSTconcepts have also been applied to a few reactions involving more
than four atoms. See the work of Brauman and co-workers mentioned
above and also Agmon(1, 2), Chesnavichet al (29), and Jarrold et al (86).
Bowers and co-workers proposed a transition-state
switching model for
ion-molecule reactions involving tight and orbiting generalized transition
states. In this modelthe existence of tight generalized transition states is
postulated even for reactions without a saddlepoint; for such reactions
there may be a local maximum
in the free-energy of activation for tight
geometries because, as the system movesalong the reaction coordinate in
the exoergic .direction, rotational-orbital motions of the reactants are
converted to vibrations. Such dynamical bottlenecks were found by Garrett
&Truhlar (58) for neutral reactions with very small barrier heights in the
entrance channel and by Rai & Truhlar (150) for the no-saddlepoint
O+OHradical-radical
reaction discussed above. As the temperature
increases, the canonical variational transition state becomestighter in such
systems; in somecases there maybe a tighter and a looser bottleneck even
at a single temperature.
The central barriers in long-lived ion-molecule complexes have been
further characterized by Wolfeet al (195) and Squires et al (164).
In attempting to use VTSTconcepts in a qualitative sense one should be
careful to distinguish free energies of activation from free energies of
formation. Thus, as the generalized transition state tends to reactants or
products, it should not be assumedthat the free energy of activation tends to
zero and to the free energy of reaction, respectively. The difference arises
because the free energy of activation is a quasithermodynamic quantity
referring to transition states, which are missing one degree of freedom,
whereas free energies of reaction and formation include all degrees of
freedom.
RELATED
TOPICS
Abovewe have reviewed recent developments in transition state theory. We
nowbriefly consider recent developmentsin a few closely related subjects.
Wedo not attempt to present self-contained discussions of these subjects in
their owncontext but rather discuss them in relation to VTSTconcepts.
Related
Dynamical
Approximations
The unified statistical (US) theory provides a generalization of VTSTto the
case of two (113) or more (114) dynamical bottlenecks. Unlike VTST,
unified statistical theory does not give a boundeven in classical mechanics.
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITION
STATE
THEORY
177
Pollak &Levine [(140, 141), see also Davis (44)] have emphasizedthat
UStheory for classical systems can be derived by information theory where
the average number of crossings of a critical surface is imposed as a
constraint, and they have also proposed a generalization involving a second
constraint, which maybe computedfrom the entropies of the reactive and
nonreactive state-to-state probability matrices. They(141) also suggested
canonical generalization by replacing microcanonical fluxes by canonical
ones. A canonical unified statistical theory was also suggested by Garrett &
Truhlar (63), whotested its predictions against accurate classical dynamics
for several collinear atom-transfer reactions. Garrett et al (68) tested the
original classical unified statistical theory against accurate dynamicsfor a
reaction with two saddlepoints and found that it overestimates the extent
of recrossing and hence underestimatesthe rate constant. Truhlar et al (180)
discussed the incorporation of quantization and tunneling effects in the
original and canonical unified statistical theories. Theyalso reported that
the quantized canonical unified statistical theory does not systematically
improve on the canonical variational theory in accuracy tests against
accurate quantal equilibrium rate constants for collinear reactions, although it can change predicted kinetic isotope effects by a non-negligible
amount. In the limit of a strongly bound intermediate between the
dynamicalbottlenecks, the unified statistical theory reduces to the statistical theory of Pechukas&Light (126) and Nikitin (121). That statistical
theory was originally formulated f~or loose dynamical bottlenecks so
that the fluxes through the dividing surfaces were proportional to the
asymptotically available phase space, but it was generalized to tight
generalized transition states by Lin &Light (99). Webb&Chesnavich(189)
have used models involving both tight and orbiting transition states in
generalized statistical
phase-space theory calculations on the energy
dependenceof the cross sections for the reaction C+ +
2. D
Chesnavich(28) proposed a theory related to VTSTin which, rather than
varying the dividing surface location, he fixed its location in the entrance
channel and varied its boundary. He obtained upper bounds on cross
sections for atom-diatom exchangereactions.
Animportant remaining problem in generalized transition state theory is
to estimate recrossing corrections. It would be very convenient if these
could be estimated from local properties of the potential surface. Miller
(114a) attempted to do this using the curvature of the minimum-energy
reaction path at the saddlepoint or the point of maximumcurvature;
unfortunately, as discussed elsewhere(180), the predictions of his formulas
do not correlate well with accurate classical dynamics. Global trajectories
provide a more reliable, but more expensive, guide to recrossing effects.
Bowman
et al (19) have evaluated transmission coefficients from trajec-
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
178
TRUHLAR
& GARRETT
tories starting in asymptotic regions, and Truhlar & Garrett I-as discussed
in (180, 181)] have evaluated them from trajectories beginning at variational transition states. Further work using this approach is in progress.
Lee, Bowman, and colleagues (18, 97) suggested using reduceddimensionality accurate quantal calculations to obtain transmission
coefficients for full-dimensional TSTcalculations. In further work, Walker
& Hayes (187) and Bowmanet al (21) presented reduced-dimensionality
calculations for reaction of H with vibrationally excited H2. The reduction
in dimensionality was achieved by treating bending degrees of freedom
adiabatically, and it corresponds to using generalized transition state
theory for bending and rotational degrees of freedom and full dynamicsfor
the two most strongly coupled degrees of freedom. Miller &Schwartz (119)
and Skodje & Truhlar (161) have presented improved system-bath
decompositions of reaction-path Hamiltonians that might be used for this
kind of approximation.
Kuppermann(94), Christov (34), and Truhlar et al (180) have provided
discussions of the relation of transition state theory to accurate collision
theory. The goal of this kind of analysis is to provide further insight into the
dynamical corrections to TST, such as those considered in the previous
paragraphs.
In microcanonical transition state theory one calculates a rate constant
for each total energy of the transition state. Miller (115-117) has pointed
out that one should calculate a distinct microcanonical rate constant for
each irreducible representation of the transition state in the symmetry
group that applies along the reaction path, since states of different
symmetryare decoupled. The difference between the rate constants for
different symmetries is largest for energies near threshold. A related
practical point is that it is sometimesbetter to base transition state theory
on a reference path through a saddlepoint with two imaginary frequencies.
This kind of reference path has been used for the unimolecular decomposition of H2CO
(116) and for 37C1q-H35C1(74). Celli et al (202)
Sakimoto (203) have calculated ion-dipole capture rate constants using
average-free-energy-function method and an adiabatic method, respectively; both methodsare closely related to/~VT(31).
Tunneling
As discussed above, accurate transmission coefficients that account for
tunneling contributions are an important ingredient in transition state
theory calculations for many cases. Wehave already mentioned some
aspects of new developmentsin the theory of tunneling in conjunction with
variational transition state theory calculations.
The most significant qualitative points to emerge from recent work on
tunneling in chemical reactions are: 1. Thc tunneling contributions are
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITION
STATETHEORY
179
usually larger than would bc expected by most workers. 2. Reaction-path
curvatureeffects are often very large, i.e., accurate transmissioncoefficients
can be calculated only by using dominant tunneling paths systematically
displaced from the minimum-energyreaction path. 3. Tunneling probabilities for multidimensional systems can nevertheless bc calculated
reliably in most or all cases by rcduccd-dimensionality semi-classical
methods.
The simplest and most commonly used methods for approximating
tunneling effects in conventional TSTare the methodsof Wigner(191) and
Bell (10). In both these methods the potential along the minimumenergy
path in the vicinity of the saddlepoint is approximated as a truncated
parabola. The transmission coefficient is obtained by BoltzmannTaveraging
the scmiclassical barrier penetration probabilities. Wigncr’s tunneling
correction is a semiclassical approximation to lowest order in h. This
correction factor is valid only whenthe correction is small, typically less
than a factor of 2. Bell’s methodhas a larger region of validity but the
expressions arc discontinuous and contain divergences. Rcccntly, Skodjc &
Truhlar (160) have presented a continuous, divcrgcncc-frcc analytic
expression for the transmission coefficient for a truncated parabolic barrier
that approximates the accurate uniform semiclassical transmission coefficients over a wide range of parameters (see also 11). The methodis also
applicable to unsymmetricbarriers and is shownto bc useful for barriers
with shapes other than parabolic. They also found that it is best to fit the
barrier to a parabola using the effective parabolic width of the nonparabolic
barrier at energies that contribute appreciably to the transmission
coefficient.
The first successful approximation for tunneling in systems with
significant reaction-path curvaturc was dcvelopcd by Marcus & Coltrin
(108) and extended by Garrctt & Truhlar (54, 57, 60, 62). This method
calculates the tunneling action integral along the caustic envelope of a
family of unboundtrajectories with quantized adiabatic vibrations; this is
called the Marcus-Coltrin path (MCP). More recently, Gray and coworkers (76) developed a semiclassical adiabatic modelusing the reactionpath Hamiltonian (118) and treating the kinetic energy terms containing
curvature coupling by second-order classical perturbation theory. They
used this second-order (SO) tunneling method involving the adiabatic
barrier to study the unimolecular isomerization of HNCto HCN(76) and
the unimolecular decomposition of formaldehyde (77). Forst (52)
treated tunneling in formaldehyde decomposition; however, he used the
classical barrier and neglected reaction-path curvature. Ccrjan ct al (25)
unified the semiclassical perturbation approximationwith the infinite order
sudden approximation applied to the reaction-path Hamiltonian to obtain
an expressionfor the total reaction probability that takes the formof a zero-
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
180
TRUHLAR & GARRETT
curvature adiabatic reaction probability times a curvature-dependent
correction factor.
Skodje et al (162, 163) developeda semiclassical adiabatic modelthat
valid for systems with small reaction-path curvature, and they derived a
criterion for the validity of the adiabatic approximation in curvilinear
natural collision coordinates. The small-curvature (SC) tunneling method
is similar to the collinear-reaction methodof Marcus&Coltrin (108) but
can be applied without singularities to systems with large reaction-path
curvature, and it is expected to be more accurate for small-curvature
systems. In addition it is applicable to noncollinear systems with reactionpath curvature components in more than one degree of freedom, for
exampleit has been applied to calculate large tunneling corrections for the
reactions OH+H2and isotopic analogs, which have curvature components in four of the five vibrational coordinates (84, 181). Skodjeet al also
compared, both formally and numerically, the MCP,SC, and SOtunneling
methods, a previously suggested method [the phase average (PA) method
(118)], and three new methods [the vibrational average (VA),
dynamical-path vibrational average (DA), and the semiclassical optical
potential (SOP) methods]. These adiabatic methodsmay be classified into
two general groups, depending upon the method used to remove from the
kinetic energy term the dependence upon the coordinates of the bound
degrees of freedom orthogonal to the reaction coordinate. One class of
models (PA, SO, VA, DA, and SOP) accomplishes this by "averaging" the
reaction-path Hamiltonian over the vibrational coordinates either classically or quantally. The other class (MCPand SC), which is systematically
moresuccessful, defines single values of the vibrational coordinates for each
value along the reaction coordinate. These "vibrational-collapse" models
have the physical interpretation that the tunneling is forced to occur along a
specified path through the interaction region.
For reactions with large reaction-path curvature, the adiabatic approximation breaks down. Large reaction-path curvature occurs, for
example, in systems in which a light atom is transferred between two heavy
atoms or molecules. Babamov, Marcus, and Lopez (6-8) developed
methodfor computingthe reaction probability for this type of system, and
they applied it to study tunneling probabilities in the threshold region as
well as the oscillations of the reaction probability as a function of energy for
energies above threshold. Garrett, and co-workers (74) developed a similar
method, which they called the large-curvature (LC) method, and they used
it to calculate thermally averaged tunneling correction factors for VTST.
The physical model for a collinear atom-diatom reaction is that the
tunneling occurs by the most direct path (a straight line) connecting the
reactant and product regions. Motion in the bound vibrational coordinate
(rather than translational motion along the reaction coordinate) promotes
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITION
STATETHEORY
181
tunneling, and for a fixed total energy, tunneling can begin at a wide range
of geometries along the caustic parallel to the reaction coordinate from the
asymptotic reactant region to the turning point in the adiabatic potential.
This method was demonstrated to work well for the C1 + HC1reaction and
isotopic variants in both 1D (17) and 3D (74).
Garrett & Truhlar (65) unified the LC method with the vibrationalcollapse adiabatic models by developing a least-action (LA) tunneling
method. In this methodthe optimumtunneling path is chosen from a set of
parameterized paths by requiring it to be the one that accumulates the
least imaginary action along the tunneling path. This methodwas found to
be extremely successful for a system with small-to-large reaction-path
curvature. In practice, transmission coefficients based on the LCand LA
approximations are based on the ground state and are called LCGand
LAG,respectively. See 080) for formulas for applying these methods to
general polyatomic systems.
The methods described above have been applied to reactions with
barriers in the regions of large reaction-path curvature, and reaction-path
curvature effects on tunneling probabilities have been found to be very
important in manycases. For reactions with no barrier, or barriers far into
the reactant and product regions, it is possible to simplify the treatment of
reaction-path curvature. Illies, Jarrold, and Bowers(83, 87) proposed
tunneling model to describe the unimolecular fragmentation of CH,~and
NH~-,which are reactions with loose transition states. They approximated
the potential in the tunneling region by a dipole term plus a rotational
barrier from free internal rotation of the molecular fragments and orbital
rotation.
Using this model they obtained good agreement between
calculated rate constants and experimental ones.
Heller & Brown(80) presented a method to estimate surface-hopping
probabilities from a boundstate on an upper surface to a boundstate on a
steep lower surface that does not cross (or avoids crossing) the upper surface
in the classically allowed region, for the case in which a single path
dominates the tunneling. Althoughthe problem is formally quite different
from the problem of single-surface reactive scattering for which the LA
method was developed, the semiclassical solution has some points in
common,especially with our small-curvature limit. Cross-fertilization of
the two methods may provide clues as to how to extend both to a wider
range of problems.
Transition state theory with tunneling has also been used to examine
intramolecular hydrogen-transfer reactions. LeRoy(98) used a phenomenological model to calculate the rate of transfer of hydrogenatoms between
two nonequivalent sites in large polyatomic molecules. The physical model
is that vibrational stretching of the bondbeing brokeninitiates the reaction,
although it is not necessary that 100~ of the energy in the vibration is
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
182
TRUHLAR
& (3ARRETT
available for promotingthe reaction. The effects of the degrees of freedom
orthogonal to the reaction coordinate enter the rate expression through a
steric factor. Adjusting the steric factor, the percentage of vibrational
energy available to the reaction coordinate, and the effective onedimensional potential along the tunneling path~ LeRoy found he could
reproduce experimental data for several intramolecular H atom transfer
reactions. Bicerano et al (12) studied a similar problem, the transfer of
hydrogen atom between equivalent sites in malonaldehyde. For this
symmetric system the potential along the minimumenergy path between
the two .equilibrium geometries is a symmetric double well potential.
Tunneling was included in the vibrationally adiabatic approximation with
potential parameters taken from ab initio electronic structure calculations.
Instead of calculating the rate of transfer from one well to another, they
computedthe effect of tunneling upon the energy level splitting, obtaining a
result within a factor of two of the experimental one. The polyatomic VTST
formalism discussed above can also be applied to multidimensional
unimolecular isomerizations, with SC, LC, or LA tunneling corrections
(180; F. B. Brownand D. G. Truhlar 1984, unpublished).
Vibrationally
Adiabatic
Barriers
The free energy of activation curve as a function of reaction coordinate
reduces at 0 K to the vibrationally-rotationally
adiabatic ground-state,
s-wave potential curve, or, for short, the vibrationally adiabatic groundstate potential curve Vf(s). Whenthe shape of this curve is dominatedby
the s dependence of high-frequency modes, then the barriers of V~(s)
provide a guide to the location of dynamical bottlenecks at nonzero
temperature or nonzero microcanonical energy. Similarly, the barriers of
vibrationally adiabatic excited-state curves mayprovide dynamical bottlenecks for reactions of vibrationally excited species.
Agmon(3) suggested using analternative coordinate system to calculate
V~(s), with the goal of improving the accuracy of the separability of the
reaction coordinate that must be assumed in TSTor VTST. A difficulty
with Agmon’scoordinate system is that the kinetic energy operator of the
generalized transition state is complicated because the vibrational coordinates are curved. Reaction-path Hamiltonians based on the minimumenergy path and non-curved vibrational coordinates (56, 57, 70, 84, 104107, 118) allow for more convenient calculations of the vibrational energies
of the generalized transition states ; yet, in a quadratic expansionabout any
point on the reaction path, the potential energy contains no cross-term
coupling the reaction coordinate to the vibrational coordinates.
Pollak (130, 132-134) calculated vibrationally adiabatic potential curves
and transmission probabilities for the vibrationally excited collinear
H-t-HE reaction and isotopic analogs by quantizing pods, and also, in
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITIONSTATETHEORY
183
Jacobi coordinates, by treating the one-dimensional bound vibrational
motion quantum mechanically and the reaction-coordinate motion by a
parabolically uniformized semiclassical approximation. He obtained
qualitatively similar results to earlier state-selected vibrationally adiabatic
calculations (57), but better agreement with accurate quantal results for
reaction probabilities of vibrationally excited species. The earlier calculations had been carried out by the Morse I approximation applied to
locally straight dividing surfaces in coordinates based on the minimumenergy reaction path. The quantitative differences were attributed to the
Morse I approximation and to the neglect of important curvature
corrections that are contained in pods. Garrett &Truhlar (1984, J. Chem.
Phys. In press) have performed calculations employing straight-line
dividing surfaces perpendicular to the minimum-energy
path and using the
WKB
approximation for vibrational energies; the new calculations yield
excellent agreement with adiabatic barrier heights obtained by quantizing
pods and also with accurate quantal rate constants for the vibrationally
excited case. This showsthat curved generalized transition states are not
necessary for high accuracy. As mentionedin a previous section, the Morse
I approximation, which is very convenient, is usually adequate for thermal
reactions, but WKB
or quantal vibrational eigenvalues maybe required for
good accuracy for excited states.
Pollak (130, 132), Lee et al (96), and Ronet al (153) also used pods
vibrational energy calculations in Jacobi coordinates to evaluate adiabatic
barriers for collinear and reduced-dimensionality calculations on the
reactions F + H2 and isotopic analogues and O + Hz. A disadvantage of
Jacobi coordinates is that they yield accurate adiabatic barriers only
relatively far out in the reactants and products channels; and a disadvantage of pods is that they exist only for collinear atom-diatom
reactions. Methods based on minimum£energyreaction paths are more
general, although they may be inappropriate in regions of very large
reaction-path curvature; fortunately we have found in applications that
this is not a problembecause the variational transition state tends not to be
located in such regions.
Garrett &Truhlar (57) and Pollak (132) also used adiabatic transmission
probabilities to calculate the cumulative reaction probability, whichhas a
step-like character due to channel openings; these steps should not be
confusedwith oscillations in the state-selected reaction probabilities, which
are due to interference effects such as resonances, but may sometimesbe
explained (46) by invoking only quantal discreteness.
The main reason that quantized VTSTis more reliable than standard
trajectory calculations for thermal rate constants is that it incorporates
quantized energy requirements at dynamical bottlenecks, i.e. it incorporates the constraints of quantized adiabatic barriers. Schatz (154) suggested
Annual Reviews
www.annualreviews.org/aronline
184
TRUHLAR & GARRETT
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
incorporating such energy constraints for low-frequency, classically nonadiabatic bending modesas ad hoc additions to the potential energy surface
for three-dimensional trajectory calculations.
State-selected
Reactions
The discussion in the main part of this review is centered on thermal rate
constants, which are the traditional domainfor transition state theory. The
methodsof variational transition state theory and related methodsare also
useful for understanding excited-state reactivity in certain cases. For
example, adiabatic barriers, as discussed above, may be used to interpret
excited-state reactivity and product-state distributions (130, 133, 197). Full
rate constant calculations for vibrationally excited species may also be
performed by invoking the vibrationally adiabatic or diabatic approximation for one degree of freedomand variational transition state theory for
others; such calculations have been performed for several collinear
reactions (57, Garrett &Truhlar 1984, J. Chem.Phys. In press), for threedimensional H÷H2and D+H
2 (B. C. Garrett and D. G. Truhlar 1984,
unpublished), and for three-dimensional OH+ z (179). F or OH +2 (n
= 1), conventional TSTpredicts a vibrational rate enhancementof > 104,
whereas state-selected VTSTpredicts 102, which is in good agreement with
experiment (75, 199). State-selected VTSTcalculations (179) for OH
also imply that the large-non-Arrhenius behavior for OH+ H2 is not a
consequence of the rate enhancementfor vibrationally excited H2, as had
been suggested (198).
Pollak & Pechukas (143) showed that one may map out the reactant and
product classical vibrational energy distributions by studying trajectories
initiated in the immediatevicinity of variational transition states.
Resonances
The vibrationally adiabatic potential curves of variational transition state
theory are also very useful for predicting and classifying collisional
resonances in manychemical reactions, especially for thermoneutral and
nearly thermoneutral reactions for which reaction-path curvature is small
or intermediate (9, 64, 71, 158, 159). Variational transition states provide the
barriers to decay of the resonance in the one-dimensional vibrationally
adiabatic model. For large reaction-path curvature or strongly exothermic
reactions, approaches based on resonant periodic orbits or on adiabaticity
in hyperspherical coordinates appear more useful [see, for example, (139,
144, 152) and references therein].
Vibrational
Bondin9
Vibrational bonding has received considerable attention in the last couple
of years, and it is interesting to point out how,like resonancephenomena,it
can often be predicted and understood in terms of the same concepts and
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL
TRANSITION
STATE
THEORY
185
quantities as developedfor variational transition state theory. In particular,
vibrational bonding maybe considered as the extreme of a pre-threshold
resonance. For example, we observed pre-threshold resonances for collinear reactions with mass combinations H + FH and D ÷ FD on a lowbarrier potential energy surface, and these can be understood vibrationally
adiabatically (71, 159). If the masscombinationis changedto heavy-lightheavy, for which generalized-transition-state vibrational energy requirements show the most pronounced minimumin the interaction region (58),
the vibrational energy of the resonance will decrease, and the resonance
energy maydrop below the zero point energy of the atom-diatomreactants
and thus becomea true bound state, even though the lowest point on the
potential surface still occurs for the asymptotic atom-diatom reactants.
This is called vibrational bonding, as opposed to ordinary bonding, with an
equilibrium geometry corresponding to the minimumin the potential
energy surface. A vibrational bondingstate was first reported for collinear
IHI (101), and shortly thereafter a vibrational bonding state for threedimensional IHI was calculated (39, see also 101, 137). Variational
transition states mayserve as effective barriers that contribute to localizing
a vibrational-bonding state in the strong interaction region. Adiabatic
bondingis expected more generally in excited-state vibrationally adiabatic
curves than in ground-state ones; if the adiabatically bound state in an
excited-state vibrationally adiabatic potential curve lies below the asymptote, it maystill decaynonadiabatically(158, 159), and thus the state is only
quasibound. Only whenvibrational effects are largest does one expect to
find states belowthe asymptoteof Vff(s), and hence vibrational bondingwill
occur far less frequently than the similar resonanceeffect.
CONCLUDING
REMARKS
In the last few years it has been shownthat variational transition state
theory can be implementedusefully for practical calculations of chemical
reaction rates from potential energy surfaces. Whencombined with
accurate semiclassical tunneling calculations, VTSTis the most accurate
practical methodavailable for such calculations. Variational-transitionstate constructs are also useful for quantitative interpretations of excitedstate reactivity and resonances.
ACKNOWLEDGMENT
The work at the University of Minnesota was supported in part by the US
Departmentof Energy, Office of Basic EnergySciences, under contract no.
DE-AC02-79ER10425.The work at Chemical Dynamics Corporation was
supported by the US ArmyResearch Office under contract no. DAAG-298 l-C-0015.
Annual Reviews
www.annualreviews.org/aronline
186
TRUHLAR & GARRETT
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
Literature Cited
1. Agrnon,N. 1980. J. Am. Chem.Soc. 102 :
2164-67
2. Agmon,N. 1981. Int. J. Chem. Kinet.
13 : 333-65
3. Agmon,N. 1983. Chem. Phys. 76:20318
4. Anderson, J. B. 1973. J. Chem. Phys.
58 : 4684-92
5. Asubiojo, O. I., Brauman,J. I. 1979. J.
Am. Chem. Soc. 101 : 3715-24
6. Babamov,V. K., Lopez, V., Marcus, R.
A. 1983. J. Chem. Phys. 78:5621-28
7. Babamov,V. K., Lopez, V., Marcus, R.
A. 1983. Chem.Phys. Lett. 101 : 507-11
8. Babamov,V. K., Marcus, R. A. 1981. J.
Chem. Phys. 74:1790-1803
9. Basilevsky, M. V., Ryaboy,V. M. 1981.
Int. J. QuantumChem. 19:611-35
10. Bell, R. P. 1959. Trans. Faraday Soc.
55 : 1-4
11. Bell, R. P. 1980. The Tunnel Effect in
Chemistry, appendix C. London: Chapman & Hall
12. Bicerano,J., Schaefer, H. F. III, Miller,
W. H. 1983. J. Am. Chem. Soc. 105:
2550-53
13. Blais, N. C., Truhlar, D. G., Garrett, B.
C. 1981. J. Phys. Chem. 85 : 1094-96
14. Blais, N. C., Truhlar, D. G., Garrett, B.
C. 1982. J. Chem. Phys. 76:2768-70
15. Blais, N. C., Truhlar, D. G., Garrett, B.
C. 1983. J. Chem.Phys. 78 : 2363-67
16. Bondi, D. K., Clary, D. C., Connor,J. N.
L., Garrett, B. C., Truhlar, D. G. 1982.J.
Chem. Phys. 76 : 4986-95
17. Bondi, D. K., Connor,J. N. L., Garrett,
B. C., Truhlar, D. G. 1983. J. Chem.
Phys. 78 : 5981-89
18. Bowman,J. M., Ju, G.-Z., Lee, K. T.
1982. J. Phys. Chem.86 : 2232-39
19. Bowman,J. M., Ju, G.-Z., Lee, K. T.,
Wagner, A. F., Schatz, G. C. 1981. J.
Chem. Phys. 75 : 141-47
20. Deleted in proof
21. Bowman,J. M., Lee, K. T., Walker, R.
B. 1983. J. Chem. Phys. 79: 3742-45
22. Deleted in proof
23. Bunker, D. L., Pattengill, M. 1968. J.
Chem. Phys. 48 : 772-76
24. Cates, R. D., Bowers, M. T., Huntress,
W. T. Jr. 1981. J. Phys. Chem. 85:313
15
25. Cerjan, C. J., Shi, S.-H., Miller, W. H.
1982. J. Phys. Chem. 86:2244-51
26. Chapman,S., Hornstein, S. M., Miller,
W. H. 1975. J. Am. Chem.Soc. 97 : 89294
27. Chesnavich, W. J. 1978. Chem. Phys.
Lett. 53 : 300-3
28. Chesnavich, W. J. 1982. J. Chem. Phys.
77 : 2988-95
29. Chesnavich, W. J., Bass, L., Su, T.,
Bowers, M. T. 1981. J. Chem. Phys.
74:2228-46
30. Chesnavich, W. J., Bowers, M. T. 1979.
Statistical
methods in reaction dynamics. In Gas-Phase Ion Chemistry,
ed. M. T. Bowers, pp. 119-51. New
York : Academic
31. Chesnavich, W. J., Bowers, M. T. 1982.
Pro.q. React. Kinet. 11 : 137-267
32. Chesnavich,W. J., Su, T., Bowers,M. T.
1979. In Kinetics of Ion-Molecule
Reactions, ed. P. Ausloos, pp. 31-53.
NewYork : Plenum. 508 pp.
33. Chesnavich,W. J., Su, T., Bowers,M. T.
1980. J. Chem. Phys. 72 : 2741-55
34. Christov, S. G. 1980. J. Res. lnst. Catal.
Hokkaido Univ. 28:119-36
35. Clary, D. C. 1981. Chem. Phys. Lett.
80: 271-74
36. Clary, D. C. 1981. Mol. Phys. 44 : 106781
37. Clary, D. C. 1981. Mol. Phys. 44: 108397
38. Clary, D. C. 1982. Chem.Phys. 71 : 11725
39. Clary, D. C., Connor, J. N. L. 1983.
Chem. Phys. Lett. 94:81-84
40. Clary, D. C., Drolshagen, G. 1982. J.
Chem.Phys. 76 : 5027-33
41. Clary, D. C., Garrett, B. C., Truhlar, D.
G. 1983. J. Chem.Phys. 78 : 777-82
42. Cohen, N., Westberg, K. R. 1982.
AerospaceReport A TR-82( 7 888)-3, pp.
39-44. E1 Segundo: Aerospace Corp.
43. Costley, J., Pechukas, P. 1981. Chem.
Phys. Lett. 83 : 139-44
44. Davis, J. P. 1981. J. Chem. Phys.
75 : 2011-12. Erratum: 1982. 76 : 753
45. Duchovic, R. J., Hase, W. L., Schlegel,
H. B., Frisch, M. J., Raghavachari, K.
1982. Chem. Phys. Lett. 89:120-25
46. Duff, J. W., Truhlar, D. G. 1975. Chem.
Phys. Lett. 36:551-54
47. Eliason, M.A., Hirschfelder, J. O. 1959.
J. Chem. Phys. 30:1426-36
48. Evans, M. G. 1938. Trans. Faraday Soc.
34 : 49-57, 73
49. Eyring, H. 1935. J. Chem.Phys. 3:10715
50. Eyring, H. 1962. Discussion. In The
Transition State, Chem. Soc. Special
Publ. 16, p. 27. London: Chem.Soc.
51. Farneth, W. E., Brauman,J. I. 1976. J.
Am. Chem.Soc. 98 : 7891-98
52. Forst, W. 1983. J. Phys. Chem. 87:
4489-94
53. Deleted in proof
54. Garrett, B. C., Truhlar, D. G. 1979. J.
Phys. Ch~m.83 : 200-3 ; Erratum 83 :
3058
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
55. Garrett, B. C., Truhlar, D. G. 1979. d.
Chem. Phys. 70:1593-98
56. Garrett, B. C., Truhlar, D. G. 1979. J.
Phys. Chem. 83: 1052-79; Errata 83 :
3058, 87:4553
57. Garrett, B. C., Truhlar, D. G. 1979. J.
Phys. Chem.83 : 1079-1112; Errata 84 :
692-86, 87:4553-54
58. Garrett, B. C., Truhlar, D. G. 1979. J.
Am. Chem. Soc. 101:4534-48
59. Garrett, B. C., Truhlar, D. G. 1979. J.
Am. Chem.Soc. 101 : 5207-17
60. Garrett, B. C., Truhlar, D. G. 1979.
Proc. Natl. Acad. Sci. USA76 : 4755-59
61. Garrett, B. C., Truhlar, D. G. 1980. d.
Phys. Chem. 84:805-12
62. Garrett, B. C., Truhlar, D. G. 1980. J.
Chem.Phys. 72 : 3460-71
63. Garrett, B. C., Truhlar, D. G. 1982. J.
Chem. Phys. 76:1853-58
64. Garrett, B. C., Truhlar, D. G. 1982. J.
Phys. Chem. 86:1136-41; Erratum:
1983. 87:4554
65. Garrett, B. C., Truhlar, D. G. 1983. J.
Chem. Phys. 79 : 4931-38
66. Deleted in proof
67. Garrett, B. C., Truhlar, D. G., Grey, R.
S. 1981. Determination of the bottleneck regions of potential energy surfaces for atom transfer reactions by
variational transition state theory. In
Potential Energy Surfaces and Dynamics
Calculations, ed. D. G. Truhlar, pp.
587-637. NewYork : Plenum. 866 pp.
68. Garrett, B. C., Truhlar, D. G., Grev, R.
S. 1981. J. Phys. Chem. 85:1569-72
69. Deleted in proof
70. Garrett, B. C., Truhlar, D. G., Grev, R.
S., Magnuson, A. W. 1980. J. Phys.
Chem. 84:1730-48; Erratum: 1983.
87:4554
71. Garrett, B. C., Truhlar, D. G., Grev, R.
S., Schatz, G. C., Walker, R. B. 1981. J.
Phys. Chem. 85 : 3806-17
72. Garrett, B. C., Truhlar, D. G., Magnuson, A. W. 1981. J. Chem. Phys. 74:
1029-43
73. Garrett, B. C., Truhlar, D. G., Magnuson, A. W. 1982. d. Chem. Phys. 76:
2321-31
74. Garrett, B. C., Truhlar, D. G., Wagner,
A. F., Dunning,T. H. Jr. 1983. J. Chem.
Phys. 78:4400-13
75. Glass, G. P., Chaturvedi, B. K. 1981. J.
Chem. Phys. 75 : 2749-52
76. Gray, S. K., Miller, W. H., Yamaguchi,
Y., Schaefer, H. F. III. 1980. J. Chem.
Phys. 73 : 2733-39
77. Gray, S. K., Miller, W. H., Yamaguchi,
Y., Schaefer, H. F. IlL 1981. J. Am.
Chem.Soc. 103 : 1900-4
78. Grimmelmann,E. K., Lohr, L. L. 1977.
Chem.Phys. Lett. 48 : 487-90
187
79. Hase, W. L. 1983. Acc. Chem. Res. 16:
258-64
80. Heller, E. J., Brown,R. C. 1983.J. Chem.
Phys. 79 : 3336-51
81. Hirschfelder, J. O. 1983. Ann. Rev. Phys.
Chem. 34:1-29
82. Horiuti, J. 1938. Bull. Chem.Soc. Jpn.
13:210-16
83. Illies, A. J., Jarrold, M.F., Bowers,M.T.
1982..L Am. Chem. Soc. 104:3587-93
84. Isaacson, A. D., Truhlar, D. G. 1982. J.
Chem. Phys. 76:1380-91
85. Jaffe, R. L., Henry,J. M., Anderson,J. B.
1973. J. Chem.Phys. 59 : 1128-41
86. Jarrold, M. F., Bass, L. M., Kemper,P.
R., van Koppen, A. M., Bowers, M. T.
1983. J. Chem.Phys. 78 : 3756-66
87. Jarrold, M.F., Illies, A. J., Bowers,M.T.
1982. Chem.Phys. Lett. 92 : 653-58
88. Jasinski, J. M., Brauman,J. I. 1980. J.
Am. Chem. Soc. 102:2906-13
89. Keck, J. C. 1960. J. Chem. Phys. 32:
1035-50
90. Keck, J. C. 1962. Discuss. FaradaySoc.
33 : 173-82, 291-93
91. Keck, J. C. 1967. Adv. Chem. Phys.
13:85-121
92. Klein, F. S., Persky,A., Weston,R. E. Jr.
1964. J. Chem.Phys. 41 : 1799-1807
93. Kneba, M., Wolfrum, J. 1979. J. Phys.
Chem. 83 : 69-73
94. Kuppermann, A. 1979. J. Phys. Chem.
83 : 171-87
95. Laidler, K. J., King, M.C. 1983. J. Phys.
Chem. 87 : 2657-64
96. Lee, K. T., Bowman,J. M., Wagner, A.
F., Schatz, G. C. 1982. J. Chem. Phys.
76 : 3563-82
97. Lee, K. T., Bowman,J. M., Wagner, A.
F., Schatz, G. C. 1982. J. Chem. Phys.
76 : 3583-96
98. LeRoy, R. J. 1980. J. Phys. Chem. 84:
3508-16
99. Lin, J., Light, J. C~1966. J. Chem.Phys.
45 : 2545-59
100. Liu, B. 1973. J. Chem.Phys. 58 : 1925-37
101. Manz, J., Meyer, R., R6melt, J. 1983.
Chem.Phys. Lett. 96 : 607-12
102. Marcus, R. A. 1965. J. Chem.Phys. 43 :
1598-1605
103. Marcus, R. A. 1966. J. Chem.Phys. 45 :
2630-38
104. Marcus, R. A. 1966. J. Chem.Phys. 45 :
4493-99
105. Marcus, R. A. 1966. J. Chem.Phys. 45 :
4500-4
106. Marcus, R. A. 1968. Discuss. Faraday
Soc. 44 : 7-13
107. Marcus, R. A. 1974. Activated-complex
theory : Current status, extensions, and
applications. In Investigation of Rates
and Mechanisms of Reaction, Techniques of Chemistry, ed. E. S. Lewis,
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
188
TRUHLAR & GARRETT
6(Pt. 1): 13-46. NewYork: Wiley-Interscience
108. MarcusR. A., Coltrin, M. E. 1977. J.
Chem. Phys. 67:2609 13
109. Martin, D. L., Raft, L. M. 1982. J. Chem.
Phys. 77 : 1235-47
110. Melander, L., Saunders, W. H. Jr. 1980.
Reaction Rates of Isotopic Molecules,
pp. 29-36. NewYork : Wiley. 2nd ed.
111. Miller, J. A. 1981. J. Chem. Phys. 74:
5120-32
112. Miller, J. A. 1981. J. Chem. Phys. 75:
5349-54
113. Miller, W. H. 1976. J. Chem.Phys. 65:
2216-23
114. Miller, W. H. 1981. Reaction path
Hamiltonian for polyatomic systems:
Further developments and applications. In Potential Energy Surfaces
and Dynamics Calculations, ed. D. G.
Truhlar,
pp. 265-86. New York:
Plenum. 866 pp.
l14a. Miller, W. H. 1982. J. Chem. Phys.
76 : 4904-8
115. Miller, W. H. 1983. J. Am. Chem. Soc.
105 : 216-20
116. Miller, W. H. 1983. J. Phys. Chem. 87:
21-22
117. Miller, W. H. 1983. J. Phys. Chem. 87:
2731-33
118. Miller, W. H., Handy, N. C., Adams,J.
E. 1980. J. Chem. Phys. 72:99-112
119. Miller, W. H., Schwartz, S. 1982. J.
Chem.Phys. 77 : 2378-82
120. Natanson, G. 1982. Mol. Phys. 46:481
512
121. Nikitin, E. E. 1965. Teor. Eksp. Khim.
1 : 135-43
122. Olmstead,W. N., Brauman,J. I. 1977. J.
Am. Chem. Soc. 99 : 4219-28
123. Pechukas, P. 1976. Statistical approximations in collision
theory. In
Dynamicsof Molecular Collisions, Part
B, Modern Theoretical Chemistry, ed.
W. H. Miller, 2:269-322. NewYork:
Plenum. 380 pp.
124. Pechukas, P. 1981. Ann. Rev. Phys.
Chem. 32:159-77
125. Pechukas, P. 1982. Ber. Bunsenges.
Phys. Chem. 86:372-98
126. Pechukas,P., Light, J. C. 1965. J. Chem.
Phys. 42 : 3281-91
127. Pechukas, P., McLafferty,F. J. 1973. J.
Chem. Phys. 58 : 1622-25
128. Pechukas, P., Pollak, E. 1979. J. Chem.
Phys. 71 : 2062-67
129. Pellerite, M. J., Brauman,J. I. 1982.
Nucleophilic substitution. In Mechanistic Aspects of Inorganic Reactions,
Am. Chem. Soc. Syrup. Set. 198, ed. D.
B. Rorabacher, J. F. Endicott, pp.
81-95. Washington:Am. Chem. Soc.
130. Pollak, E. 1981. J, Chem. Phys. 74:
5586-94
131. Pollak, E. 1981. J, Chem. Phys. 74:
676%70
132. Pollak, E. 1981. J. Chem. Phys.
75:4435-40
133. Pollak, E. 1981. Chem. Phys. Lett. 80:
45-54
134. Pollak, E. 1981. Chem.Phys. 61 : 305-16
135. Pollak, E. 1982. J, Chem. Phys. 78:
1228-36
136. Pollak, E. 1982. Chem. Phys. Lett. 91:
27-33
137. Pollak, E. 1983. Chem. Phys. Lett. 94:
85 89
138. Pollak, E. 1984. Periodic orbits and the
theory of reactive scattering. In The
Theory of Chemical Reaction Dynamics,
ed. M. Baer. Boca Raton, FL: CRC
Press. In press
139. Pollak, E., Child, M. S. 1981. Chem.
Phys. 60: 23-32
140. Pollak, E., Levine, R. D. 1982. Bet.
Bunsenges. Phys. Chem. 86 : 458-64
141. Pollak, E., Levine, R. D. 1982. J. Phys.
Chem. 86:4931 37
142. Pollak, E., Pechukas, P. 1978. J. Chem.
Phys. 69 : 1218-26
143. Pollak, E., Pechukas, P. 1983. J. Chem.
Phys. 79:2814-21
144. Pollak, E., Wyatt, R. E. 1982. J. Chem.
Phys. 77: 2689-91
145. Quack, M., Troe, J. 1974. Bet. Bunsenfles. Phys. Chem.78 : 240-52
146. Quack, M., Troe, J. 1975. Bet. Bunsenges. Phys. Chem.79 : 170-83
147. Quack, M., Troe, J. 1975. Bet. Bunsenges. Phys. Chem.79 : 469-75
148. Quack, M., Troe, J. 1977. Bet. Bunsenges. Phys. Chem.81 : 329-37
149. Quack, M., Troe, J. 1977. Gas Kinetics
Energy Transfer: Specialist Periodical
Report, 2 : 175-238. London: Chemical
150. Rai, S. N., Truhlar, D. G. 1983. J. Chem.
Phys. 79 : 6046-59
151. Ravishankara, A. R., Nicovich, J. M.,
Thompson,R. L., Tully, F. P. 1981. J.
Phys. Chem. 85 : 2498-2503
152. R6melt, J. 1983. Chem. Phys. 79 : 197209
153. Ron, S., Baer, M., Pollak, E. 1983. J.
Chem. Phys. 78 : 4414-22
154. Schatz, G. C. 1983. J. Chem. Phys.
79 : 5386-91
155. Schatz, G. C., Elgersma, H. 1980. Chem.
Phys. Lett. 73 : 21-25
156. Schatz, G. C., Kuppermann,A. 1976. J.
Chem. Phys. 65 : 4668-92
157. Siegbahn, P., Liu, B. 1978. J. Chem.
Phys. 68 : 2457-65
158. Skodje, R. T., Schwenke, D. W.,
Annual Reviews
www.annualreviews.org/aronline
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
VARIATIONAL TRANSITION STATE THEORY
Truhlar, D. G., Garrett, B. C. 1984. J.
Phys. Chem. 88:628-36
159. Skodje, R. T., Schwenke,D. W., Truhlar, D. G., Garrett, B. C. 1984. J. Chem.
Phys. 80 : 3569-73
160. Skodje, R. T., Truhlar, D. G. 1981. J.
Phys. Chem,85 : 624-28
161. Skodje, R. T., Truhlar, D. G. 1983. J.
Chem.Phys. 79 : 4882-88
162. Skodje, R. T., Truhlar, D. G., Garrett, B.
C. 1981. J. Phys. Chem. 85:3019-23
163. Skodje, R. T., Truhlar, D. G., Garrett, B.
C. 1982. J. Chem.Phys. 77 : 5955-76
164. Squires, R. R., Bierbaum, V. M.,
Grabowski,J. J., dePuy, C. H. 1983. J.
Am. Chem. Soc. 105 : 5185-92
165. Su, T., Chesnavich,W. J. 1982. J. Chem. .
Phys. 76:5183-85
166. Sverdlik, D. I., Koeppl, G. W. 1978.
Chem.Phys. Lett. 59 : 449-53
167. Sverdlik, D. I., Stein, G. P., Koeppl,G.
W. 1979. Chem.Phys. Lett. 67 : 87-92
168. Swamy,K. N., Hase, W. L. 1982. J.
Chem. Phys. 77 : 3011-21
169. Swarc, M. 1962. Discussion. In The
Transition State, Chem.Soc. Spec. Publ.
16, pp. 25-27. London: Chem.Soc.
170. Troe, J. 1981. J. Chem.Phys. 75 : 226-37
171. Truhlar, D. G. 1970. J. Chem. Phys.
53 : 2041-44
172. Truhlar, D. G. 1979. J. Phys. Chem.
83 : 199
173. Truhlar, D. G., Garrett, B. C. 1980. Acc.
Chem. Res. 13:440-48
174. Truhlar, D. G., Garrett, B. C., Blais, N.
C. 1984. J. Chem.Phys. 80 : 232-40
175. Deleted in proof
176. Truhlar, D. G., Grev, R. S., Garrett, B.
C. 1983. J. Phys. Chem. 87:3415-19
177. Truhlar,I). (3., I-Iase, W.L., bIynes, J.
1983. J. Phys. Chem.87 : 2664-82;1983.
Erratum: 87 : 5523
178. Truhlar, D_G., Horowitz, C. J. 1978. J.
Chem.Phys. 58 : 2466-76. Erratum 71 :
1514
179. Truhlar, D. G., Isaacson, A. D. 1982. J.
Chem. Phys. 77 : 3516-22
180. Truhlar, D. G., Isaacson, A. D., Garrett,
B. C. 1984. Generalized transition state
theory. See Ref. 138, In press
181. Truhlar, D. G., Isaacson, A. D., Skodje,
R. T., Garrett, B. C. 1982. J. Phys. Chem.
189
86:2252-61 ; Erratum: 1983. 87:4554
182. Truhlar, D. G., Kilpatrick, N. J.,
Garrett, B. C. 1983. J. Chem.Phys. 78 :
2438-42
183. Truhlar, D. G., Kuppermann,A. 1971.
J. Am. Chem. Soc. 93:1840-51
184. Truhlar, D. G., Kupperman,A. 1972. J.
Chem.Phys. 56 : 2232-52
185. Tweedale, A., Laidler, K. J. 1970. J.
Chem.Phys. 53 : 2041-44
186. Walch, S. P., Dunning, T. H. 1980. J.
Chem. Phys. 72 : 1303-11
187. Walker, R. B., Hayes, E. F. 1983. J.
Phys. Chem. 87:1255 63
188. Walker, R. B., Light, J. C. 1980. Ann.
Rev. Phys. Chem. 31:401-33
189. Webb,D. A., Chesnavich, W. J. 1983. J.
Phys. Chem. 87:3791-98
190. Westheimer, F. H. 1961. Chem. Rev.
61:265-73
191. Wigner, E. 1932. Z. Phys. Chem. B 19:
203-16
192. Wigner,E. 1937. J. Chem.Phys. 5 : 72025
193. Wigner, E. 1938. Trans. Faraday Soc.
34: 29-41
194. Wolf, R. J., Hase, W. L. 1980. J. Chem.
Phys. 72 : 316-31
195. Wolfe,S., Mitchell, D. J., Schlegel, H. B.
1981. J. Am. Chem.Soc. 103 : 7694-96
196. Wong, W. H., Marcus, R. A. 1971. J.
Chem.Phys. 55 : 5625-29
197. Zeiri, Y., Shapiro, M., Pollak, E. 1981.
Chem.Phys. 60 : 239-47
198. Zellner, R. 1979. J. Phys. Chem.83 : 1823
199. Zellner, R., Steinert, W. 1981. Chem.
Phys. Lett. 81:568-72
References added in proof:
200. Kreevoy, M. M., Truhlar, D. G. 1984.
Transition state theory. In Investigation
of Rates and Mechanismsof Reactions,
ed. C. F. Bernasconi. NewYork : Wiley.
4th ed. In press
201. Melius, C. F., Blint, R. J. 1979. Chem.
Phys. Lett. 64:183-89
202. Celli, F., Weddle,G., Ridge, D. P. 1980.
J. Chem.Phys. 73 : 801-12
203. Sakimoto, K. 1984. Chem. Phys.
85 : 273-78
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.
Annu. Rev. Phys. Chem. 1984.35:159-189. Downloaded from arjournals.annualreviews.org
by University of Minnesota- Law Library on 01/09/07. For personal use only.