You are on page 1of 12

Ab initio and diatomics in molecule potentials for and

Jiri Vala, Ronnie Kosloff, and Jeremy N. Harvey

Citation: The Journal of Chemical Physics 114, 7413 (2001); doi: 10.1063/1.1361248
View online: https://doi.org/10.1063/1.1361248
View Table of Contents: http://aip.scitation.org/toc/jcp/114/17
Published by the American Institute of Physics

Articles you may be interested in


The electronic structure of the triiodide ion from relativistic correlated calculations: A comparison of different
methodologies
The Journal of Chemical Physics 133, 064305 (2010); 10.1063/1.3474571

Photodissociation dynamics of the triiodide anion


The Journal of Chemical Physics 113, 2255 (2000); 10.1063/1.482040

Photodissociation of gas-phase : Comprehensive understanding of nonadiabatic dissociation dynamics


The Journal of Chemical Physics 126, 204311 (2007); 10.1063/1.2736691

Intermolecular interactions in the condensed phase: Evaluation of semi-empirical quantum mechanical methods
The Journal of Chemical Physics 147, 161704 (2017); 10.1063/1.4985605

Improving the accuracy of Møller-Plesset perturbation theory with neural networks


The Journal of Chemical Physics 147, 161725 (2017); 10.1063/1.4986081

Product energy deposition of CN + alkane H abstraction reactions in gas and solution phases
The Journal of Chemical Physics 134, 214508 (2011); 10.1063/1.3595259
JOURNAL OF CHEMICAL PHYSICS VOLUME 114, NUMBER 17 1 MAY 2001

Ab initio and diatomics in molecule potentials for I2À , I2 , I3À , and I3


Jiri Vala and Ronnie Kosloff
Department of Physical Chemistry and The Fritz Haber Research Center, The Hebrew University,
Jerusalem, 91904, Israel
Jeremy N. Harveya)
School of Chemistry, University of Bristol, Cantock’s Close, Bristol BS8 1TS, United Kingdom
共Received 8 December 2000; accepted 15 February 2001兲
The electronic structure of the I⫺ ⫺
3 molecular anion and its photoproducts I2 , I2 , and I3 were studied.
Ab initio calculations were carried out using the multireference configuration interaction 共MRCI兲
method for the valence electrons together with a relativistic effective core potential. The ab initio
wave functions were also used to compute some spin–orbit coupling matrix elements, as well as
approximate valence bond wave functions, used as guidelines in the construction of a 108-state
diatomics in molecule 共DIM兲 description of the electronic structure of I⫺ 3 . In the DIM model,
spin–orbit coupling was introduced as a sum of atomic operators. For I⫺ 2 the ab initio and the DIM
ground-state potentials show excellent agreement with the experimental results. The results for I2
are also in very good agreement with experimental data. For I⫺ 3 , the MRCI calculations give a very
good description of the spectroscopic constants and agree with the vertical excitation energies,
provided spin–orbit coupling is included. The DIM description fails both quantitively by leading to
erroneous spectroscopic constants, and qualitatively by not even reproducing the MRCI ordering of
the excited-states. The failure of the DIM is attributed to the omission of ionic states. The overall
qualitative picture of the excited-state potentials shows a maze of dense avoided crossings which
means that all energetically allowed photoproducts will be present in the experiment. The ground
electronic state of I3 was calculated to be a collinear and centrosymmetric 2 ⌸ u,3/2 . The collinear
state is stabilized by spin–orbit coupling relative to a bent configuration. Calculated vertical
transition energies from the ground to low-lying excited states of the radical are in excellent
agreement with the experimental data. The spin–orbit assignment of these states is provided.
© 2001 American Institute of Physics. 关DOI: 10.1063/1.1361248兴

I. INTRODUCTION sists in two distinct bands centered at about 290 and 360 nm.
Insight in physical chemistry is typically obtained by After some controversy,9,14–16 these two bands were assigned
comprehensive studies of a particular molecular system to the spin–orbit-induced mixture of 1 ⌺ 0⫹u and 3 ⌸ 0⫹u
where a close feedback is established between experiment states,11,12,17 where the former state carries the transition di-
and theory. The triodide molecular anion I⫺ 3 is emerging as
pole moment responsible for optical absorption.
one of these systems. Its strong optical absorption in the near More recently, Ruhman and co-workers have used fem-
ultraviolet 共UV兲 makes it readily accessible to photochemical tosecond spectroscopy to investigate the properties and dy-
studies. The photochemical products I⫺ namics of the I⫺ 3 anion. This research covered several
2 absorb in a different
spectral region. This factor and the slow dynamics induced complementary viewpoints, investigating aspects such as the
by the heavy masses have enabled the photoreaction to be light-induced vibrational dynamics on the electronic ground
followed in real time from reactants to products.1–3 The ex- state,18,19 photodissociation,1,2,4,20 and geminate
perimental observations have stimulated the development of recombination.21,22 These last studies revealed several dis-
novel theoretical tools4 addressing the problem of dissipative tinct recombination time scales and the possibility of mul-
quantum dynamics in solution. The present study is part of tiple potential-energy surfaces being involved in solution dis-
the large theoretical effort on the I⫺
3 system, addressing spe- sociation dynamics. Another aspect of this experimental
cifically the electronic structure of the parent ion and the corpus was devoted to the spectroscopy and relaxation dy-
routes to photofragments. Obtaining the ground- and excited- namics of the transient diiodide ion.23 A similar experimental
state potentials is a necessary condition for any comprehen- study of the photodissociation dynamics of I⫺ 3 was recently
sive theoretical understanding of these processes. carried out by Vöhringer and coworkers.24–26
The first experimental studies of the I⫺
3 ion were oriented The triodide anion has also been investigated using ad-
on dissociation kinetics in solution,5–7 electronic absorption vanced frequency domain spectroscopical methods,27–29
spectroscopy in the condensed phase,8–12 and thermo- which have brought considerable insight into the role of sol-
chemistry.13 The electronic spectrum of I⫺ 3 in solution con- vent effects and the dynamical symmetry breaking of the I⫺ 3
molecule.
a兲
Electronic mail: jeremy.harvey@bris.ac.uk The gas-phase properties of the di- and tri-iodide anions

0021-9606/2001/114(17)/7413/11/$18.00 7413 © 2001 American Institute of Physics


7414 J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Vala, Kosloff, and Harvey

have been extensively investigated by Neumark and co- form of configuration interaction 共CI兲 and large one-particle
workers using ultrafast photoelectron spectroscopy.3 The basis sets. However, this problem is general when accurate
photodissociation of I⫺ 2 was studied
30–34
with an emphasis on global potential-energy surfaces are needed. The specific
the electronic structure, providing experimental potentials for problem with the present iodine-containing system is due to
the X̃ 2 ⌺ ⫹
u ground state
34
and the à ⬘ 2 ⌸ g,1/2 excited state.33 the very high atomic weight of iodine, which leads to very
Photofragmentation experiments35–37 on I⫺ 3 indicate that a
important relativistic effects on the wave functions, which
complicated set of electronic states with many curve cross- cannot as yet be treated in a rigorous way, even for such a
ings is involved in the dissociation process, in agreement ‘‘small’’ system. Whilst scalar relativistic effects can be ad-
with the results of solution studies.21,22 Also, the neutral I3 equately reproduced using effective core potentials to de-
radical has been recently characterized for the first time.38,39 scribe the innermost, ‘‘core’’ electrons, the treatment of the
This species, which is important in the iodine atom recom- nonscalar effects, which are mostly due to spin–orbit cou-
bination reaction, has a ground state which is bound with pling, is much less satisfactory. Although several studies
respect to I⫹I2 , and has a linear, essentially centrosymmet- have used methods based on the Dirac equation to study
ric configuration. diatomic iodine species,40,41 application to I⫺ 3 , especially at
The dynamics of triatomic species is sufficiently simple the level of detail and accuracy needed for dynamical stud-
to be described with very high accuracy by fully quantum- ies, is still quite distant.
mechanical methods. It is, therefore, no surprise that the ex- In the present study, we present a first attempt to use ab
perimental work on I⫺ 3 has been accompanied by extensive
initio computations to characterize the excited-state surfaces
theoretical work. However, dynamical computations are and dynamics of I⫺ 3 , including the photoionization to neutral
critically dependent on obtaining accurate descriptions of the I3 . Because the chemistry of these species is so intimately
potential-energy surfaces involved, and of the coupling be- linked to that of diatomic I2 and I⫺ 2 , we have also performed
tween them where relevant. All of the dynamical studies so computations on the latter. The level of theory used, multi-
far have used an empirical description of the excited state reference configuration interaction, with a large one-particle
corresponding to the blue band of the I⫺ 3 electronic spectra,
basis set, should treat the correlation problem adequately if
constructed in terms of the London–Eyring–Polanyi–Sato not perfectly. The use of a relativistic effective core potential
共LEPS兲 potential.1,20 Because the surfaces have been roughly to treat the innermost electrons has the dual advantage of
fitted to experimental properties, reasonable agreement with decreasing the computational expense and accounting for
experiment has been obtained with respect to, e.g., the elec- most of the scalar relativistic effects. To treat the spin–orbit
tronic spectra characteristics of I⫺ 3 in solution,
1,20
the time- coupling, we have used two approaches both relying on a
resolved photoelectron spectra, or the vibrational energy dis- simplified one-electron operator. For certain important points
tribution of the I⫺ photofragment in the gas-phase on the potential-energy surfaces, we are able to compute
2
experiment.35–37 However, the experimental and theoretical spin–orbit coupling matrix elements between a limited num-
work is full of indications that the surfaces used are very ber of spin–orbit-free ab initio wave functions. To get a
approximate at best. For instance, the first LEPS surface more global picture of the effect of spin–orbit coupling, we
used1,20 reproduces the upper energy band of the solution have constructed a diatomics in molecules 共DIM兲 model of
electronic spectrum rather well, which is expected since it the valence electronic states of I⫺ 3 , which includes spin–
was constructed to do so. Less expectedly, it also reproduces orbit coupling within an atoms in molecules ansatz. Al-
the lower-energy band, providing the necessary shift in total though the results obtained are only of semiquantitative ac-
energy is introduced. However, it is not compatible with the curacy, they should already represent an improvement on
gas-phase experimental observations, so that a substantially previous semiempirical surfaces. Also, by giving for the first
modified LEPS surface36 has been used to account for the time a global, nonempirical description of the surfaces and
results of photofragmentation. A more serious problem is their interaction, the study should be a valuable guideline for
that both the solution21,22 and the gas-phase experiments36,37 future research in this area.
agree in predicting that the excited-state dynamics occurs on After describing the details of the ab initio computations
multiple potential-energy surfaces. and the Diatomics in Molecules model, we present the re-
In the case of triatomic systems, it is possible to perform sults below for each of the species separately, namely I⫺ 2 , I2 ,
enough accurate ab initio computations 共⬃103n⫺6 ⫽1000 I⫺
3 , and I3 .
points兲 to fully characterize the potential-energy surface.
Given the enormous experimental interest in the I⫺ 3 system, II. AB INITIO COMPUTATIONS
the attraction of determining its potential-energy surface us-
ing ab initio computations should be obvious. However, We have performed new ab initio computations on I2 ,
there are also many problems for such an approach. First of I⫺
2 , I⫺
3 , and I3 . All calculations used the MOLPRO 98 and
all, the dynamics of I⫺ 3 involves bond-breaking, so that non- 2000 program packages.42 Three types of calculations were
dynamical correlation will be important in many areas of the used to 共a兲 reproduce the energy curves in the absence of
potential-energy surface which need to be described. To- spin–orbit coupling; 共b兲 compute spin–orbit coupling matrix
gether with the expected role of dynamical correlation in elements between certain specified electronic states; and 共c兲
obtaining an accurate description of the potential, this means obtain approximate valence-bond wave functions for aid in
that the ab initio method used to describe the system will the construction of and comparison with the DIM model.
need to be an expensive multireference method with some The energy calculations were performed at the multiref-
J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Potentials for I2⫺ , I2 , I3⫺ , and I3 7415

erence singles and doubles configuration interaction level Arn .I⫺


2 ,
50
and the I3 radical.51,52 To be specific, these last
共MRCISD, referred to as MRCI below兲, with the added models rely on diatomic data for I2 and I⫺ 2 in the ⍀, ␻ rep-
Davidson correction to provide approximately size- resentation, that is, Hund’s case 共c兲 adiabatic potential-
consistent results. The CI wave function was built on a com- energy curves including the effect of spin–orbit coupling.
plete active-space self-consistent field 共SCF兲 reference, with The advantage is that some of these curves, the lower-lying
the active space comprising all the explicitly described elec- ones in particular, are very well known from experiment, and
trons, which for the triodide anion, for example, means 22 of course, therefore, incorporate exactly all the effects such
electrons in 12 orbitals. The spin–orbit coupling matrix ele- as correlation, relativity, spin–orbit 共SO兲 coupling, etc.,
ments were computed between CASSCF wave functions which are hard to obtain from computation.
computed with the same active space. The approximate VB However, evaluating a DIM model for I⫺ 3 constructed in
wave functions were constructed by first performing a this way by comparing it to spin–orbit-free ab initio compu-
CASSCF computation with a slightly smaller active space in tations would be difficult. For this reason, we use instead a
which the 5s orbitals were constrained to be doubly occupied DIM model based on ⌳,⌺-eigenstates of the diatomic spe-
in all configurations. The resulting orbitals were then local- cies. Because these are not very realistic models of the
ized within the reduced-symmetry C 2 v point group, and the potential-energy curves, we add a simple spin–orbit coupling
orbital and CI coefficients of the wave function within this operator to the DIM Hamiltonian 共this is referred to below as
orbital set examined. Localization was found to lead to al- the DIM⫹SO model兲. For comparison of the DIM model
most entirely pure 5p-like orbitals, localized on the indi- with computations, the simple spin–orbit free results can be
vidual iodine atoms. used 共referred to simply as DIM below兲. Naumkin44 has ar-
In all these calculations, the same 46-electron relativistic gued that the ⌳,⌺ approach to DIM models for iodine-
ECP43 was used, in the form implemented in MOLPRO.42 The containing species is actually better than the other approach
valence 5s and 5p electrons were, however, described by 共provided atomic spin–orbit coupling is included兲, because
different basis sets. In the energy calculations, the the construction of the diatomic Hamiltonian requires fewer
(7s7 p2d1 f )/ 关 3s3 p2d1 f 兴 basis associated with the ECP43 arbitrary assumptions.
was used after decontracting the two most diffuse contracted The basis set for the DIM Hamiltonian is constructed
s and p primitives, and replacing the standard polarization from all the spin-singlet and 共triply degenerate兲 spin-triplet
functions with a 共loosely energy-optimized for neutral and states which can be constructed from two iodine atoms and
anionic iodine atom兲 set of 3 d ( ␣ ⫽1, 0.4, and 0.16兲, 2 f one iodide anion. This leads to a total of 27⫹3⫻27⫽108
( ␣ ⫽0.632, 0.252兲, and 1 g ( ␣ ⫽0.45) uncontracted primi- primitive states to be included in the calculation. The fitting
tives, to yield a (7s7 p3d2 f 1g)/ 关 5s5 p3d2 f 1g 兴 basis. In of the ab initio data to the DIM Hamiltonian is for the most
the spin–orbit coupling calculations, supplementary spin– part straightforward and follows an identical approach to that
orbit coupling parameters for the ECP were used, as docu- of Naumkin.44 In particular, the correspondence Table be-
mented in the MOLPRO manual,42,43 together with the stan- tween molecular states and atomic configurations for I2 is
dard (7s7p2d1 f )/ 关 3s3 p2d1 f 兴 basis set. Calculations with mostly taken from his work. We note, however, that our
this basis yield a 2 P 3/2 to 2 P 1/2 splitting of 7601 cm⫺1 , in assignment of the ⌸ g and ⌸ u states is taken instead from
excellent agreement with the experimental value of 7603 Ref. 53, and is thus reversed with respect to that of Ref. 44,
cm⫺1 . The approximate valence bond 共VB兲 computations which appears to be incorrect.
also used the standard basis set, but with the f functions As in both these references, we assume that there is no
removed. mixing between the two 1 ⌺ ⫹ g configurations or between the
Gradients for geometry optimization and second deriva- two 3 ⌺ ⫺ configurations. This is based on the large energy
u
tives for frequencies were obtained by numerical differentia- gap between the two configurations.44 To test the validity of
tion of the Davidson-corrected MRCI energies; note that this assumption, we computed approximate ab initio valence
only the stretching frequencies were considered. Fully state- bond wave functions at the I2 ground-state minimum. The
optimized orbitals were used throughout, except for compu- weight of atomic configurations such as Px Px in the X 1 ⌺ ⫹ g
tation of spin–orbit coupling matrix elements, excitation wave function, and of the Pz Pz configuration in the II 1 ⌺ ⫹ g
transition dipole moments, and second or higher roots within wave function, are very small. These small, if nonzero,
a given symmetry species, where state-averaged orbitals weights justify the approximation used here.
were used. Bond dissociation energies were calculated with Analytical functions fitted to the ab initio data are used
respect to the appropriate supermolecular asymptote. for the diatomic curves. Curves with a significant minimum
共deeper than 0.001 Hartree兲 are fitted to an extended Morse
III. DIM curve of the form D 0 兵 1⫺exp(⫺␤1(r⫺r0)⫺␤2(r⫺r0)2⫺␤3(r
⫺r0)3)其2. The purely repulsive curves are fitted to extended
The DIM model used is very similar to the one used by exponential curves of the form A exp(⫺␣r⫺␣⬘r2).54
Naumkin44,45 to describe the Ar.I2 and Ar.I⫺ 2 species. For a The spin–orbit coupling is treated approximately as a
general description of the DIM method and the common ap- sum of atomic operators, independent of molecular geom-
proximations used in its construction, the reader is referred etry, which is added to the normal DIM Hamiltonian. The
to the literature.46–48 It should be pointed out that the DIM form of the operator for one iodine atom in terms of the P x ,
model used here is considerably different from other DIM P y , P z basis set can be found in the literature.44 The ap-
models which have been used to describe Xen .I2 , 49 proximation we use is valid when the atomic orbitals making
7416 J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Vala, Kosloff, and Harvey

It is to be noted that within the simplified treatment of spin–


orbit coupling included here, the 2 ⌸ g,3/2 state is not affected,
so that the ab initio and DIM⫹SO constants are identical.
The bond length and energy are in good agreement with
experimental values. The main effect of the spin–orbit cou-
pling on the ground state is to stabilize the atomic asymptote
and thus decrease the bond energy. At the minimum, al-
though spin–orbit coupling of the 2 ⌺ ⫹ u state with the ⌸ u
2

states is still strong, the latter are so much higher in energy


that the effect on the ground-state energy is minimal, that is,
it is lowered in energy with respect to the spin–orbit free
species by only 0.074 eV. The bond length is slightly in-
creased, by 0.024 Å, and the harmonic frequency slightly
decreased. Overall, the agreement with experiment is rather
good, and this extends to the whole potential-energy curve.
Fitting to the extended Morse curve D e 关 1⫺exp兵兺i⫽1 5
⫺i
⫺␤i⌬r 其兴 yields ␤ parameters 共in Å ) of 1.175 84,
i 2
FIG. 1. DIM⫹SO potential-energy curves for I⫺
2 .
⫺0.012 335 8, ⫺0.004 880 43, ⫺0.000 263 595, and
0.000 192 631. Apart from the slight difference in bond
up the molecular orbitals are not substantially distorted with length and energy, this curve is almost superimposable on
respect to the isolated atom orbitals, and comparison with a the experimental one.34
few ab initio computed spin–orbit coupling matrix elements As a final note on the I⫺ 2 species, we note that whilst the
suggests that the approximation performs very well. In fact, present bond length and energy are in good agreement with
the ab initio values computed with MOLPRO are also only experiment, this has not been the case in all ab initio studies.
approximate, because they are not derived from the full This is to be expected for the bond energy, which is strongly
Breit–Pauli spin–orbit coupling Hamiltonian 共which in- subject to correlation effects, but somewhat less so for the
cludes one-electron and two-electron terms兲, but from a sim- bond length, which can usually be well reproduced even with
plified one-electron operator. However, unlike the more dras- fairly low levels of theory. We found the bond length to be
tic approximation used in the DIM approach, the ab initio quite sensitive to basis set and correlation treatment, so that
computation accounts rigorously for orbital relaxation in the at the complete active space self-consistent field 共CASSCF兲
different electronic states, so the agreement between the two level with the same basis set, it is extended to 3.257 Å,
suggests that both approaches are at least qualitatively quite whereas at the MRCI level with a reduced spd basis, it is as
reliable. much as 3.291 Å. This no doubt accounts for previously
noted discrepancies. Finally, test calculations suggest even
IV. I2À better agreement can be reached—for example, MRCI calcu-
lations using an added core-polarization potential 共CPP兲40,43
The diiodide anion has been the subject of many previ- lead to a bond length of 3.164 Å and a D e of 1.259 eV.
ous computational studies,55–60 including at least two which Assuming the spin–orbit coupling correction to be the same
have calculated all the states originating from the ground- as above leads to a prediction of R e ⫽3.188 Å and D e
state atoms, with and without including spin–orbit ⫽1.017 eV. Coupled-Cluster 关CCSD共T兲兴 calculations yield
coupling.59,60 Our four MRCI curves are similar to previous almost identical results to MRCI computations with the same
results, and are not shown here. The six DIM⫹SO curves are basis.
shown in Fig. 1, and the spectroscopic constants of the two
states displaying significant minima are included in Table I.
V. I2
TABLE I. Ab initio and DIM⫹SO computed spectroscopic constants for I⫺ 2 The iodine molecule has been the object of innumerable
and I2 . DIM⫹SO results for the ⌸ states listed here are identical to the investigations, both experimentally and computationally,
MRCI results.
only a few of which can be cited here.40,41,61,62 Several stud-
Species State Method R e 共Å兲 D e 共eV兲 ␻ e (cm⫺1 ) ies have addressed all 12 spin–orbit free and 23 spin–orbit
coupled potential energy curves correlating to two ground-
2
⌺⫹ MRCI 3.192 1.226 115
I⫺ 2
⌺⫹
u
DIM⫹SO 3.216 0.984 109 state iodine atoms.63,64 Our spin–orbit free results are similar
2 u,1/2
2 ⫹
⌺ u,1/2 exp. a 3.205 1.014 110⫾2 to those found in these studies and are not shown here. The
2
⌸ g,3/2 MRCI 3.984 0.130 97 spin–orbit coupled curves are shown in Fig. 2. The com-
1 ⫹
⌺g MRCI 2.659 1.978 221 puted spectroscopic constants for the ground and excited B
I2 X̃ 0 ⫹
g
DIM⫹SO 2.672 1.444 213 states are shown in Table I for reference. The major discrep-
X̃ 0 ⫹
g
exp. b 2.666 1.55 215 ancy concerns the ground-state bond energy, which is
3
⌸u MRCI 3.060 0.414 81 slightly too low, as expected from previous results which
a
Reference 34. have shown it to converge only slowly with the size of the
b
Reference 77. one-particle basis set.40
J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Potentials for I2⫺ , I2 , I3⫺ , and I3 7417

TABLE III. Energies 共in eV兲 of selected states of I⫺


3 relative to the ground-
state minimum, at the MRCI and DIM levels.

State E rel 共MRCI兲 E rel 共DIM兲 Expt.a


1
⌺⫹g 0.000 0.000
1
⌺⫹u 4.496 2.003
1
⌸g 2.924 2.222
1
⌸u 3.620 2.455
II 1 ⌸ g 4.594 4.647
1
⌬g 5.736 4.653
3 ⫹
⌺u 2.582 1.726
3
⌸g 2.626 1.985
3
⌸u 3.517 2.277
II 3 ⌸ g 4.355 4.557
3
⌬u 5.939 4.801
I2 ⫹I⫺ 1.247 0.715 1.31⫾0.06
I⫺
2 ⫹I3/2 2.090 1.467 1.85⫾0.06
I⫹I⫹I⫺ 3.209 2.693 2.85⫾0.06
2
⌸ I3 4.414 4.226⫾0.013

Spectral bands E rel 共MRCI⫹SO兲 Expt.b

0.84 1 ⌺ ⫹
u ⫹0.16 ⌸
3
4.654 4.28
1 ⫹
0.16 ⌺ u ⫹0.84 3 ⌸ 3.661 3.44

FIG. 2. DIM⫹SO potential-energy curves for I2 . The ground 1 ⌺ ⫹


g and
a
References 32 and 37.
b
excited 3 ⌸ u ,0 states are shown in bold. Reference 9.

The spin–orbit coupling between the ground state and periment, a fact which should not be badly affected by the
the excited states is almost completely quenched at the mini- omission of spin–orbit coupling, because this is a closed-
mum, due to the large energy separation between them. This shell species which also dissociates into closed-shell frag-
leads to the residual SOC stabilization at the minimum being ments. This can be verified by comparing the values obtained
of only 0.095 eV, a value comparable to that obtained in with and without spin–orbit coupling with the DIM model.
previous studies.64 As discussed below, the absolute values at these levels are
rather incorrect, but the difference between them should pro-
vide a useful indication of the likely magnitude of spin–orbit
VI. I3À
effects in this species. As can be seen in the Table, neither
The triodide ion has been the subject of a number of the bond length, the stretching frequencies, nor the bond en-
previous computational studies. Many of these have focused ergy are much changed. For the latter quantity, this reflects
only on the electronic ground state, considering such aspects the fact that the residual stabilization by spin–orbit coupling
as its geometry, bond energy, and electronic 共hypervalent兲 at the minimum, of 0.109 eV, is almost identical to that of I2
structure.65–68 However, a number of studies have also ad- at its minimum.
dressed the vertical excitation energies11,55 and part of the Turning to the DIM results, the immediate conclusion to
dissociation pathways.16 be drawn from Table II is that whilst the bond length of the
Unlike for the diatomic species, where the DIM model ground-state I⫺ 3 is well reproduced, neither the bond energy
itself exactly reproduces the ab initio data, for I⫺
3 , the DIM nor the vibrational frequencies are correct. The suggested
results will not match the MRCI results, and the two sets of reason for this poor agreement is discussed below.
data will need to be considered separately and compared. Computed vertical excitation energies 共up to ⬃5 eV兲 of
The computed spectroscopic constants of I⫺ 3 are com- the lower-lying excited states, together with energies of the
pared in Table II to available experimental data. As can be relevant dissociation asymptotes and of the I3 neutral, are
seen, the MRCI results are in excellent agreement with ex- presented in Table III.
Starting with the ab initio results, it can first of all be
seen that I⫺3 supports a fairly large number of excited states
TABLE II. Spectroscopic constants of ground-state I⫺
3 . which are below or close to the level of the neutral. How-
Method R e 共Å兲 D e 共eV兲 ␻ e 共sym兲 ␻ e 共asym兲 ever, some of the states shown, including the important 1 ⌺ ⫹ u
one, and all of the other computed states not shown here
MRCI 2.930 1.247 114 147
because they lie at higher energy, are in formal terms reso-
DIM 2.949 0.715 100 101
DIM⫹SO 2.966 0.729 95 104 nances and not stable states.69
Expt. 2.93a 1.31⫾0.06b 112⫾1c 145d Of the states shown, the lower one with 1 ⌺ ⫹ u symmetry
a
is important because it is the only one to have an appreciable
In crystals, see Ref. 14 and reference therein.
b excitation transition dipole moment, of 11.53 Debye, with
Reference 78.
c
Reference 32. the ground state. Optical transitions to most of the other
d
In solution, Ref. 8. states are symmetry forbidden, and the transition moment for
7418 J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Vala, Kosloff, and Harvey

the allowed excitation to the 1 ⌸ u state is very weak, at 0.16 it should still be manageable computationally, and the num-
Debye. ber of extra ab initio diatomic curves required would not be
Turning to the DIM results, it can immediately be seen insuperable. The main problem would be the tedious and
that the agreement of the spin–orbit free results with the ab arbitrary fitting work which would be needed to derive the
initio data is extremely poor. Part of this disagreement is due necessary Hamiltonian matrix elements for a global model of
to the poor description of the ground-state, discussed above, good quality. It is essentially because of the unstraightfor-
which is the origin of energy in Table III. Had the dissoci- ward nature of this task that we did not extend the model in
ated atom asymptote been chosen as a zero of energy, the this way.
overall agreement would have been slightly better. However, The overall effect of spin–orbit coupling is difficult to
this would not have altered the fact that the ordering of the discuss, because only the DIM⫹SO approach gives a global
states in the DIM method is qualitatively and quantitatively picture, yet is unreliable due to the underlying DIM prob-
wrong. In terms of the photochemistry, it is particularly no- lems. However, a fairly accurate idea can be derived from
ticeable that the 1 ⌺ ⫹
u state is predicted to lie close to or even the MRCI energies, and a limited number of ab initio calcu-
below the 3 ⌸ states, and also well below the three-atom lations of individual matrix elements, as to what the effect of
dissociation limit. spin–orbit coupling is on the absorption spectrum. As could
Analysis of the DIM and ab initio wave functions sug- be expected from symmetry considerations, the 1 ⌺ ⫹ u state
gests that the reason for the poor quality of the DIM results is couples strongly only to ⌸ u states, and only one of these is
3

the limited size of the DIM basis set. The latter contains only at low enough energy to mix with it substantially. The matrix
those states derived from two neutral iodine atoms and one element with this state is computed to be of 1361 cm⫺1 . It is
iodide anion. The true wave function also contains ‘‘ionic’’ to be noted that despite the poor energetics and simple spin–
states with two iodide anions and one cationic iodine, such as orbit coupling operator, the DIM method gives a rather simi-
I⫺ •••I⫹ •••I⫺ . If one considers the ground 1 ⌺ ⫹ g and excited lar result, of 1487 cm⫺1 .
1 ⫹
⌺ u states, the extent to which this ‘‘ionic’’ configuration, Upon diagonalizing a model Hamiltonian with the
which is strongly stabilized by Coulombic interactions, MRCI energies as diagonal elements, and the ⌺/⌸ and ⌸/⌸
mixes into both of them can readily explain the mismatch 共⫽2428 cm⫺1 ) SOC matrix elements in the off-diagonal po-
between the ab initio results, which include this mixing, and sitions, one obtains one pure ⌸ state and two mixed 3 ⌸ u /
1 ⫹
the DIM results, which do not. The ground state of I⫺ 3 is best ⌺ u states. The lower of these lies 3.661 eV above the 共as-
described as an in-phase linear combination of I⫺I•••I⫺ and sumed to be unperturbed兲 ground state, and is 0.84 by weight
I⫺ •••I⫺I, but symmetry allows other gerade configurations 3
⌸ in nature, with the 1 ⌺ contributing 0.16 to the weight.
such as the ionic one above, to mix in as well. The excited The upper state has the reverse composition and lies 4.654
state is best described as the phase-reversed combination of eV above the ground state. This leads to a predicted spec-
these two covalent states, with the mixing-in of the ionic trum containing one low-energy, low-intensity band, and one
configuration indicated above symmetry forbidden. It is im- high-energy, high-intensity band. These results are in excel-
portant to realize that the DIM method implicitly attributes a lent agreement with the experimental spectrum11,37 peaked
certain stabilizing effect of ionic configurations to all states around 3.44 eV 共360 nm兲 and 4.28 eV 共290 nm兲. The pre-
described in part by the diatomic ground-state potential- dicted ratio of the intensities is also of the correct order of
energy curve of I2 . This is because the corresponding di- magnitude. This suggests that the essential physics of the
atomic Hamiltonian matrix element is directly derived from absorption process is contained by considering only the mix-
the MRCI potential-energy curve, which owes part of its ing between the two states. It is to be noted that an identical
stability to an optimum mix of covalent and ionic states. assignment was suggested already by Okada,11 based on
Therefore, DIM predicts both the ground and excited states Hückel calculations with added spin–orbit coupling; our cal-
of triodide to be significantly stabilized by admixture of ionic culations provide a much firmer theoretical footing to this
states. In fact, the ground state is more ionic than predicted conclusion. The same assignment was also derived in the
by DIM, because of the very favorable nature of the charge- context of an experimental study17 of the electronic absorp-
alternating structure described above. On the other hand, the tion spectroscopy and magnetic circular dichroism of the ion.
excited state is considerably less ionic than predicted, be- Turning now to the dissociation behavior of the different
cause the only symmetry allowed ionic configurations, such states, we first show in Fig. 3 the MRCI potential along the
as I⫺ •••I⫺ •••I⫹ , are highly destabilized. Therefore, the symmetric stretch coordinate for several of the lower-lying
ground state is more stable than predicted by DIM, and the states of I⫺ 1 ⫹
3 . Both the ⌺ u and ⌸ u states are quite strongly
3

excited state less, as can be seen in Table III. Approximate repulsive at the ground-state minimum. However, the former
valence-bond wave functions generated for iodine and trio- state barely displays a minimum along the dissociation path-
dide are in full agreement with this analysis. Note also that way, whereas the ⌸ state has a rather deep minimum, of
similar arguments can be used, e.g., to understand the over- ⬃0.8 eV, at a bond length close to 3.5 Å. This is rather
estimated stability of the 3 ⌸ u state. similar to the behavior of the LEPS excited-state potential.
The remedy for this poor behavior of DIM is obvious: During photodissociation, such a minimum might be ex-
Include ionic states in the basis.70 There is very little doubt pected to 共partly兲 ‘‘focus’’ the dissociating wave packet,
that, done properly, this would yield much better results, whose momentum is initially directed almost entirely to-
although probably still not comparable in quality to the wards the three-body asymptote, into the two-body valleys
MRCI data. Although the DIM Hamiltonian would be larger, instead. If this is the case, photodissociation via the lower of
J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Potentials for I2⫺ , I2 , I3⫺ , and I3 7419

FIG. 3. MRCI computed potential energy curves along the symmetric


stretch dissociation curve, for I⫺ 1 ⫹ 3
3 . Shown in bold are the X ⌺ g , ⌸ u , and FIG. 4. DIM potential energy curves along the symmetric stretch dissocia-
1 ⫹
⌺ u states. tion curve, for I⫺ 1 ⫹ ⫹
3 . Shown in bold are the ground ⌺ g state, and all 0 u states
which are of same Hund’s case 共c兲 symmetry as the 1 ⌺ ⫹ u state which has a
strong transition dipole moment with the ground state.

the two absorption bands would be expected to lead to some-


what less intensity in the three-body channel than that via the
upper band, which is mainly 1 ⌺ ⫹ u in nature. This rather sim- experimental observation37 that photodissociation via either
plistic dynamical prediction ignores, however, the expected of the bands always leads both to ground and excited iodine
strong effect of spin–orbit coupling. atoms.
As expected from the poor prediction of vertical ener- Overall, Fig. 4 highlights just how many states are
gies, the corresponding diagonal curves computed with the present in the energy régime at which dissociation takes
spin–orbit free DIM model are in medium to poor agreement place. This shows that nonadiabatic behavior is almost cer-
with the MRCI results. They are, therefore, not shown here. tain to occur, as is observed experimentally, so that the suc-
However, the curves computed using the DIM⫹SO model cess of previous modeling studies based on single surfaces is
are of some interest because they should give a qualitative probably due, at least in part, to a fortunate averaging out of
idea of the ‘‘real’’ potential-energy curves. The 63 nonde- the multistate dynamics. It is also possible, of course, that the
generate curves are displayed in Fig. 4. As can be seen, the DIM⫹SO surfaces are so incorrect that they also exaggerate
state with maximum 1 ⌺ ⫹ u character is very different from the the density of states in the excitation and dissociation win-
MRCI-computed curve in that it has a deep minimum near dows, and thereby also exaggerate the complexity of the ex-
3.3 Å. This is mainly due to the incorrect underlying spin– pected dynamics.
orbit-free DIM description, which also leads to a deep mini- The final element of the triodide potential-energy sur-
mum, so that no real conclusions can be drawn from this faces that we shall consider is the nonsymmetrical, two body
observation. What is more significant is that the second 0 ⫹ u dissociation pathway leading either to I2 ⫹I⫺ or to I⫺ 2 ⫹I. The
state, which is mainly 3 ⌸ u in nature, can be seen to dissoci- DIM⫹SO potential-energy curves are plotted in Fig. 5 as a
ate to the first excited asymptote (I⫺ ⫹I⫹I*兲. It thereby un- function of one I–I bond length, and for two different values
dergoes a weakly avoided crossing with a 1 ⌺ ⫹ u state, whose of the other, fixed, bond length. In Fig. 5共a兲, one bond is
underlying electronic structure is that of a phase-reversed fixed at the equilibrium value in ground-state triodide 共2.95
combination of the diatomic II 1 ⌺ ⫹ g states of I2 . The weak Å兲, while in Fig. 5共b兲, the fixed bond has the extended length
nature of the avoided crossing can, therefore, be traced back of 3.3 Å. This should give a better picture of the potential
to the weak interaction between the diatomic X 1 ⌺ ⫹ g and II experienced during photodissociation, because the excited-
1 ⫹
⌺ g states. 共In the DIM model, these states are assumed not state potentials are initially very repulsive along the symmet-
to mix at all, directly, but they do mix a bit indirectly due to ric stretch coordinate, so that a substantial extension of both
the spin–orbit coupling; this should roughly match the real I–I bonds will occur initially whatever the excited state and
coupling, which is expected to be weak, as discussed in the the details of the potential.
DIM section.兲 In the real potentials, it is likely that either the In Fig. 5共a兲, the energy ordering of the two lowest dis-
1 ⫹
⌺ u excited state or the 1 ⌸ u , or both, undergo weakly sociation asymptotes is as expected: I2 ⫹I⫺ lies below I⫺ 2 ⫹I.
avoided crossings of this type. This could explain the These latter, excited products, are formed adiabatically from
7420 J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Vala, Kosloff, and Harvey

FIG. 5. DIM⫹SO potential-energy curves along the nonsymmetric dissociation coordinate, for I⫺ 3 . One of the bond lengths is set equal to 2.95 Å 共a兲 or 3.30
Å 共b兲. The curves of states of 0 ⫹ symmetry are shown with bold lines. The nature of the two lowest dissociation asymptotes is shown in both cases.

dissociation of the I⫺3 excited states. However, in Fig. 5共b兲,


of complicated photodissociation pathways as obtained in
corresponding to the case where the diatomic iodine–iodine recent experimental observations in solution21,22 and the gas
distance is extended, the energy ordering of the asymptotes phase.32,37
is reversed: The I⫺2 ⫹I products actually lie lowest in energy.
This conclusion concerning the diatomic asymptotes is actu-
VII. I3
ally true despite the severe shortcomings of the DIM method
in the triatomic region. This is because the potentials towards The triodide radical has been the object of extremely few
the right-hand side of Fig. 5 are essentially diatomic. The computational investigations, with none of the published re-
reversal in order of the states is due to the fact that at r sults being at a particularly reliable level of theory. Viste71
⫽3.3 Å, the neutral I2 bond is strongly extended, whereas I⫺ 2 studied the I⫹I2 system using relativistically parametrized
is near its minimum, thereby compensating for the greater extended-Hückel calculations, and found a linear transition
intrinsic stability of the former state. From Fig. 5共b兲, it is state at about r e ⫽3.2 Å on the lowest-lying potential-energy
apparent that there must be a quite strongly avoided crossing surface, which was assigned a 2 ⌫ 3/2 symmetry within the
between the two states along the dissociation pathway, and C ⬁ v point group. A DIM model for the ground state of the
this is indeed clearly so from an analysis of the DIM wave radical has been proposed,51,52 which has assumed that this
functions. state has a 2 ⌺ symmetry (J⫽1/2), and some extended
As a result of this, the observed production of I⫺ 2 ⫹I Hückel computations have been reported.72 In a recent ex-
during gas-phase and solution photodissociation experiments perimental study,39 unpublished higher-level computations
becomes at once easier and more difficult to understand. On are cited which seem to agree with our conclusion, below,
the one hand, Fig. 5 is suggestive in explaining why the that the triodine radical has a bent symmetric structure in the
solution experiments do not lead to the ‘‘ground-state’’ prod- absence of spin–orbit coupling.
ucts, namely molecular iodine and the iodide molecule, de- We have optimized the structure of I3 within the D⬁h ,
spite the expected collisions with the solvent. This appears to C ⬁ v , and 共bent兲 C 2 v point groups. Starting from a linear
be because the nature of the dynamics means that the system structure with two different bond lengths, convergence to the
only visits regions of the potential-energy surface where the symmetric geometry is observed, with r e ⫽2.836 Å and a
‘‘excited’’ products are actually more stable. On the other 2
⌸ u electronic state. This state lies just 0.019 eV below the
hand, though, the gas-phase experiments must involve I2 ⫹I dissociation asymptote, and has vibrational stretching
surface-hopping to the lower state if ground-state I⫺ 2 and I frequencies of 122 共sym兲 and 209 共asym兲 cm⫺1 . The former
are formed, as is the case at least part of the time. The exact value is in fair agreement with the tentatively assigned ex-
dynamics describing this crossing are however hard to pre- perimental value of 115⫾5 cm⫺1 . The next highest state
dict at present given the poor accuracy of the DIM potential within the D⬁h point group is a 2 ⌺ ⫹ g state, with r e ⫽2.860 Å,
in the Franck–Condon region. However, the theoretical con- but this lies 0.150 eV above the asymptote. At the minimum
clusions are in qualitative accord with the inferred existence of the 2 ⌸ u state, 2 ⌸ g and 2 ⌺ ⫹
u excited states are found to lie
J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Potentials for I2⫺ , I2 , I3⫺ , and I3 7421

TABLE IV. The ground and low-lying excited states of I3 . A similar simple model for the spin–orbit coupling can
be used for the bent structures. The 2 A1 component of the
State E rel(MRCI⫹SO兲 Expt.a 2
⌸ u state lies 0.253 eV below the linear minimum at a bend-
2
⌸ u,3/2 0.000 0.000 ing angle of 145.4 degrees, whereas the 2 B1 component lies
2
⌺ g,1/2 0.276 0.279
0.470 eV higher in energy at the 2 A1 minimum, where the
2
⌸ u,1/2 0.605 0.615
2
⌸ g,3/2 0.775 0.679 two states are thus split by 0.724 eV. As well as this strong
2
⌸ g,1/2 1.624 splitting, the spin–orbit coupling between the two states de-
2
⌺ u,1/2 3.356 creases drastically, to be of only 930 cm⫺1 . This leads to a
a much smaller spin–orbit stabilization, of only 0.018 eV, at
Reference 30.
the 2 A1 minimum, compared to the 0.335 effect for the 2 ⌸ u
linear state. This leads to a reversal of the order of stability
of the linear and bent minima upon including spin–orbit cou-
at high energy, respectively, 0.731 and 2.942 eV above the pling, with the linear state lying some 0.06 eV lower. This is
dissociated asymptote. in agreement with experimental results38 which suggest that
As in the case of the Br3 radical,73 however, the lowest the I3 radical is a linear centrosymmetric species. Similar
energy is found for the bent structure, due to Renner–Teller conclusions were reached for the Cl3 species,74 which is also
coupling which splits the two components of the 2 ⌸ u state to predicted to be linear 共but not centrosymmetric兲 due to spin–
form 2 A1 and 2 B1 states. The former of these, with an r e of orbit coupling, despite the overall spin–orbit free minimum
2.808 Å, and an angle of 145.4° is found to lie 0.272 eV corresponding to a bent configuration.
below the asymptote. As mentioned before, a DIM model for I3 has been sug-
These results are all obtained without spin–orbit cou- gested, and assumes that the ground state is linear and has a
pling, which, as in the other species studied here, is expected value of 1/2 for the projection of the total angular momen-
to have major effects on the relative energies of different tum onto the molecular axis. In terms of Hund’s case 共a兲
states. To assess this effect, we have computed the spin– states, this would mean the ground state is mostly 2 ⌺ in
orbit coupling elements between the lowest 2 ⌺ g,u and 2 ⌸ g,u nature. This would be in stark contrast with the situation for
states at the linear 2 ⌸ u minimum, and diagonalized the re- both Cl3 74–76 and Br3 , 73 which both have ground states cor-
sulting matrix 共with the MRCI energies as diagonal ele- responding to 2 ⌸ states. The computations described here
ments兲. We obtain six eigenvalues, the lowest of which cor- allow an unambiguous assignment of the ground state as
responds to the 2 ⌸ u,3/2 component of the spin–orbit free being primarily 2 ⌸ u,3/2 in nature. This is because this state is
ground state, and is stabilized by the value of the coupling the lowest in energy in the spin–orbit free ab initio compu-
between the two components, i.e., by 2703 cm⫺1 or 0.335 tations, and it is the only one which is strongly stabilized by
eV. A similar limited SOC calculation for the I2 ⫹I asymp- spin–orbit coupling. The 2 ⌺ ⫹ g state does not lie much higher
tote would lower it by roughly 0.314 eV 共due to the atomic in energy, but is only coupled to very high-lying 2 ⌸ g states,
coupling, neglecting the effect on I2 ). This leaves the ⌸ state and so will not be greatly stabilized by spin–orbit coupling.
overall at 0.040 eV below the asymptote, in moderate agree- It is unclear at this stage whether the disagreement with
ment with the experimental value of 0.143 eV.38 The adia- the DIM model is due entirely to the assumption made in that
batic electron affinity of I3 can be derived from the ab initio model that the ground state is 2 ⌺ in nature, or whether there
total energies of the anion and neutral minima 共⫺34.283 42 are similar intrinsic problems with the DIM methodology
and ⫺34.123 18 a.u., respectively兲 and the approximate SOC applied to I3 to those we found to occur for I⫺ 3 . The con-
stabilization of the neutral 共0.335 eV兲. This gives a value for struction of a DIM model based on L,S atoms with added
Ea (I3 ) of 4.025 eV, in similar agreement with the experi- spin–orbit coupling, as done here for I⫺ 3 , would be relatively
mental value of 4.226⫾0.013 eV.38 The diagonalized SOC trivial 共although the basis set constructed from all states cor-
matrix also leads to predictions for the energies of the ex- relating to 3 2 P iodine atoms would be somewhat larger, with
cited states of the neutral, which are shown in Table IV. As 216 states兲, but was not attempted here.
can be seen, the relative energies of the four lowest levels
共0.00, 0.28, 0.61, and 0.77 eV兲 are in excellent agreement
with the experimental observation of states lying at 0.279, VIII. CONCLUSIONS
0.615, and 0.679 eV. These results are summarized in A theoretical understanding of the photoreactions of I⫺ 3
Table V. requires a detailed knowledge of the potential-energy sur-
faces involved. The very large nuclear charge of iodine pre-
cludes the use of all electron methods, and leads to very
TABLE V. Spectroscopic constants of ground-state I3 2 ⌸ u . important scalar relativistic and spin–orbit coupling effects,
of the same order of magnitude as the binding energy of
Method R e 共Å兲 D e 共eV兲 ␻ e (sym) ␻ e 共asym兲 many of the species studied here. Due to these difficulties
MRCI 2.836 0.019 共0.040 兲a
122 209 any electronic structure description has to be a compromise.
Expt.b 0.143 115⫾5 The main computational approach employed here is
multireference configuration interaction 共MRCI兲 using a
a
The number in brackets refers to the bond energy of the 2 ⌸ u,3/2 state when
spin–orbit effects are approximately included. large one-particle basis set and a relativistic effective core
b
Reference 38. potential 共ECP兲. Within this framework the spin–orbit cou-
7422 J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Vala, Kosloff, and Harvey

pling has been added empirically. The MRCI approach the mixing of ionic and covalent states in a flexible way, due
yields extremely good results for the ground-state properties to the absence of ionic basis states in the DIM model.
of all species studied. The results are found to have predic- Nevertheless, the qualitative appearance of the
tive value in assigning vibrational frequencies to experimen- DIM⫹SO excited-state potentials for I⫺ 3 does have important
tal results. Also, the computed geometries agree extremely implications for the molecular dynamics. Once the system is
well with the available experimental data. It is noted that a promoted to the excited states, the mazelike nature of the
correct prediction of the bond energy and bond length re- potentials, with their numerous closely lying avoided cross-
quires an accurate treatment of correlation effects together ings, will mean that the amplitude of the excited wave func-
with rather large basis sets and an ECP. In the case of I⫺ 2 , tion will be distributed onto many different states. This
this treatment leads to a significantly improved prediction of means that the simple picture whereby the photoreaction is
both molecular descriptors compared to previous ab initio modeled by following a single potential surface from reac-
studies 共for a summary see Ref. 36, for instance兲. In particu- tants to products is misleading. Instead, the DIM⫹SO model
lar, our results describe the whole ground-state potential of predicts that all energetically allowed photofragments will be
this ion equally well, including the parts sampled by very produced, which is in accordance with existing experiments.
highly excited vibrational states.
For the excited states, the involvement of spin–orbit ACKNOWLEDGMENTS
coupling is much more crucial. Due to the very large number
We thank Professor Gabriel Balint-Kurti, Professor San-
of states it was not realistic to treat the spin–orbit coupling
ford Ruhman, and Professor Dan Neumark for many valu-
between all of them at a rigorous level. However, an empiri-
able discussions. J.V. thanks Erez Gershgoren for many in-
cal treatment of the spin–orbit coupling yields good agree-
teresting conversations about peculiarities of experimental
ment with experiment for the potentials of the fairly well-
work. This research was supported by the Israel Science
known diatomic species. For the triatomic species, the
Foundation administered by the Israel Academy of Science.
identification and computation of a few key spin–orbit cou-
The Fritz Haber Center is supported by the Minerva Gesell-
pling matrix elements allows us to rationalize some of the
schaft für die Forschung, GmbH München, FRG.
experimental data concerning these far less well understood
species. Thus, the two bands in the absorption spectrum of
the anion are explained in terms of a higher-lying, mostly
1
U. Banin, R. Kosloff, and S. Ruhman, Isr. J. Chem. 33, 141 共1993兲.
1 ⫹
2
U. Banin and S. Ruhman, J. Chem. Phys. 98, 4391 共1993兲.
⌺ u state, to which optical transition from the ground state is 3
B. J. Greenblatt, Ph.D. thesis, University of California at Berkeley, 1999.
strongly allowed, and a lower-lying 3 ⌸ u state, to which tran- 4
G. Ashkenazi, U. Banin, A. Bartana, R. Kosloff, and S. Ruhman, Adv.
sitions are formally forbidden, but which gains intensity Chem. Phys. 100, 229 共1997兲.
5
J. C. Roy, W. H. Hanillm, and R. R. Williams, Jr., J. Am. Chem. Soc. 77,
from its substantial spin–orbit coupling-induced mixing with
2953 共1955兲.
the 1 ⌺ ⫹
u state. For neutral I3 , a simple model of the SOC 6
L. I. Grossweiner and M. S. Matheson, J. Phys. Chem. 61, 1089 共1957兲.
leads to the ground state being predicted to be a weakly 7
A. Barkatt and M. Ottolenghi, Mol. Photochem. 6, 253 共1974兲.
bound, centrosymmetric and linear 2 ⌸ u,3/2 state, despite the
8
W. Gabes and H. Gerding, J. Mol. Struct. 14, 267 共1972兲.
9
W. Gabes and D. J. Stufkens, Spectrochim. Acta A 30, 1835 共1974兲.
overall MRCI minimum being a bent structure, due to better 10
T. Okada, J. Phys. Soc. Jpn. 50, 582 共1981兲.
spin–orbit stabilization at the linear geometry. This is in 11
T. Okada and J. Hata, Mol. Phys. 43, 1151 共1981兲.
agreement with experiment.38 The energies of the lowest- 12
M. Mizuno, J. Tanaka, and I. Harada, J. Phys. Chem. 85, 1789 共1981兲.
lying excited electronic states are also predicted using the
13
L. E. Topol, Inorg. Chem. 10, 736 共1971兲.
14
W. Gabes and M. A. M. Nijman-Messter, Inorg. Chem. 12, 589 共1973兲.
simple SOC model, and are found to be in excellent agree- 15
H. L. Friedman, J. Chem. Phys. 21, 319 共1953兲.
ment with experiment.39 16
P. W. Tasker, Mol. Phys. 180, 1 共1994兲.
The diatomics in molecules model for electronic struc-
17
H. Isci and M. R. Mason, Inorg. Chem. 24, 271 共1985兲.
18
U. Banin, A. Bartana, S. Ruhman, and R. Kosloff, J. Chem. Phys. 101,
ture has gained considerable popularity. Since it supplies a
8461 共1994兲.
wave functionlike description of both the ground and excited 19
E. Gershgoren, S. Ruhman, J. Vala, and R. Kosloff, J. Phys. Chem. A 共to
potential-energy surfaces, it has been used extensively for be published兲.
simulating nonadiabatic dynamics in clusters and condensed
20
I. Benjamin, U. Banin, and S. Ruhman, J. Chem. Phys. 98, 8337 共1993兲.
21
E. Gershgoren, U. Banin, and S. Ruhman, J. Phys. Chem. A 102, 9 共1998兲.
phases. In the development of the DIM model for I⫺ 3 we had 22
Z. Wang, T. Wasserman, E. Gershgoren, J. Vala, R. Kosloff, and S.
two goals. The first was to account for the spin–orbit cou- Ruhman, Chem. Phys. Lett. 313, 155 共1999兲.
pling between all electronic states, thus having a model of 23
U. Banin, R. Kosloff, and S. Ruhman, Chem. Phys. 183, 289 共1994兲.
the excited states which would enable us to follow the maze
24
T. Kühne and Vöhringer, J. Chem. Phys. 105, 10788 共1996兲.
25
T. Kühne and Vöhringer, J. Phys. Chem. A 102, 4177 共1998兲.
of potentials leading from I⫺ 3 to its photoproducts. Second, 26
T. Kühne, R. Küster, and Voḧringer, Chem. Phys. 233, 161 共1998兲.
once a DIM model was established, it could be used as a fast 27
A. E. Johnson and A. B. Myers, J. Chem. Phys. 102, 3519 共1995兲.
interpolation scheme to obtain potential-energy surfaces on 28
A. E. Johnson and A. B. Myers, J. Phys. Chem. 100, 7778 共1996兲.
which to run the dynamics of the photoreaction. An addi-
29
A. E. Johnson and A. B. Myers, J. Chem. Phys. 104, 2497 共1996兲.
30
B. J. Greenblatt, M. T. Zanni, and D. M. Neumark, Chem. Phys. Lett. 258,
tional benefit of the DIM model would be the ability to in- 523 共1996兲.
clude in the model solvation effects. However, comparison 31
M. T. Zanni, T. R. Taylor, B. J. Greenblatt, B. Soep, and D. M. Neumark,
with MRCI computations shows the DIM method to perform J. Chem. Phys. 107, 7613 共1997兲.
32
M. T. Zanni, V. S. Batista, B. J. Greenblatt, W. H. Miller, and D. M.
fairly poorly in this case. This suggests that any empirical
Neumark, J. Chem. Phys. 110, 3748 共1999兲.
DIM model for this system is likely to be of questionable 33
V. S. Batista, M. T. Zanni, B. J. Greenblatt, D. M. Neumark, and W. H.
quality. This failure seems to be due to the inability to treat Miller, J. Chem. Phys. 110, 3748 共1999兲.
J. Chem. Phys., Vol. 114, No. 17, 1 May 2001 Potentials for I2⫺ , I2 , I3⫺ , and I3 7423

34
M. T. Zanni, A. V. Davis, C. Frishkorn, M. Elhanine, and D. M. Neumark, 59
P. E. Maslen, J. Faeder, and R. Parson, Chem. Phys. Lett. 263, 63 共1996兲.
J. Chem. Phys. 112, 8847 共2000兲; 113, 8854 共2000兲. 60
S. B. Sharp and G. I. Gellene, Mol. Phys. 98, 667 共2000兲.
35
M. T. Zanni, B. J. Greenblatt, A. V. Davis, and D. M. Neumark, Proc. 61
R. S. Mulliken, J. Chem. Phys. 55, 288 共1971兲.
SPIE 3271, 196 共1998兲. 62
P. Schwerdtfeger, L. v. Szentpály, K. Vogel, H. Silberbach, H. Stoll, and
36
M. T. Zanni, B. J. Greenblatt, A. V. Davis, and D. M. Neumark, J. Chem. H. Preuss, J. Chem. Phys. 84, 1606 共1986兲.
Phys. 111, 2991 共1999兲. 63
J. Li and K. Balasubramanian, J. Mol. Spectrosc. 138, 162 共1989兲.
37
H. Choi, R. T. Bise, A. A. Hoops, and D. M. Neumark, J. Chem. Phys. 64
C. Teichteil and M. Pelissier, Chem. Phys. 180, 1 共1994兲.
113, 2255 共2000兲. 65
38
T. R. Taylor, K. R. Asmis, M. T. Zanni, and D. M. Neumark, J. Chem. G. A. Landrum, N. Goldberg, and R. Hoffmann, J. Chem. Soc., Dalton
Phys. 110, 7607 共1999兲. Trans. 1997, 3605.
39
H. Choi, T. R. Taylor, R. T. Bise, A. A. Hoops, and D. M. Neumark, J.
66
S. B. Sharp and G. I. Gellene, J. Phys. Chem. A 101, 2192 共1997兲.
67
Chem. Phys. 113, 8608 共2000兲. R. M. Lynden-Bell, R. Kosloff, S. Ruhman, D. Danovich, and J. Vala, J.
40
M. Dolg, Mol. Phys. 88, 1645 共1996兲. Chem. Phys. 109, 9928 共1998兲.
41 68
W. A. de Jong, L. Visscher, and W. C. Nieuwpoort, J. Chem. Phys. 107, A. Sanov, T. Sanford, L. J. Butler, J. Vala, R. Kosloff, and W. C.
9046 共1997兲. Lineberger, J. Phys. Chem. A 103, 10244 共1999兲.
42
MOLPRO is a package of ab initio programs written by H.-J. Werner and P. 69
When very diffuse s and p primitives ( ␣ ⫽0.015) are added to the basis
J. Knowles, with contributions from R. D. Amos, A. Berning, D. L. set, some of the higher-lying excited states are shown to be indeed reso-
Cooper et al.
43 nances by one of the occupied orbitals becoming very diffuse and con-
M. Dolg, Habilitationsschrift, Universität Stuttgart, 1997.
tinuumlike. These extra-diffuse primitives have a negligible effect on the
44
F. Y. Naumkin, Chem. Phys. 226, 319 共1998兲.
45
F. Y. Naumkin, Chem. Phys. 240, 79 共1999兲. energy of the valence electronic states studied here and were not included
46
J. C. Tully, in Potential Energy Surfaces, edited by K. P. Lawley 共Wiley, in the basis set used to produce the results described in this paper.
70
New York, 1980兲, p. 63. For an example of where ionic states improve the results from DIM, see
47
P. J. Kuntz, Ber. Bunsenges. Phys. Chem. 86, 367 共1982兲. B. L. Grigorenko, A. V. Nemukhin, and V. A. Apkarian, Chem. Phys.
48
D. Horinek and B. Dick, Phys. Chem. Chem. Phys. 1, 2667 共1999兲. 219, 161 共1997兲.
49
V. S. Batista and D. F. Coker, J. Chem. Phys. 105, 4033 共1996兲. 71
A. Viste and P. Pyykkö, Int. J. Quantum Chem. 25, 223 共1984兲.
50
V. S. Batista and D. F. Coker, J. Chem. Phys. 106, 7102 共1997兲. 72
N. N. Gorinchoi, F. Cimpoesu, and I. B. Besuker, J. Mol. Struct.:
51
C. J. Margulis, D. A. Horner, S. Bonella, and D. F. Coker, J. Phys. Chem. THEOCHEM 530, 281 共2000兲.
A 103, 9552 共1999兲. 73
P. Schuster, H. Mikosch, and G. Bauer, J. Chem. Phys. 109, 1833 共1998兲.
52
C. J. Margulis and D. F. Coker, J. Chem. Phys. 113, 6113 共2000兲. 74
A. L. Kaledin, M. C. Heaven, W. G. Lawrence, Q. Cui, J. E. Stevens, and
53
I. H. Gersonde and H. Gabriel, J. Chem. Phys. 98, 2094 共1993兲. K. Morokuma J. Chem. Phys. 108, 2771 共1998兲.
54
The parameters of these diatomic curves, as well as all the other data 75
A. L. Kaledin, M. C. Heaven, K. Morokuma, and D. M. Neumark, Chem.
described here, and the FORTRAN programs used for the DIM models, can
Phys. Lett. 306, 48 共1999兲.
be obtained upon request from JNH. 76
55
D. Danovich, J. Hrušák, and S. Shaik, Chem. Phys. Lett. 233, 249 共1995兲. J. B. Giorgi, F. Y. Naumkin, J. C. Polanyi, S. A. Raspopov, and N. S.-K.
56
M. N. Glukhovtsev, A. Pross, H. B. Schlegel, R. D. Bach, and L. Radom, Sze, J. Chem. Phys. 112, 9569 共2000兲.
J. Am. Chem. Soc. 118, 11258 共1996兲.
77
K. P. Huber and G. Herzberg, Constants of Diatomic Molecules 共Van
57
B. Braı̈da and P. C. Hiberty, J. Phys. Chem. A 104, 4618 共2000兲. Nostrand-Reinhold, New York, 1979兲.
58 78
P. W. Tasker, G. G. Balint-Kurti, and R. N. Dixon, Mol. Phys. 32, 1651 K. Do, T. P. Klein, C. A. Pommerening, and L. S. Sunderlin, J. Am. Soc.
共1976兲. Mass Spectrom. 8, 688 共1997兲.

You might also like