You are on page 1of 12

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

pubs.acs.org/JPCA

Insights into the Electronic Structure of Ozone and Sulfur Dioxide


from Generalized Valence Bond Theory: Bonding in O3 and SO2
Tyler Y. Takeshita,† Beth A. Lindquist,† and Thom H. Dunning, Jr.*,†,‡

Department of Chemistry, University of Illinois at Urbana−Champaign 600 S. Mathews Avenue, Urbana, Illinois 61801, United
States

ABSTRACT: There are many well-known differences in the physical and chemical
properties of ozone (O3) and sulfur dioxide (SO2). O3 has longer and weaker bonds than
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

O2, whereas SO2 has shorter and stronger bonds than SO. The O−O2 bond is dramatically
weaker than the O−SO bond, and the singlet−triplet gap in SO2 is more than double that
in O3. In addition, O3 is a very reactive species, while SO2 is far less so. These disparities
have been attributed to variations in the amount of diradical character in the two
molecules. In this work, we use generalized valence bond (GVB) theory to characterize the
Downloaded via 106.198.37.32 on February 19, 2020 at 09:26:29 (UTC).

electronic structure of ozone and sulfur dioxide, showing O3 does indeed possess
significant diradical character, whereas SO2 is effectively a closed shell molecule. The GVB
results provide critical insights into the genesis of the observed difference in these two
isoelectronic species. SO2 possesses a recoupled pair bond dyad in the a″(π) system,
resulting in SO double bonds. The π system of O3, on the other hand, has a lone pair on
the central oxygen atom plus a pair of electrons in orbitals on the terminal oxygen atoms
that give rise to a relatively weak π interaction.

1. INTRODUCTION the diradical character of O3 in contrast with the closed-shell


There are major differences in both the energetics and nature of SO2.
Recently, Glezakou and co-workers examined the electronic
reactivities of the ozone and sulfur dioxide molecules. For
structure of various sulfur−oxygen triatomic molecules.6 Their
example, O3 has longer and weaker bonds than O2, 1.278 Å
analysis supported this conventional wisdom, with ozone
versus 1.207 Å and 70 kcal/mol versus 118 kcal/mol,1,2 whereas
containing a non-negligible degree of diradical character while
SO2 has shorter and stronger bonds than SO, 1.432 Å versus
SO2 was overwhelmingly closed shell. Moreover, Miliordos and
1.493 Å and 127 kcal/mol versus 125 kcal/mol.1,2 In addition,
Xantheas13 recently showed that the degree of diradical
the O−O2 bond is dramatically weaker than the O−SO bond,
character from their calculations correlated very well with
23 kcal/mol versus 129 kcal/mol,2 and the singlet−triplet gap
molecular properties (bond lengths and energetics). But, why
in SO2 is more than two times that in O3, 73 kcal/mol versus
are these two valence isoelectronic species so different with
34 kcal/mol.3 Finally, O3 is a very reactive species, while SO2 is
respect to this important molecular property?
far less so. For example, O3 reacts rapidly with carbon−carbon
The previous theoretical studies of O3 and SO2 focused
multiple bonds, but SO2 is totally unreactive.3 The first two
mainly on the electronic structure of the molecules near their
points will be investigated in this paper, and the third will be equilibrium geometries, with the exception of ref 13, which
investigated in a following paper, where we consider the reported multistate potential energy scans for a variety of
addition of hydrogen atoms to O3 and SO2. triatomic sulfur-oxide molecules. In the current study, we
The underlying rationale for the above discrepancies is not analyze the bonding in both O3 and SO2 over a range of
unanimously agreed upon. The dramatic divergent properties of geometries describing formation of the O′−XO (X = O, S)
O3 and SO2 suggest different bonding motifs between the two bond using the generalized valence bond (GVB) method.14,15
molecules. However, why this should be is not immediately The GVB wave function enables us to explicitly connect the
obvious given that they are isoelectronic species. While the SO electronic structure of the parent diatomic molecule (O2 or
bonds in SO2 are often depicted in general chemistry textbooks SO) and the incoming O atom to that of the triatomic molecule
as double bonds on the basis of their length and strength,4 the (O3, SO2). In a recent article,16 we used a similar strategy to
bonding scheme that gives rise to this pattern is unclear, as the
d-orbitals of the sulfur atom are too high in energy to
meaningfully participate in bonding in SO2.4 To address this Special Issue: 100 Years of Combustion Kinetics at Argonne: A
Festschrift for Lawrence B. Harding, Joe V. Michael, and Albert F.
dilemma, various studies comparing the electronic structure of
Wagner
O3 and SO2 using both molecular orbital (MO) and valence
bond (VB) theory have been carried out.5−12 Generally (but Received: January 30, 2015
not always7), the unusual reactivity of O3 relative to SO2, as Revised: May 13, 2015
well as the difference in the singlet to triplet gap, is attributed to Published: June 11, 2015

© 2015 American Chemical Society 7683 DOI: 10.1021/acs.jpca.5b00998


J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

show that the π systems in the HSO and SO2 molecules, but functions for the electrons in the na active orbitals for a state
not in SOH and SOO, possess recoupled pair bonds.17,18 This designated by spin quantum numbers, S and M. The
finding provided a straightforward explanation for the antisymmetrizer, (̂ , ensures that the wave function is
remarkable differences in the geometries and energetics of antisymmetric with respect to the interchange of electrons.
these molecules (e.g., the weaker SO bond strength and longer As a molecule forms, the GVB orbitals and spin-coupling
SO bond distance in SOH versus HSO). These calculations coefficients (cS;k) change from those appropriate for the
showed that SO2 has a double-bonded structure, while its separated species to those appropriate for the molecule. Of
structural isomer, SOO, does not. The double bond character particular interest in this work is the perfect pairing (PP) spin
of the SO bonds arises from the ability of the sulfur atom, but function, which couples the electrons in successive orbital pairs
not the oxygen atom, to form recoupled pair π bonds.16,19,20 into singlets. That is, for a singlet, the electrons in the active
This allows sulfur to formally “expand its octet” without orbital pairs (φ2i+1φ2i+2), i = 0, 1, ..., na/2 are multiplied by the
invoking 3d atomic orbitals as valence orbitals. spin function 1/√2(αβ − βα) with the GVB wave function
The remainder of the paper is organized as follows. The being a product of many such pairs. We find that the PP spin
computational methodology used in this work is described in function is often dominant for a range of geometries near the
Section 2. In Section 3, we discuss the differences in the equilibrium geometries. This simplifies the interpretation of the
bonding motif of the ground states (1A1) of O3 and SO2 by wave function because the overlaps of GVB orbitals that are
adding an O atom to the ground states (3Σ−) of O2 and SO. singlet coupled in the PP wave function are energetically
This is followed by an analysis of the impact of the two bonding favorable, while the overlaps of GVB orbitals that are not singlet
motifs on the singlet−triplet (1A1 → 3B2) gap. We conclude in coupled represent energetically unfavorable interactions as a
Section 4. result of the Pauli Principle (Pauli Repulsion).21 A detailed
analyses of GVB theory and the behavior of GVB wave
2. METHODOLOGY functions during the formation of σ and π recoupled pair bonds
Our approach combines accurate calculations of the structure may be found in refs 17−20.
and energies of the molecules of interest using state of the art 2.2. GVB Orbital Diagrams. In this work, we schematically
coupled cluster and multireference configuration interaction represent the GVB wave function using simple orbital diagrams.
methods followed by generalized valence bond (GVB) Diagrams for the sulfur and the oxygen atom are shown in
calculations to gain insights into the critical differences between Figure 1a. The guidelines for interpreting this diagram are:
the electronic structures of the molecules of interest. (1) The in-plane lobes represent the px or pz orbitals and the
2.1. GVB Theory. Valence bond (VB) methods are circle represents the out-of-plane py orbital.
appropriate for the study of chemical bond formation because (2) The dots in these orbitals represent their electron
they can describe the transition between the molecule and its occupations.
constituent fragments. In the traditional formulation of VB (3) The electrons in the colored orbitals are coupled into
theory, the spatial orbitals are fixed as atomic orbitals. This high spin in the dominant spin function. In Figure 1a, the
constraint requires a substantial number of VB structures to be two electrons in the py and pz orbitals are triplet coupled.
included in the calculation to accurately describe molecular
(4) The S atom lone pair is depicted as one lobe with two
formation, which greatly complicates the interpretation of the
dots (electrons) separated by dashed lines and the O
wave function. The GVB method lifts this constraint by
atom lone pair a contracted and diffuse pair (see the
optimizing the orbitals as well as the spin coupling coefficients
at each nuclear configuration.14,15 As a result, the GVB wave Discussion below). In Hartree−Fock (HF) theory, both
function provides a realistic representation of the electronic of these lone pairs correspond to doubly occupied px2
structure of the molecule that connects smoothly with that of orbitals as is illustrated below the two GVB diagrams.
the fragments and can be interpreted using concepts familiar to (5) A line connecting two lobes/circles on different atoms
most chemists. indicates that these orbitals are singlet coupled into a
The GVB wave function consists of three sets of orbitals: (i) bond pair.
the doubly occupied core orbitals, {ϕci}, (ii) the doubly While not depicted in the diagrams, the electrons in both the
occupied valence orbitals, {ϕvi}, and (iii) the singly occupied doubly occupied valence s orbitals and the core orbitals are
active valence orbitals, {φai}. The active valence orbitals are, in included in the calculations.
general, nonorthogonal to one another. However, the doubly The GVB diagram for the S(3P) atom differs from the HF
occupied orbitals are orthogonal to one another as well as to representation in a very important way: the S 3px2 lone pair is
the singly occupied orbitals. The general form of the GVB wave described by a singlet-coupled pair of GVB orbitals, φS3px− and
function (un-normalized) is shown below. φS3px+ with an overlap of 0.88 (in the HF wave function S ≡ 1).
Calculations on the sulfur atom find that the S 3px2 lone pair is
ΨGVB = (̂ ϕc1ϕc1 ··· ϕcn ϕcn ϕv1ϕv1 ··· ϕvn ϕvn φa1φa2 ··· φan αβ best described by mixing S 3dxz character into the S 3p orbitals,
c c v v a
introducing angular or “left−right,” correlation into the orbitals
··· αβαβ ··· αβ ΘnSa, M (1) involved in the lone pair; see Figure 1b.22 (Note that this 3dxz
function is not the sulfur atomic 3d orbital; its size is
The GVB spatial wave function is a product of (nc+nv) spatial
comparable to that of the 3p orbital, not that of the more
functions, {ϕciϕci, ϕviϕvi}, which represent the doubly occupied
diffuse atomic 3d orbital.) We have shown in earlier work that
core and valence orbitals, and na singly occupied active orbitals,
using a GVB orbital pair to describe the S 3p2 lone pair
{φai}. This spatial function is multiplied by a spin function that
provides new insights into sulfur−oxygen bond formation (i.e.,
is a product of (nc + nv) αβ pairs and a general spin function for
f na
recoupled pair bonds).16 By contrast, in the oxygen atom, the
the electrons in the active orbitals, ΘnS,M a
= ∑k s cS;kΘnS,M;k
a
. The 2p orbital mixes with a small amount of the 3p orbital to form
na
latter is a linear combination of all f S linearly independent spin the GVB lobes (again, not the oxygen atomic 3p orbital). These
7684 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

parameters were optimized with the MRCI+Q wave function


using the aug-cc-pVTZ (oxygen) and aug-cc-pV(T+d)Z
(sulfur) basis sets.25−27 The general notation AVTZ and
AVQZ will be used to denote these basis sets.
The GVB calculations were performed with the fully
variational CASVB program written by Cooper, Thorsteinsson,
and co-workers36−38 with the Kotani spin basis.39 (The full
GVB wave function and the spin-coupled valence bond
(SCVB) wave function40,41 are identical.) In general, the only
constraint placed on the wave function was that orbitals of
different symmetries (a′, a″) were forced to be orthogonal. For
the calculations on O3(3B2), the orbitals localized on the
terminal oxygen atoms were optimized sequentially until self-
consistency was reached to allow them to retain their atomic-
like character rather than mixing to become orthogonal
bonding and antibonding combinations. This procedure
resulted in a GVB energy that was energetically equivalent
(within 0.01 kcal/mol) to the unconstrained GVB calculation.

3. RESULTS AND DISCUSSION


One of the most peculiar differences between ozone and sulfur
dioxide is the trend in the bond lengths and strengths when
compared to their parent diatomic species. In O3, the O−O
bonds are much weaker and longer than in O2, but in SO2, the
S−O bonds are shorter and slightly stronger than in SO. In
order to understand this observation, as well as the weakness of
the O−O2 bond relative to the O−SO bond, we examined the
two addition pathways: O(3P) + O2(3Σ−) → O3(1A1) and
O(3P) + SO(3Σ−) → SO2(1A1).
Both O2 and SO contain a covalent σ bond. However, their
calculated bond lengths (1.208 and 1.486 Å) and bond
strengths (118.0 and 123.2 kcal/mol) are much shorter and
stronger than would be expected for a single bond. The
Figure 1. (a) GVB and HF orbital diagrams for S(3P) and O(3P), (b) delocalization of the doubly occupied π orbitals coupled with
the angularly correlated sulfur 3p2 GVB orbitals, and (c) radially the favorable triplet coupling between the electrons in the
correlated oxygen 2p2 GVB orbitals. Basis: AVQZ. Throughout, yellow singly occupied π orbitals in O2 and SO have been invoked to
spheres denote sulfur and red spheres, oxygen. account for the very strong bonds in O2(3Σ−) and SO(3Σ−).
For reference, with the same computational methodology, the
lobes are radially or “in−out” correlated, as both lobes are bond lengths of HOOH and HSOH, in which these effects are
centered on the oxygen atom, with one being more diffuse and missing, are 1.453 and 1.667 Å, respectively, and the
the other being tighter (see Figure 1c). Prior work has showed corresponding bond energies are 53.8 and 73.4 kcal/mol,
that it is far more difficult to form recoupled pair bonds with RCCSD(T)/AVQZ.
the 2p lone pair of oxygen than the 3p lone pair of sulfur.19,20,23 We will show that the propensity of the S(3p−, 3p+) lone pair
Therefore, we depict the oxygen lone pair as a doubly occupied lobe orbitals to form a recoupled pair bond and the resulting
orbital in the molecular GVB diagrams, as in the HF orbital propensity of the singly occupied π (a″) orbitals of SO, but not
diagram (Figure 1a), although we will use two GVB orbitals to of O2, to form recoupled pair bond dyads is responsible for the
describe the oxygen lone pair in our calculations. The resulting remarkable variations in the bond lengths and strengths in the
analysis shows that representing the oxygen lone pair as triatomic molecules. In a previous paper, we demonstrated the
essentially doubly occupied is appropriate. presence of a (veiled) recoupled pair π bond in SO:16 as the
2.3. Computational Considerations. All calculations sulfur and oxygen atoms approach one another, the electron in
were performed with the Molpro suite of quantum chemical the incoming singly occupied O 2p1 π orbital forms a π bond
programs.24 The augmented correlation consistent basis set of with an electron from the S(3p−, 3p+) lone pair, leaving an
quadruple-ζ quality (aug-cc-pVQZ) was used for oxygen;25 for unpaired electron in an orbital largely localized on the sulfur
sulfur, the d-augmented quadruple-ζ set,26,27 aug-cc-pV(Q atom (i.e., a recoupled pair π bond). However, the orthogonal π
+d)Z, was used. Coupled cluster theory with perturbative system, which has an unpaired electron in a S 3p1 orbital, does
triples28−31 [CCSD(T) and RCCSD(T)] was used to optimize not form a recoupled pair π bond with the O(2p−, 2p+) lone
geometries and calculate bond energies. The potential energy pair, leaving the O 2p-like lone pair singlet coupled and the
scans for O + XO → XO2 (X = O, S) were computed with singly occupied S 3p-like orbital largely localized on sulfur. This
multireference configuration interaction32,33 wave functions bonding scheme is depicted in the GVB orbital diagram for the
with the Davidson correction34,35 [MRCI+Q] using the same SO(3Σ−) fragment in Figure 2a for the O′ + SO pathway with
active space as used for the GVB calculations at Re as described the orange-shaded orbitals on the sulfur atom being triplet
in Section 3 (four a′-symmetric orbitals and four a″-symmetric coupled. As can be seen, the coupling pattern of the GVB lobe
orbitals). For each point in the scan, all other geometric orbitals on the sulfur atom in SO has changed significantly
7685 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

compared to that in Figure 1a: in the S atom the S(3px−, 3px+)


orbitals are singlet coupled, but in the SO molecule the
S3p−(a″)-like orbital is triplet coupled with the sulfur 3p(a′)-
like orbital in the molecular plane. In the previous studies, we
found that the presence or absence of this recoupled pair π
bond explained the large variations in S−O bond strength and
length in both the HSO/SOH and OSO/SOO isomers. In the
following sections, we will show in detail how the π orbitals of
SO are modified by formation of SO2, and then we will analyze
O2 and O3 in a similar way.
3.1. O′ + SO → SO2. In the SO(3Σ−) state, the net result of
the formation of the recoupled pair π bond is that both singly
occupied orbitals are centered on the sulfur atom. This orbital
configuration enables the incoming oxygen atom (O′) to form
a π bond as well as a σ bond with the S atom in the SO
molecule. In Figure 3, the GVB orbitals for the O′(3P) +
SO(3Σ−) addition are plotted as a function of ΔR = R(O′−SO)
− Re(SO2); Re(SO2) = 1.437 Å. The GVB orbitals in Figure 2b
and 3 are labeled consistently for clarity. All six orbitals are
included in the active space [the O′−S σ bonding orbitals, (φ1,
φ2), and the S−O and O′−S π orbitals, (φ5, φ6) and (φ3, φ4)]
and, at Re, where the SO bond lengths are equivalent, both σ
bond pairs are included. At large distances, ΔR(O′−SO) = 1.6
Å, the GVB orbitals of a″(π) symmetry describe an SO bond
pair involving the S 3p+-like sulfur lobe orbital (φ4) and an O
2p1-like orbital (φ3), with φ4 polarized toward and delocalized
onto the oxygen atom, resulting in a very polar π bond. The
remaining GVB a″ orbitals correspond to the S 3p−-like sulfur
Figure 2. (a) GVB diagram for the O′(3P) + SO(3Σ−) → SO2(1A1) lobe orbital (φ5) and the incoming O′2p1 orbital (φ6). The spin
formation pathway. Shading of the same color in the lobes indicates function that singlet couples the spins of the electrons in the
triplet coupling between the orbitals at R(OS−O′) = ∞, while the SO π bond pair (φ3, φ4) and triplet couples those in the two
dashed lines represent bond formation as R(O′−SO) decreases. (b)
GVB diagram for SO2(1A1) at Re.
orbitals remaining on the SO fragment (φ2, φ5) as well as those
on the O′ atom (φ1, φ6) constituents 96.3% of the wave

Figure 3. GVB orbitals for O′(3P) + SO(3Σ−) → SO2(1A1), where ΔR = R(O′−SO) − Re(SO2); Re(SO2) = 1.445 Å. All other degrees of freedom
are optimized. Dashed lines separate orbital pairs that are singlet coupled by the PP spin function.

7686 DOI: 10.1021/acs.jpca.5b00998


J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

function. This is shown in Figure 2a and describes the O(3P) = 0.64 and S35 = 0.21, which represent antibonding interactions
and SO(3Σ−) fragments. (The orbital labeling scheme in Figure in the GVB wave function of the isolated SO fragment.
2, panels a and b, are identical.) This coupling pattern can be As the internuclear distance decreases, φ2 localizes on the
schematically written as αβααββ when the GVB orbitals in sulfur atom and polarizes into the O′−SO bonding region.
Figure 3 are ordered in the sequence (φ3, φ4, φ2, φ5, φ1, φ6) in Concomitantly, the spin of the electron in this orbital becomes
the GVB spatial orbital product. In the SO2 molecule (Figure increasingly singlet coupled with that of the electron in φ1,
2b), the dominant spin function is the perfect pairing (PP) spin representing the formation of an SO σ bond (see Figure 4).
function, αβαβαβ, with the orbital ordering (φ1, φ2, φ3, φ4, φ5, Because φ5 is localized on the sulfur atom at large ΔR(O′−
φ6). As R(O′−SO) decreases, the weight of the αβααββ SO), a π bond involving φ5 and φ6, the singly occupied π(a″)
configuration decreases and that for the PP coupling grows (see orbital on the incoming O atom, can also form, imparting
Figure 4). The PP spin function accounts for 93.9% of the wave double bonded character to both SO bonds in SO2. Therefore,
transition of the spin coupling coefficients from a recoupled
pair π bond plus two triplet pairs to a perfectly paired spin
function represents formation of a second O′−SO σ bond
(involving φ1 and φ2), and a second π bond (involving φ5 and
φ6). In addition, the original recoupled pair π bond, (φ3, φ4), as
well as the other SO σ bond (not shown) is maintained,
without any qualitative changes.
The overlaps of the a″ GVB orbitals are plotted as a function
of R(O′−SO) in Figure 5, where blue lines indicate the

Figure 4. GVB spin coupling weights of O′−SO as a function of


R(O′−SO) for O′(3P) + SO(3Σ−) → SO2(1A1).

function at Re. The switch in the dominant spin function


reflects the transition from the diatomic fragments to the
bonded SO2 molecule (i.e., changing from the coupling pattern
in Figure 2a to that in Figure 2b). Note that these two spin
functions are not orthogonal because of the different ordering
of the orbitals and, therefore, the sum of the squares of their
Figure 5. Overlaps of the GVB a″(π) orbitals for O′(3P) + SO(3Σ−)
coefficients do not sum to unity. However, the GVB energy is → SO2(1A1) as a function of R(O′−SO).
unaffected by any permutation of the spatial orbitals as long as
all of the spin functions are included in the calculations (as is
done here).
As the SO2 molecule forms, the GVB orbitals, as well as the energetically favorable overlaps between the orbitals of a
spin coupling coefficients, change in response to the approach singlet-coupled pair of electrons in the PP wave function and
of the O′ atom. In Figure 3, dashed lines separate the orbitals red lines indicate energetically unfavorable overlaps due to
that are singlet coupled in the PP wave function for SO2 (even exchange-repulsion between the electrons in the orbitals in
if they are not predominantly singlet coupled at the given ΔR). different singlet coupled pairs.17,18,21 As the second σ bond
At large internuclear distances, the orbitals of the SO fragment forms, the overlap between φ5 and φ6 (S56), representing the
are not significantly perturbed by the presence of the O′ atom second π bond, increases monotonically to a value of 0.70 at Re.
and the orbitals in Figure 3 at ΔR = 1.6 Å are essentially the Moreover, the energetically unfavorable overlaps that were
orbitals of O(3P), (φ1, φ6), and SO(3Σ−), (φ2, φ3, φ4, φ5). In present in SO are decreased by bond formation as φ5 is
SO(3Σ−), φ2 and φ5 would be equivalent by symmetry, though polarized toward the incoming oxygen atom, O′ (see S35 and
they are slightly different in this calculation due to the choice of S45 in Figure 5). These two effects, coupled with the lack of any
active space. At ΔR = 1.6 Å, φ2 has antibonding character substantial additional unfavorable overlaps resulting from bond
because it is orthogonal to the doubly occupied 2pa′ lone pair formation (i.e., S36 and S46 are small), result in the SO bonds in
on O, ϕv1, in Figure 2. By contrast, all of the valence a″ orbitals SO2 being stronger and shorter than that of diatomic SO. The
in Figure 3 are included in the active space, so φ5 does not have polarity of the SO π bonds due to the difference in
to be orthogonal to the a″ orbitals of the sulfur and oxygen electronegativity between the central sulfur atom and the
atoms and, thus, does not possess antibonding character. terminal oxygen atoms aids in reducing the energetically
Rather, φ5 is entirely localized on the sulfur atom but with large unfavorable overlap between the orbitals involved in the bond
energetically unfavorable overlaps with the other a″ orbitals: S45 pairs. It is the stability of the two π bonds (a recoupled pair π
7687 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

bond dyad16,19) in SO2 that accounts for its essentially the orbital ordering (φ4, φ5, φ3, φ2, φ1, φ6) relative to Figures 6
complete lack of diradical character. and 8. As Figure 8 shows, this coupling pattern accounts for
3.2. O′ + O2 → O3. We now turn to the same analysis for 98.2% of the wave function at R(O′−OO) = Re + 2.0 Å. This is
O3. Figure 6 shows the O′(3P) + O2(3Σ−) → O3(1A1) in contrast to O′ + SO, where there was a clear π bond pair on
the SO fragment and is a consequence of the lack of recoupled
pair bonding in the O2 fragment. The lack of a recoupled pair π
bond is consistent with prior studies that have shown that it is
much less energetically favorable to form a recoupled pair bond
with an oxygen 2p2 pair than with a sulfur 3p2 pair.19,20,23
The absence of recoupled pair π bonding in O2(3Σ−) has
important consequences for the electronic structure of O3(1A1).
As O′ approaches O2 in Figure 7, φ2 localizes onto the central
oxygen atom in order to form the σ bond. This breaks the g and
u symmetry of the GVB a″ orbitals, localizing φ3 and φ4, the
“in−out” correlated oxygen (2p−, 2p+) orbitals, on the central
oxygen atom. At intermediate ΔR, the “out” correlated a″
orbital, φ3, maintains a π-bonding shape over O2, but as R(O′−
OO) approaches Re, this orbital localizes on the center oxygen
atom but with significant delocalization onto the two terminal
oxygen atoms. The addition of the O′ atom results in a new σ
bond (φ1, φ2). However, formation of a π bond is not observed
as ΔR approaches zero because the unpaired a″(π) orbital, φ5,
is centered on the terminal, not the central, oxygen atom of the
O2 fragment. Although this diradical configuration clearly
dominates at all values near Re, previous work by Cooper et
al.38,42 have shown that, at the equilibrium geometry, there
exists a GVB wave function of nearly equal energy with a π
bonding pattern similar to SO2. This alternative GVB wave
function and its relation to the GVB wave function described in
this section are discussed further in the Appendix. Here, we
focus on the wave function depicted in Figure 6 as it provides a
consistent description of the formation of O3 from O + O2.
As with SO2, the coefficient of the spin-coupling pattern
associated with the O′ + O2 fragments decreases as a function
Figure 6. (a) GVB diagram for the O′(3P) + O2(3Σ−) → O3(1A1)
formation pathway, where the spin coupling pattern of O2 has been of R(O′−OO), crossing that for the PP spin function, the latter
simplified (see text for full description). Shading of the same color in comprises 95.0% of the GVB wave function of O3 at Re. These
the lobes indicates triplet coupling between the orbitals at R(OO−O′) two coupling patterns are depicted in the GVB orbital diagrams
= ∞, while the dashed lines represent bond formation as R(O′−OO) in Figure 6 (panels a and b), where the growth in the weight of
decreases. (b) GVB diagram for O3(1A1) at Re. the PP pairing spin function indicates formation of the new
O′O σ bond and the new, but weak, O′O π interaction between
formation pathway diagrams where the orbital labeling scheme φ5 and φ6.
in Figures 6 and 7, the GVB orbitals as a function of ΔR, are The bonding pattern of ozone results in a different overlap
consistent for clarity. The a″ orbitals localized on the O2 pattern compared to SO2 formation and is shown in Figure 9.
fragment at large R(O′−OO) bear some similarity to those of Instead of two bond pairs, there is one large energetically
SO; compare the top rows of Figures 3 and 7. However, for favorable overlap corresponding to the lone pair centered on
ozone, φ2 and φ5 are effectively identical, despite the differing the central oxygen atom (S34), which decreases slightly with
active spaces for the a′ and a″ orbitals. The antibonding decreasing R(O′−OO) and a weakly overlapping pair localized
character of φ5 in O2 but not SO is not due to any constraint on on the terminal atoms (S56), which increases from 0.0 to 0.16 as
the GVB wave function; rather, this is a consequence of the R(O′−OO) decreases from 3.5 Å to Re. The overlaps
symmetry and bonding in the O2 fragment. Unlike in SO(3Σ−), representing repulsive interactions between the singlet coupled
the O 2p2 pairs in O2(3Σ−) are not recoupled by the incoming pairs (φ3, φ4) and (φ5, φ6) are shown as the red lines in Figure
O 2p1 orbitals but rather are simply delocalizing onto the 9. Compared to SO2, all of the energetically unfavorable
opposite O atom as the MO description of bonding suggests. overlaps increase dramatically as O′−OO bond formation
Because φ5 is essentially orthogonal to φ3 and φ4 at large occurs, particularly S35 and S36. These repulsive interactions and
R(O′−OO), the dominant spin function for O2 at large R is the loss of the favorable exchange interactions in O2 are the
one in which the electrons in the “in-correlated” bonding major factors in the weakening and lengthening of the OO
orbital (φ4) and the antibonding orbital (φ5) are coupled into a bonds from O2 to O3. The lengthening of the OO bonds from
triplet, which then couple with the electron in the “out- 1.208 Å in O2 to 1.269 Å in O3 [RCCSD(T)/AVQZ] results in
correlated” orbital (φ3) to make an overall doublet (this the observed reduction in the S35 and S45 overlaps from about
coupling pattern is signified by the checkered pattern in Figure R(O′−OO) = 1.7 Å to Re (see below).
8a). The electrons in the remaining orbitals are spin coupled as 3.3. Geometry and Energetics Along the O′−XO
required for the O2(3Σ−) and O(3P) fragments to form a net Reaction Coordinate, (X = O, S). In the Introduction, we
singlet state. We express this coupling pattern as ααβαββ with noted that O3 and SO2 have different trends in terms of bond
7688 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

Figure 7. GVB orbitals for O′(3P) + O2(3Σ−) → O3(1A1), where ΔR = R(O′−OO) − Re(O3); Re(O3) = 1.281 Å. All other degrees of freedom are
optimized. Dashed lines separate orbital pairs that are singlet coupled by the PP spin function. Orbitals (φ3, φ4) provide a GVB description of the
oxygen lone pair (see Figure 1a).

Figure 8. Spin coupling weights of O′−OO as a function of R(O′−


OO).
Figure 9. Overlaps of the GVB a″ orbitals for O′(3P) + O2(3Σ−) →
O3(1A1) as a function of R(O′−OO).
length and strength with respect to their parent diatomic
molecules, XO (X = O, S). In this section, we consider these
properties as a function of the O′−XO distance. The calculated Figure 10b shows the change in the bond length of the XO
potential energy curve [MRCI+Q/AVQZ] for formation of fragment as the O′ atom approaches XO for the geometries in
XO2 from O′ and XO as a function of ΔR(O′−XO) is shown in the scan in Figure 10a. This figure is a plot of ΔR(O′X−O)
Figure 10a, where the other molecular parameters (RXO, θO′XO) relative to its value at ΔR(O′−XO) = Re + 2.0 Å, which is
are optimized at each point along the curve [MRCI+Q/AVTZ]. approximately the bond length of diatomic XO. As ΔR(O′−
For SO2, the energy associated with this addition is quite large: XO) decreases, the O′X−O bond length decreases monotoni-
131.0 kcal/mol, compared to the calculated bond energy of SO cally until ΔR(O′−XO) ≈ 0.6 Å in both cases, likely as a result
of 123.2 kcal/mol [RCCSD(T)/AVQZ]. By contrast, for O3, of the reduction in the antibonding character of φ2 in Figures 3
the O′−O2 bond is markedly weaker than that calculated for and 7 as it polarizes toward the O′−XO bonding region. For
O2: 118.0 kcal/mol for O2 versus 23.8 kcal/mol for O′−O2 both formation pathways, this trend reverses around ΔR(O′−
[RCCSD(T)/AVQZ]. XO) ≈ 0.6 Å as the magnitude of the repulsive orbitals overlaps
7689 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

are stronger than that of SO(3Σ−). In SO2, the incoming a″


orbital can form a recoupled pair π bond dyad because the
recoupled pair bond present in the a″ system of SO localizes
both of the singly occupied orbitals on the sulfur atom.
Conversely in O3, the interaction between the electron in the
incoming oxygen 2pa″ orbital and the two singlet-coupled
electrons centered on the central oxygen atom is repulsive. In
addition, this orbital has only a very small overlap with the
orbital with which it is singlet coupled. We will comment on
these results in the context of prior work investigating the
nature of bonding in O3 in Section 4.1. The implications of this
difference in the electronic structure of O3 and SO2 for the
singlet to triplet gap in these two molecules will be considered
in the following section.
3.4. Energetics of the Singlet−triplet Energy Gap of
the 1A1 and 3B2 States of O3 and SO2. The different degrees
of diradical character in O3 and SO2 have a major impact on the
singlet−triplet splitting in these molecules. In O3, only a weakly
interacting singlet coupled pair (S = 0.16) must be broken to
form the corresponding 3B2 triplet state, whereas in SO2, two
much more strongly interacting singlet pairs (S = 0.70) must be
disrupted in order to form such a state. In this section, we
compare the properties of the 3B2 states of O3 and SO2. This
state is, in fact, the lowest triplet state in O3 with the lowest-
lying 3B1 and 3A2 states being 6.8 and 0.7 kcal/mol higher in
energy, respectively. The 3B2 state in SO2 is the third lowest
triplet excited state, although the 3B1 and 3A2 states of SO2 are
only 7.0 and 5.2 kcal/mol lower in energy (in that order) than
the 3B2 state. We calculate the energy splitting between the 1A1
and 3B2 state of O3 to be 29.7 kcal/mol compared to 81.7 kcal/
mol in SO2 [RCCSD(T)/AVQZ]. The triplet states have
longer bond lengths than the ground states as well, by 0.084 Å
in O3 and 0.122 Å in SO2 at the same level of theory. In the
triplet states, the S−O bonds of SO2 are still stronger than the
O−O bonds of O3 (relative to the ground state of the atoms),
Figure 10. (a) Potential energy curves relative to the O′ + XO ground but the difference has been reduced from 112.3 kcal/mol in the
state asymptote and (b) dependence of the O′X−O bond length on singlet states to 60.3 kcal/mol in the triplet states, indicative of
ΔR(O′−XO) = R − Re(XO) for X = O, S; Re(O3) = 1.281 Å and a major loss of bonding in SO2(3B2).
Re(SO2) = 1.445 Å. Calculations: MRCI+Q/AVQZ//MRCI+Q/ The GVB orbitals for the 3B2 states are shown in Figure 11,
AVTZ. panel a (O3) and panel b (SO2). The GVB orbitals for the
O3(3B2) state are qualitatively similar to the singlet ground
between the fragments become nontrivial. For O3, R(O′O−O) state, with the electrons in the lone pair orbitals localized on the
increases monotonically, yielding O−O bonds that have central atom, (φ1, φ2), singlet coupled, and those in the orbitals
lengthened significantly from 1.208 Å in O2 to 1.269 Å in on the terminal atoms, (φ3, φ4), triplet coupled. The latter
O3. However, for SO2, R(O′S−O) also begins to increase but orbitals are now orthogonal instead of weakly overlapping. This
then decreases again. At sufficiently short R(O′−SO), bonding pattern provides an accurate description of the GVB
formation of the O′−SO π bond shortens the O′S−O bond wave function for the π system (wPP = 99.0%). Because the
and reduces the energetically unfavorable overlaps associated terminal orbitals are high spin coupled, their interaction is
with the recoupled pair π bond in SO as shown in Figure 5. repulsive, although the magnitude of the interaction is small
The behavior of the curve between R(O′−SO) = 0.4−0.6 Å because of the spatial separation of the orbitals. In response, the
reflects the transition from repulsion between the O′ + SO O 2p2-like pair on the central oxygen atom contracts: the
fragments to double bond formation and occurs at approx- singlet-coupled pair on the central oxygen atom has a higher
imately the crossing of the two spin couplings plotted in Figure overlap in the O3(3B2) state, 0.85 versus 0.80 in the ground
4. Ultimately, the calculated S−O bond lengths decrease from state, and is more localized on the central oxygen atom than the
1.486 to 1.437 Å [RCCSD(T)/AVQZ] upon O′ addition. corresponding orbitals in the 1A1 state. These effects contribute
In summary, O3 can be described as having a singlet-coupled to the weaker O−O bonds in the 3B2 state than the 1A1 state.
lone pair on its central atom and a pair of electrons in the In contrast, the SO2(3B2) orbitals possess a qualitatively
orbitals on each of the terminal atoms that are coupled together different dominant spin coupling pattern than the SO2(1A1)
into a singlet; the latter representing a very weak π interaction orbitals, as formation of a recoupled pair bond dyad is no
(see the Appendix however). This bonding scheme combined longer possible in the 3B2 state (see Figure 11b). Instead of SO
with the loss of the favorable triplet interaction on O2 leads to double bonds, the PP-like spin coupling pattern (αβαα, 91.4%
O−O bonds in O3 that are substantially weaker than in O2. of the wave function for the π system) has the electrons in the
Conversely, SO2 has two highly polar S−O double bonds that two “left−right” lobe orbitals, (φ1, φ2), coupled into a singlet
7690 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

the overlap of the two orbitals giving rise to the diradical


character to be 0.16. The magnitude of this overlap is small but
is not zero as it would be for a pure diradical. The presence of
non-trivial closed shell character is consistent with the
observation that the singlet state is more than 30 kcal/mol
lower in energy than the triplet state, as the singlet and triplet
states would be degenerate if S ≡ 0 due to the spatial
distribution of the orbitals, i.e. large spatial separation.
Kalemos and Mavridis7 suggested that the character of O3 is
attributable to a dative bonding interaction of the O(1D) atom
with the 1Δg state of O2, and Miliordos and Xantheas13
interpreted their potential energy scans of the 1A′ state of the O
+ O2 interaction as agreeing with this proposal. However, the
GVB calculations for O + O2 do not show any signs of
involvement from this higher lying asymptote, which lies 91.9
kcal/mol7 above the O3(1A1) minimum (MRCI+Q). Rather,
the GVB wave function correlates O3 with the ground state
fragments across the entirety of the potential energy surface
that was scanned and gives rise to the diradical character as a
logical outcome of molecular formation.
Of course, the wave functions arising from the O(3P) +
O2(3Σg−) and O(1D) + O2(1Δg) asymptotes have the same spin
and symmetry and therefore can mix; the question is what is
the magnitude of the mixing. Given the large energy differences
shown in Figure 1 of Miliordos and Xantheas,13 the
Hamiltonian matrix element coupling the two diabatic states
would have to be quite large for there to be substantial mixing
between the two states. Unfortunately, the magnitude of this
matrix element is not known. In its absence, we simply note
that the GVB calculations show a smooth transition from
O(3P) + O2(3Σg−) to O3, suggesting that the state arising from
the O(1D) + O2(1Δg) limit does not have a strong impact on
the ground state orbitals and energetics of O3. It should,
Figure 11. GVB orbitals and orbital overlaps for the (a) O3(3B2) and however, be noted that since we only use one set of orbitals and
(b) SO2(3B2) states. Dashed lines separate orbitals that are singlet the overall GVB wave function must be consistent with the
coupled by the PP spin function. In (a), orbitals (φ1, φ2) provide a symmetry of the molecule, our calculations would not be able
GVB description of the oxygen lone pair (see Figure 1a). to describe the asymmetric bonding pattern associated with
formation of a dative bond at Re. However, even at larger R,
and centered on the sulfur atom with a large overlap of 0.90. where symmetry constraints are not an issue, we see no
The two terminal GVB orbitals, φ3 and φ4, are then triplet evidence of an asymmetric bond associated with the O(1D) +
coupled. The GVB(PP) energy is just 2.8 kcal/mol higher than O2(1Δg) asymptote.
that from the corresponding full GVB calculation using the As we have shown, the GVB description of the formation of
same orbitals, so the GVB(PP) wave function provides a O3 from O(3P) + O2(3Σg−) provides a natural explanation for
reasonable description of the molecule. the diradical character of the ground state (1A1) of O3. In the
It should be noted that the GVB wave function for the O(3P) + O2(3Σg−) limit, Figure 6 shows the singly occupied a″
SO2(3B2) state used here cannot describe a bonding pattern orbital localized on the incoming oxygen atom (O′) with the
with a recoupled pair π bond associated with one SO bond and singly occupied a″ orbital on the O2(3Σg−) molecule being
a triplet-coupled pair associated the other SO bond; this would delocalized over the two oxygen atoms. As R(O′−O2)
require a symmetry projected GVB wave function.43,44 While it decreases, both of these orbitals largely localize on the terminal
is possible that this type of bonding motif might be found if a oxygen atoms with the spins of the electrons in these orbitals
projected GVB wave function was used, it is nonetheless becoming singlet coupled. While the orbitals that are centered
impossible for a new π bond to be formed as R(O′−SO) on the terminal oxygen atoms do not directly overlap very
decreases on the triplet surface. This fundamental change in the strongly with one another, they both interact with the same pair
electronic structure of SO2 accounts for the large singlet−triplet of electrons on the central atom. This is in keeping with the
splitting in the molecule. GVB(PP/SO) calculations of Dunning, Goddard, and Hay,5
where, because the orbitals on the terminal atoms had to be
4. CONCLUSIONS orthogonal to the electron pair on the central atom, they had a
4.1. Why is the Ground State of O3 a Diradical? Our nontrivial overlap with one another (for this restricted wave
calculations are consistent with the conclusion of prior work function, S = 0.28). The full GVB results presented herein are
that concluded that O3(1A1) has significantly more diradical more complicated: since all of the a″ orbitals can overlap, the
character than SO2(1A1). However, O3 is far from a pure favorability of the singlet state is manifest both in the presence
diradical: its diradical character was calculated to be only 44% of the stabilizing overlap between the terminal orbitals (S56) as
or 18% in two recent studies.6,13 In agreement with this, we find well as the reduction in the repulsive overlaps between the
7691 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

terminal orbitals and the central atom orbitals (S35, S36, S45, 4.3. Why are the Singlet−triplet Energy Gaps of the
1
S46), where the numbering is that in Figure 7. However, the A1 and 3B2 States of O3 and SO2 so Different? The two
overall conclusion is the same: singlet coupling the terminal different bonding patterns in SO2 and O3, two π bonds in SO2
orbitals in the π system is energetically favorable. If we use the and one delocalized π pair on the central atom plus a pair of
simpler GVB(PP/SO) wave function where this effect is weakly overlapping π orbitals on the terminal atoms in O3, also
encapsulated in S56, we see that this overlap is in fact directly accounts for the much larger singlet−triplet gap in SO2 versus
related with the 1A′ → 3A′ gap over the O + O2 scan. The O3. For ozone, forming the triplet state only requires breaking
dependence of the 1A′ → 3A′ gap on S56 along the O′−O2 the singlet coupling between the electrons in two weakly
minimum energy curve is plotted in Figure 12 and shows that overlapping orbitals centered on the terminal oxygen atoms (S
these two quantities are nearly linearly related after the PP spin = 0.16). However, in SO2, formation of the triplet state requires
function becomes dominant in the full GVB calculation, which breaking actual π bonds (S = 0.70), and the increased singlet−
occurs around S56 ≈ 0.08. triplet gap in SO2 relative to O3 reflects that. Clearly, the ability
of the sulfur atom to form recoupled pair bonds and recoupled
pair bond dyads with very electronegative ligands is quite
significant and must be included in a comprehensive
explanation of the differences between O3 and SO2, including,
as we shall show in a subsequent paper, differences in their
reactivities.

■ APPENDIX
In an important paper on the electronic structure of the ozone
molecule, Thorsteinsson et al.38 reported GVB calculations for
the four electrons in the π(a″) system of O3 and noted that it is
possible to construct two different wave functions for O3 at its
equilibrium geometry that have nearly the same energy. The
active spaces for these two GVB calculations differ in the nature
of the virtual orbital for correlating the 2pa″-like lone pair on
the central oxygen atom. The first active space, AS1, contains a
3pa″-type virtual orbital largely localized on the central oxygen
atom resulting in the in−out correlated orbital pair (φ3, φ4) as
Figure 12. 1A′ → 3A′ energy gap as a function of the overlap of the shown in Figure 7. The second active space, AS2, replaces the
orbitals localized on the terminal orbitals (S56) in the GVB(SO/PP) 3p-type virtual orbital with a 3d-type virtual orbital resulting in
wave function for R(O′−OO) = [Re − 0.2 Å, Re + 2.0 Å]. S56 and the
singlet−triplet gap (computed with MRCI+Q/AVQZ using state-
two π bonds similar to those in SO2: (φ1, φ2) and (φ3, φ4) in
averaged orbitals) were calculated at the geometries from the scan in Figure 13. The energy of the first GVB wave function, the
Figure 8.

4.2. Why are the Ground States of O3 and SO2 so


Different? In contrast to O3, where a small amount of closed
shell character arises from the weakly overlapping orbitals on
the terminal oxygen atoms, in SO2, the closed shell character is
a result of the formation of a recoupled pair bond dyad in the π
system, effectively resulting in two SO double bonds. These
important differences in the bonding motif associated with the
π system are manifest in the GVB orbitals and spin couplings of
the two species and are a direct consequence of the presence or
absence of a recoupled pair π bond in the parent diatomic
species, SO and O2, respectively. This difference in bonding is
reflected in the vastly different strengths of the O′−SO and
O′−O2 bonds, 131.0 versus 23.8 kcal/mol.
As in our studies of other hypervalent species (e.g., the SFn, n
= 1−6),45 the recoupled pair bond dyad in SO2 allows the Figure 13. O3 (1A1) GVB(AS2) orbitals and orbital overlaps. Orbitals
sulfur atom to form an expanded octet with five nominal that are singlet coupled in the GVB(PP) wave function are separated
electron pairs. However, the SO bonds, especially those of the by dashed lines. Basis: AVQZ.
recoupled pair dyad, are quite polar, reducing electron density
around the sulfur atom. The GVB orbitals clearly show that diradical solution discussed here, was of lower energy, but the
these bonds, while quite polar, are actual π bonds and not O energy difference was found to be just 1.5 kcal/mol. Because of
2p2 lone pairs, in agreement with the earlier SCVB calculations this small energy difference, Thorsteinsson et al. concluded that
of Cooper and co-workers.46 The ability of the sulfur atom to neither wave function “is likely to provide a very satisfactory
form recoupled pair bonds and recoupled pair bond dyads with description of the bonding for this system.”
very electronegative ligands explains how the central sulfur We believe that, in cases such as this, either description may
atom in SO2 can possess two double SO bonds in addition to a be used (compare for example, the σ−π versus bent bond
σ lone pair without appealing to atomic 3d-orbitals. descriptions of unsaturated organic compounds), but that other
7692 DOI: 10.1021/acs.jpca.5b00998
J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

factors that influence the description of the electronic structure


of the molecule must also be taken into account. To better
■ AUTHOR INFORMATION
Corresponding Author
understand the nature of the bonding in O3, we repeated the 4- *E-mail: thdjr@uw.edu, thom.dunning@pnnl.gov.
electron, 4-orbital GVB calculations of Thorsteinsson et al.38
Present Address
and then extended these calculations to include the orbitals ‡
T.H.D., Jr.: Northwest Institute for Advanced Computing,
involved in the σ bonds in the active space (8-electron, 8-orbital
Pacific Northwest National Laboratory, 127 Sieg Hall and
GVB calculations). If only the π orbitals are included in the
Department of Chemistry, University of Washington, Seattle,
GVB calculations, we find that the energy difference,
WA 98195-1700.
EGVB(AS2) − EGVB(AS1), is 1.04 kcal/mol, which is in
reasonable agreement with the result of Thorsteinsson et al.38 Notes
The authors declare no competing financial interest.


given the differences in methods and basis set. The spin-
coupling pattern in these two wave functions is, however, rather
different. The GVB(AS1) wave function couples the spins of ACKNOWLEDGMENTS
the four π electrons as shown in Figure 6b with the perfect This work was supported by the Distinguished Chair for
pairing (PP) weight being equal to unity (by symmetry). For Research Excellence in Chemistry and the National Center for
the GVB(AS2) wave function, the spins of the electrons are Supercomputing Applications at the University of Illinois at
coupled as described in Figure 13 with wPP = 0.91. Thus, Urbana−Champaign. One of the authors (B.A.L.) is the
describing the π system of ozone as having two π bonds is on grateful recipient of a National Science Foundation Graduate
Research Fellowship.


the borderline (although there is no hard-and-fast rule for
making this decision, most well-defined bonding cases have wPP
> 0.9). The overlaps between the GVB(AS2) orbitals are also REFERENCES
instructive. First, the overlap between the two orbitals involved (1) Herzberg, G. Molecular Spectra and Molecular Structure. I. Spectra
in the π bonds is just 0.56. This indicates that the π bonds are of Diatomic Molecules; D. Van Nostrand and Company, Inc.: Princeton,
rather weak. Second, the overlap of the two orbitals that are NJ, 1950.
(2) Herzberg, G. Molecular Spectra and Molecular Structure. III:
largely localized on the central oxygen atom is quite high, 0.63.
Electronic Spectra and Electronic Structure of Polyatomic Molecules; D.
To the extent that the bond pairs are singlet coupled, this Van Nostrand and Company, Inc.: Princeton, NJ, 1966.
overlap represents substantial Pauli repulsion between the two (3) Lan, Y.; Wheeler, S. E.; Houk, K. N. Extraordinary Difference in
π bond pairs, further weakening them. Reactivity of Ozone (OOO) and Sulfur Dioxide (OSO): A Theoretical
We then included the electrons and orbitals involved in the σ Study. J. Chem. Theory Comput. 2011, 7, 2104−2111.
bonds in the calculation. Increasing the GVB active space to (4) Purser, G. H. Lewis Structures Are Models for Predicting
eight orbitals increases ΔEGVB(AS2−AS1) slightly, to 1.39 kcal/ Molecular Structure, Not Electronic Structure. J. Chem. Educ. 1999, 76,
mol. This again suggests that there is little to distinguish the 1013.
two descriptions based on energetics alone. However, (5) Goddard, W. A.; Dunning, T. H.; Hunt, W. J.; Hay, P. J. Hay.
Generalized Valence Bond Description of Bonding in Low-Lying
expansion of the active space decreases wPP in the GVB wave States of Molecules. Acc. Chem. Res. 1973, 6, 368−376.
functions to 0.95 (AS1) and 0.87 (AS2). The magnitude of wPP (6) Glezakou, V.-A.; Elbert, S. T.; Xantheas, S. S.; Ruedenberg, K.
for the GVB(AS1) wave function is consistent with a molecule Analysis of Bonding Patterns in the Valence Isoelectronic Series O3, S3,
that has two σ bonds, a lone pair orbital on the central oxygen SO2, and OS2 In Terms of Oriented Quasi-Atomic Molecular Orbitals.
atom, and a singlet coupled pair of orbitals on the terminal J. Phys. Chem. A 2010, 114, 8923−8931.
oxygen atoms. The much smaller value for wPP for the (7) Kalemos, A.; Mavridis, A. Electronic Structure and Bonding of
GVB(AS2) wave function, on the other hand, implies that the Ozone. J. Chem. Phys. 2008, 129, 054312.
GVB(AS2) wave function is not well-described as having two (8) Miliordos, E.; Ruedenberg, K.; Xantheas, S. S. Unusual Inorganic
Biradicals: A Theoretical Analysis. Angew. Chem., Int. Ed. 2013, 52,
double bonds. The overlaps among the π orbitals from the 8- 5736−5739.
electron, 8-orbital GVB calculation are similar to those from the (9) Braida, B.; Walter, C.; Engels, B.; Hiberty, P. C. A Clear
4-electron, 4-orbital calculation [e.g., 0.54 (bonding orbitals) Correlation Between the Diradical Character of 1,3-Dipoles and Their
and 0.60 (orbitals on the central oxygen atom)]. For these Reactivity Toward Ethylene or Acetylene. J. Am. Chem. Soc. 2010, 132,
reasons, we favor the GVB(AS1) wave function, which, as we 7631−7637.
showed in the paper, provides a consistent description of the (10) Ess, D. H.; Houk, K. N. Theory of 1,3-Dipolar Cycloadditions:
formation of O3 from O + O2 and of the differences between Distortion/Interaction and Frontier Molecular Orbital Models. J. Am.
O3 and SO2. Chem. Soc. 2008, 130, 10187−10198.
(11) DeBlase, A.; Licata, M.; Galbraith, J. M. A Valence Bond Study
The logical next question is the following: What are the of Three-Center Four-Electron π Bonding: Electronegativity vs
corresponding energy differences and spin coupling weights in Electroneutrality. J. Phys. Chem. A 2008, 112, 12806−12811.
SO2? The difference in the GVB energy for AS1 and AS2 in (12) Christodouleas, C.; Xenides, D.; Simos, T. E. Trends of the
SO2 are EGVB(AS2) − EGVB(AS1) = 14.19 kcal/mol. Further, Bonding Effect on the Performance of DFT Methods in Electric
wPP = 0.99 for AS2. Thus, there is a strong preference in SO2 Properties Calculations: A Pattern Recognition and Metric Space
for the double bond description. Approach on Some XY2 (X = O, S and Y = H, O, F, S, Cl) Molecules.
In summary, these results, as well as those in previous J. Comput. Chem. 2009, 412−420.
sections, show that the diradical bonding description of O3 (i) (13) Miliordos, E.; Xantheas, S. S. On the Bonding Nature of Ozone
(O3) and Its Sulfur-Substituted Analogues SO2, OS2, and S3:
smoothly describes the formation of O3 from O + O2, (ii) is the Correlation Between Their Biradical Character and Molecular
lowest energy configuration as a function of R(O′−O2), and Properties. J. Am. Chem. Soc. 2014, 136, 2808−2817.
(iii) provides a consistent description of O3 at its equilibrium (14) Goddard, W. A., III Improved Quantum Theory of Many-
geometry. It is for these reasons that the O3 diradical Electron Systems. I. Construction of Eigenfunctions of S2 Which
configuration was the focus of the present study. Satisfy Pauli’s Principle. Phys. Rev. 1967, 157, 73.

7693 DOI: 10.1021/acs.jpca.5b00998


J. Phys. Chem. A 2015, 119, 7683−7694
The Journal of Physical Chemistry A Article

(15) Goddard, W. A., III Improved Quantum Theory of Many- (37) Cooper, D. L.; Thorsteinsson, T.; Gerratt, J. Adv. Quantum
Electron Systems. II. Phys. Rev. 1967, 157, 81−93. Chem.; Lowdin, P.-O., Ed.; Academic Press: San Diego, 1998; Vol. 32,
(16) Lindquist, B. A.; Takeshita, T. Y.; Woon, D. E.; Dunning, T. H. pp 51−67.
Bonding in Sulfur-Oxygen CompoundsHSO/SOH and SOO/OSO: (38) Thorsteinsson, T.; Cooper, D. L.; Gerratt, J.; Karadakov, P. B.;
An Example of Recoupled Pair π Bonding. J. Chem. Theory Comput. Raimondi, M. Modern Valence Bond Representations of CASSCF
2013, 9, 4444−4452. Wavefunctions. Theor. Chem. Acc. 1996, 93, 343−366.
(17) Dunning, T. H., Jr; Xu, L. T.; Takeshita, T. Y. Fundamental (39) Kotani, M.; Amemyia, A.; Ishiguro, E.; Kimura, T. Tables of
Aspects of Recoupled Pair Bonds. I. Recoupled Pair Bonds in Carbon Molecular Integrals; Maruzen Co: Tokyo, 1963.
and Sulfur Monofluoride. J. Chem. Phys. 2015, 142, 034113. (40) Gerratt, J. General Theory of Spin-Coupled Wave Functions for
(18) Dunning, T. H., Jr; Takeshita, T. Y.; Xu, L. T. Fundamental Atoms and Molecules. Adv. At. Mol. Phys. 1971, 7, 141−221.
Aspects of Recoupled Pair Bonds. II. Recoupled Pair Bond Dyads in (41) J, G.; Lipscomb, W. N. Spin-Coupled Wave Functions for
Atoms and Molecules. Proc. Natl. Acad. Sci. U. S. A. 1968, 59, 332−
Carbon and Sulfur Difluoride. J. Chem. Phys. 2015, 142, 034114.
335.
(19) Takeshita, T. Y.; Dunning, T. H., Jr. Generalized Valence Bond
(42) Cooper, D. L.; Garrett, J.; Raimondi, M. The Electronic
Description of Chalcogen-Nitrogen Compounds. I. NS, F(NS) and
Structure of 1,3-Dipoles: Hypervalent Atoms. J. Chem. Soc., Perkin
H(NS). J. Phys. Chem. A 2015, 119, 1446−1455. Trans. 2 1989, 1187.
(20) Takeshita, T. Y.; Dunning, T. H., Jr. Generalized Valence Bond (43) Voter, A. F.; Goddard, W., III A Method for Describing
Description of Chalcogen-Nitrogen Compounds. II. NO, F(NO) and Resonance Between Generalized Valence Bond Wavefunctions. Chem.
H(NO). J. Phys. Chem. A 2015, 119, 1456−1463. Phys. 1981, 57, 253−259.
(21) Gillespie, R. J.; Popelier, P. L. A. Chemical Bonding and (44) Voter, A. F.; Goddard, W., III The Generalized Resonating
Molecular Geometry: From Lewis to Electron Densities; Oxford Valence Bond Method: Barrier Heights in the HF + D and HCl + D
University Press: New York, 2001. Exchange Reactions. J. Chem. Phys. 1981, 75, 3638.
(22) Hay, J. P. Generalized Valence Bond Studies of the Electronic (45) Woon, D. E.; Dunning, T. H., Jr. Theory of Hypervalency:
Structure of SF2, SF4, and SF6. J. Am. Chem. Soc. 1977, 99, 1003−1012. Recoupled Pair Bonding in SFn (n = 1−6). J. Phys. Chem. A 2009, 113,
(23) Woon, D. E.; Dunning, T. H., Jr. A Comparison Between Polar 7915−7926.
Covalent Bonding and Hypervalent Recoupled Pair Bonding in (46) Cunningham, T. P.; Cooper, D. L.; Gerratt, J.; Karadakov, P. B.;
Diatomic Chalcogen Halide Species {O,S,Se} × {F,Cl,Br}. Mol. Phys. Raimondi, M. Chemical Bonding in Oxofluorides of Hypercoordinate
2009, 107, 991−998. Sulfur. J. Chem. Soc., Faraday Trans. 1997, 93, 2247−2254.
(24) Werner, H. -J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Schütz,
M.; Celani, P.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.;
et al. MOLPRO, version 2010.1, a package of ab initio programs; 2010.
(25) Dunning, T. H. Gaussian Basis Sets for Use in Correlated
Molecular Calculations. I. The Atoms Boron Through Neon and
Hydrogen. J. Chem. Phys. 1989, 90, 1007.
(26) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in
Correlated Molecular Calculations. III. The Atoms Aluminum
Through Argon. J. Chem. Phys. 1993, 98, 1358.
(27) Dunning, T. H.; Peterson, K. A.; Wilson, A. K. Gaussian Basis
Sets for Use in Correlated Molecular Calculations. X. The Atoms
Aluminum Through Argon Revisited. J. Chem. Phys. 2001, 114, 9244.
(28) Purvis, G. D.; Bartlett, R. J. A Full Coupled-Cluster Singles and
Doubles Model: The Inclusion of Disconnected Triples. J. Chem. Phys.
1982, 76, 1910.
(29) Pople, J. A.; Head-Gordon, M.; Trucks, G. W.; Raghavachari, K.
A Fifth-Order Perturbation Comparison of Electron Correlation
Theories. Chem. Phys. Lett. 1989, 157, 479.
(30) Knowles, P. J.; Hampel, C. A.; Werner, H.-J. Coupled Cluster
Theory for High Spin, Open Shell Reference Wave Functions. J. Chem.
Phys. 1993, 99, 5219.
(31) Watts, J. D.; Gauss, J.; Bartlett, R. J. Coupled-Cluster Methods
with Noniterative Triple Excitations for Restricted Open-Shell
Hartree−Fock and Other General Single Determinant Reference
Functions. Energies and Analytical Gradients. J. Chem. Phys. 1993, 98,
8718.
(32) Werner, H.-J.; Knowles, P. J. An Efficient Internally Contracted
Multiconfiguration−Reference Configuration Interaction Method. J.
Chem. Phys. 1988, 89, 5803.
(33) Knowles, P. J.; Werner, H.-J. An Efficient Method for the
Evaluation of Coupling Coefficients in Configuration Interaction
Calculations. Chem. Phys. Lett. 1988, 145, 514−522.
(34) Langhoff, S. R.; Davidson, E. R. Configuration Interaction
Calculations on the Nitrogen Molecule. Int. J. Quantum Chem. 1974, 8,
61−72.
(35) Davidson, E. R.; Silver, D. W. Size Consistency in the Dilute
Helium Gas Electronic Structure. Chem. Phys. Lett. 1977, 52, 403−406.
(36) Cooper, D. L.; Thorsteinsson, T.; Gerratt, J. Fully Variational
Optimization of Modern VB Wave Functions Using the CASVB
Strategy. Int. J. Quantum Chem. 1997, 65, 439−451.

7694 DOI: 10.1021/acs.jpca.5b00998


J. Phys. Chem. A 2015, 119, 7683−7694

You might also like