Está en la página 1de 10

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/pccp | Physical Chemistry Chemical Physics

Hydrogen bonding in electronically excited states: a comparison between


formic acid dimer and its mono-substituted thioderivatives
Alvaro Cimas,a Otilia Mo,b Manuel Yanez,b Nazario Mart ncd and Ines Corral*b

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

Received 2nd June 2010, Accepted 30th July 2010


DOI: 10.1039/c0cp00772b
A multi-state complete active space second order perturbation theory (MS-CASPT2) study on the
valence singlet electronic excited states of formic acid dimer is presented. The electronic spectrum
of this dihydrogen bonded system is dominated by np* and pp* intramonomer and charge
transfer excitations and consists of a very intense pp* transition at 8.25 eV and three weaker
np*(2) and pp*(1) electronic excitations at 6.21 eV, 9.13 eV, and 9.93 eV, respectively.
The characteristic np*np*pp*pp*. . . pattern found in the formic acid dimer electronic
spectrum is altered when a sulfur atom is introduced in the molecule. Furthermore, carbonylby-thiocarbonyl or hydroxyl-by-thiohydroxyl substitution is predicted to strongly aect the
intensity of the above electronic transitions. The eect of oxygen-by-sulfur substitution on
the geometry of the rst excited state (S1) has been investigated at the CC2 and CASSCF
levels of theory. Although the two methods qualitatively predict the same geometrical changes
upon np* excitation, the geometries of the S1 state are found to dier considerably between the
two levels.

1. Introduction
Hydrogen bonded donoracceptor assemblies have emerged as
suitable models for understanding and controlling electron
and energy transfer in photochemistry and photobiology.
Moreover, some of these studies have shown that electronic
communication through hydrogen bonded interfaces can be
even more ecient than that found in comparable s or p
bonding networks.14 Besides enhancing the charge transfer
(CT) process, one of the requirements these hydrogen bonded
spacers must meet is not to compete with the photoinitiated
hydrogen bond mediated electron-transfer process but rather
to act as spectators in the overall photochemical process. The
overlap between the electronic absorption spectra of the
hydrogen bonded connector and the electron donor moiety
may result in competitive photoinitiated processes within
the hydrogen bonded moiety which can aect the quantum
yield of the electron transfer process. It is for this reason
that knowing beforehand the photophysics of the hydrogen
bonded linkers turns out to be crucial for the design of
hydrogen bond connected donoracceptor interfaces. Here,
we report on the valence singlet excited states of formic acid
dimer, which can be considered as the simplest prototype for
an intrasupramolecular dihydrogen bonded system.

Centro de Investigacao en Qumica, Departamento de Qumica e


Bioqumica, Facultade de Ciencias, Universidade do Porto,
Rua do Campo Alegre, 687, 4169-007 Porto, Portugal
b
Departamento de Qumica, C-9, Universidad Autonoma de Madrid,
Cantoblanco, 28049, Madrid, Spain. E-mail: ines.corral@uam.es;
Fax: +34-91-497-5238
c
Departamento de Qumica Organica, Facultad de Qumicas,
Universidad Complutense, 28040-Madrid, Spain
d
Instituto Madrileno de Estudios Avanzados en Nanociencia
(IMDEA-Nanociencia), Cantoblanco, 28049-Madrid, Spain

This journal is

the Owner Societies 2010

In the search for an adequate hydrogen bonded bridge


between the donor and the acceptor moieties in molecular
wires, we have considered it interesting to extend this simple
model and evaluate the eect in the electronic absorption
spectrum of introducing a sulfur atom in the system via
hydroxyl or carbonyl substitution by thio-hydroxyl or
thio-carbonyl groups, respectively.
Finally, we have characterized the equilibrium geometry of
the minimum of the rst excited state potential energy surface,
which is np* in nature in all the dimers considered, in order to
analyze the impact of np* electron promotion on the structure
of these dimers, paying special attention to the changes
induced in the double hydrogen bond, distinctive of carboxylic
acids and whose study has motivated a large number of
experimental and theoretical publications.
In fact, carboxylic acids constitute the simplest examples for
molecular aggregates via two hydrogen bonds, and have
served as suitable prototypes for DNA base pairs or other
species susceptible of undergoing double proton transfer.
Furthermore, formic acid is an important constituent of
clouds and fog and is widely used in manufacturing, which
explains the large number of theoretical and experimental
studies devoted to the characterization of its dimers with
water,510 pyridine,11 the hydroperoxyl radical12 or the
formamide13 molecule, for instance. Formic acid dimers have
been also extensively studied from both the theoretical and the
experimental points of view. The structure of formic acid
dimer was experimentally determined through the identication
and assignment of the rotational recurrences of dimeric
HCOOH in femtosecond degenerate four-wave mixing spectra
by Matylitsky et al.14 Several groups have also explored its
ground state potential energy surface15 or have addressed from
a theoretical perspective the question of the nature of its
intramolecular hydrogen bond,1618 whose heat of formation
Phys. Chem. Chem. Phys., 2010, 12, 1303713046

13037

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

was estimated to be 14.8 kcal mol1 by Clague and Bernstein19 in


1969. Finally, the infrared spectrum of formic acid dimer was
recorded and some frequencies were assigned by Halupka and
Sander,20 using matrix isolation under supersonic jet conditions.
Contrary to formic acid monomers electronic spectrum,
which has been intensively investigated,2134 studies on the
excited states of formic acid dimer are scarce. To the best of
our knowledge only three experimental works22,35,36 are available
in the literature reporting on the electronic absorption spectrum
of formic acid dimer, the latest published in 1992.35 Apparently,
the strong mixing of the monomer and dimer species in the
gaseous sample makes the assignment of the electronic
spectrum dicult and therefore only the rst two bands were
characterized and assigned to np* and pp* absorptions,
respectively.
The only theoretical studies on the low-lying excited states
of formic acid dimer were carried out in the groups of
Morokuma28 and Sathyamurthy,37 who also considered other
carboxylic acid dimers in their studies. Iwata and Morokuma
predict the np* vertical transition energy at 6.3 eV. Lourderaj
et al.37 compare the calculated vertical energies for the rst 3
singlet and triplet states at dierent levels of theory, namely,
time-dependent density functional theory (TD-DFT) with two
dierent functionals, LDA and B3LYP, complete active
space self-consistent eld (CASSCF) and multireference
conguration interaction (MR-CI), and conclude that for the
rst absorbing band, S2, TD-B3LYP provides the best
agreement with the most recent experiment,36 while CASSCF
overestimates this energy by more than 4 eV. Our results show
that both the appropriate choice of the orbitals of the active
space and the inclusion of dynamical correlation are crucial
for obtaining reliable excitation energies.

2. Computational details
Equilibrium ground state (S0) geometries for all dimers were
optimized using the hybrid functional B3LYP, which combines
three-parameter Beckes exchange potential38 and LeeYang
Parr non-local correlation functional.39 Two dierent basis
sets, namely 6-311++G(d,p) and 6-311++G(3df,2p), were
used for the description of formic acid dimer and HCOSH
HCOOH and HCSOHHCOOH, respectively. The inclusion
of additional sets of d- and f-type functions for sulfur-containing
dimers is based on the conclusions of previous works4042
which determine that the bonding in sulfur compounds and
their energies are signicantly inuenced by the addition of f
functions. Geometry optimizations of HCOOHHCOOH,
HCOSHHCOOH and HCSOHHCOOH were performed
under C2h, Cs, and Cs symmetry constraints, respectively.
The geometries, shown in Fig. 1, were subsequently employed
to calculate the absorption spectra using the methods
described below.
Absorption spectra were calculated using CASSCF
method43 combined with the atomic natural orbital ANO-L
basis set, contracted to C,O [4s3p2d]/H[3s2p]44 for the formic
acid dimer and the 6-311++G(3df,2p) basis set for the
HCSOHHCOOH and HCOSHHCOOH dimers.
The size of the active spaces used were (12,14), (12,12) and
(12,11) for formic acid dimer, HCOSHHCOOH and
13038

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

HCSOHHCOOH, respectively. These active spaces have in


common 8 orbitals; two pairs of pCQX (X = O, S) and p*CQX
(X = O, S) orbitals, the 2 lone pairs nCQX (X = O, S) and the
2 lone pairs perpendicular to the plane of the molecule nCXH
(X = O, S), summarized in Fig. 2. The rest of the orbitals up
to the size specied above are virtual orbitals, included
to avoid intruder states. Further eects connected to the
incompleteness of the active space were tackled through the
use of a real level-shift45 parameter of 0.3 a.u.
The CAS calculations were performed state average (SA)
over 6, 3, 2 and 3 roots of Ag, Au, Bu and Bg symmetries for
formic acid dimer, 13 and 8 roots of A 0 and A00 symmetries for
HCOSHHCOOH and, nally, 12 and 8 roots of A 0 and A00
symmetries for HCSOHHCOOH. Multi-state complete
active space second order perturbation theory, MS-CASPT2,
was employed to account for dynamical correlation. Oscillator
strengths were calculated from the transition dipole moments
obtained with the perturbative modied CASSCF wavefunctions and MS-CASPT2 energies.
Pre-optimization of the ground state (S0) and rst excited
state (S1) of all dimers was done at the CASSCF/6-31G(d)
level of theory, employing the (12,8) active space of Fig. 2.
Real vibrational frequencies conrm that the stationary points
found correspond to minima of the ground and excited
potential energy surfaces. These geometries were rened using
the bigger 6-311++G(3df,2p) basis set.
The comparison of the CASSCF and B3LYP geometries for
the S0 states showed up the poor performance of the CASSCF
method for the description of inter- and intramolecular
hydrogen bonds, already predicted by Song et al.46 and
Fores et al.47 Thus, dynamical correlations inuence on the
S1 minima geometries was evaluated through the reoptimization
of these structures using TD-BP86 and the resolution of
the identity second-order approximate coupled cluster,48
RI-CC2, methods combined with the def-TZVP basis sets,
respectively.
The atoms in molecules (AIM)49 and the electron localization
function (ELF)5052 theories were used to investigate the bond
activations undergone by the ground state minima upon
excitation. For this purpose, in the framework of the AIM
theory, we have evaluated the charge density r(r) at the
corresponding bond critical points (BCPs). This analysis will
be complemented with that carried out in terms of the
lengthening or shortening of the bond lengths.
The ELF theory permits the partition of the molecular space
in basins with a maximum probability of nding a pair of
electrons, so that monosynaptic basins, with the participation
of the valence shell of a single atom, correspond to core
electrons or electron pairs, whereas disynaptic basins, with
contribution from the valence shell of two atoms, correspond
to covalent bonds. ELF grids and basin integrations have been
evaluated with the TopMod package,53 using an ELF value
around 0.80.
All the MS-CASPT2/SA-CASSCF calculations were
performed using the MOLCAS 6.4 quantum chemistry
package.54 The B3LYP and CASSCF geometry optimizations
were carried out with the Gaussian 03 suite of programs,55 and
the TD-BP86 and RI-CC2 excited state optimizations were
done with Turbomole.56
This journal is

the Owner Societies 2010

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

Fig. 1 Optimized geometries for the ground (S0) state (rst column) and the rst excited (S1) state (second and third columns) of formic acid dimer
and its thio-derivatives. B3LYP values are given in normal font, TD-BP86/def-TZVP values in bold, CASSCF(12,8)/6-311++G(3df,2p) values in
parentheses, and RI-CC2/def-TZVP parameters in curly brackets. When available, experimental values are given in brackets. Note that only
TD-BP86 (in bold) and RI-CC2 (curly brackets) values are available for the excited states of formic acid dimer and HCSOHHCOOH (see text).
Bond distances are given in Angstrom and angles in degrees.

3. Results and discussion


3.1

Formic acid dimer

The most relevant experimental equilibrium geometrical parameters14 for the ground state of formic acid dimer along with
B3LYP/6-311++G(d,p), BP86/def-TZVP, CASSCF(12,8)/
6-311++G(3df,2p) and RI-CC2/def-TZVP optimized values
are reported in Fig. 1. In general, the B3LYP and CC2 single
reference methods provide calculated structures in good
agreement with the experiment. The bigger discrepancies are
found for the OH and H  O bond distances involved in the
hydrogen bond, which are slightly underestimated by DFT
and CC methods. As already anticipated, the bigger deviations
from the experiment are found for the CASSCF method.
Interestingly, the deviations from the experimental bond
distances provided by CASSCF have, in many cases, opposite
sign to those predicted by B3LYP and CC2 theories, and
for particular bond distances we nd that CASSCF and
B3LYP/CC2 errors can dier in one order of magnitude.
This journal is

the Owner Societies 2010

For instance, H  O hydrogen bond distance is 0.209 A overestimated by CASSCF, whereas RI-CC2 underestimates this
distance by 0.034 A, in comparison with the experiment.
Due to the better performance of the B3LYP method,
formic acid dimer vertical excitation energies were determined
at the ground state equilibrium geometry optimized at the
B3LYP/6-311++G(d,p) level of theory.
The MS-CASPT2/SA-CASSCF FranckCondon excitation
energies calculated with the (12,14) active space together with
the associated oscillator strengths are collected in Table 1.
In principle one would expect to distinguish two clear
regions in the absorption spectrum; the rst one, at lower
energies, arising from electronic excitations occurring within
one of the monomers of the dimer, and a second, at higher
energies, governed by CT electronic transitions involving the
two monomers. In practice, these two regions, which extend
from 6.13 eV to 8.25 eV and from 9.13 eV onwards, are
composed of two pairs of degenerate nCQO - p*CQO and
pCQO - p*CQO transitions. The high symmetry of the dimer
Phys. Chem. Chem. Phys., 2010, 12, 1303713046

13039

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

Fig. 3 Simulated absorption bands arising from MS-CASPT2


vertical valence transitions for HCOOHHCOOH (black solid line),
HCOSHHCOOH (red dashed line) and HCSOHHCOOH (blue
dotted line) dimers. A half-width at half-maximum value of 1.0 was
used for all the Lorentzians. The inset shows the calculated
MS-CASPT2 valence transitions for formic acid dimer superimposed
on the experimental absorption spectrum provided in ref. 22.

Fig. 2 CASSCF molecular orbitals common to the active spaces


employed for the calculation of the valence vertical transition energies
in the 3 dimers considered in this work. The orbitals in the gure were
extracted from formic acid dimer calculations. Similar orbitals or
linear combinations of them were employed for the calculations of
S-substituted heterodimers.

(C2h), however, results in the mixing of the orbitals of the two


monomers, making dicult the distinction between intramonomer and CT transitions. Also, the degeneracy between
pairs of electronic transitions of the same nature is a

consequence of the high symmetry of the molecule (see


Table 1). Nevertheless, upon inspection of Table 1 small
energy gaps, which amount to ca. 0.1 eV, between pairs of
excitations can be observed, which we attribute to small
dierences in the active spaces used for the calculation of
excitations of dierent symmetries.
Fig. 3 shows the experimental absorption spectrum of
formic acid dimer which covers the biggest range of wavelengths in the literature22 (inset of Fig. 3) and the absorbing
MS-CASPT2 calculated vertical transitions (black solid
spectrum in Fig. 3 and stick spectrum in the inset of Fig. 3).
In the experimental spectrum two structureless bands can be
distinguished: a weak band centered at 210 nm36 that was
ascribed to np* excitations, followed by a broad band ranging

Table 1 MS-CASPT2 excitation energiesa DE (eV, nm), main one-electron excitations with the conguration interaction (CI) coecients and
oscillator strengths f, for the low lying excited states of the formic acid dimer
MS-CASPT2//CASSCF(12,14)/ANO-L
State symmetry
1

1 Bg (S1)
11Au (S2)
1

2 Ag (S3)
11Bu (S4)
21Au (S5)
21Bg (S6)
31Ag (S7)
21Bu (S8)

Main conguration
nBuCQO
nAgCQO
nBuCQO
nAgCQO
pAuCQO
pBgCQO
pBgCQO
pAuCQO
nAgCQO
nBuCQO
nAgCQO
nBuCQO
pBgCQO
pAuCQO
pAuCQO
pBgCQO

p*AuCQO
p*BgCQO
p*BgCQO
p*AuCQO
p*AuCQO
p*BgCQO
p*AuCQO
p*BgCQO
p*AuCQO
p*BgCQO
p*BgCQO
p*AuCQO
p*BgCQO
p*AuCQO
p*BgCQO
p*AuCQO

CI coecient

DE/eV

DE/nm

0.76
0.55
0.69
0.64
0.73
0.48
0.70
0.61
0.67
0.62
0.75
0.55
0.72
0.49
0.67
0.58

6.13

202

0.0000

6.21

200

0.0010

7.56

164

0.0000

8.25b

150

0.6553

9.13

136

0.0015

9.17

135

0.0000

9.80

127

0.0000

9.93

125

0.0051

Ground state total energy: 378.841141 hartrees. b These two values should be identical because the states are formally degenerate. The origin of
this energy gap is the too low reference weight for the S3 state in the CASPT2 calculation.

13040

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

This journal is

the Owner Societies 2010

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

from 125170 nm of pp* nature. In agreement with these


ndings, MS-CASPT2 predicts 4 bright transitions within the
spectral region experimentally recorded (see inset of Fig. 3).
Although weak, the rst transition with non-zero oscillator
strength corresponds to the S2 state. This excitation lies at
6.21 eV (200 nm), and would account for the weak absorption
at low energies found in the experimental spectra. Three more
bright excitations follow the 11Au state in the theoretical
spectrum. At 8.25 eV (150 nm), coinciding with the beginning
of the most intense experimental band, the pCQO - p*CQO
transition peaks, with the highest oscillator strength of all the
transitions calculated. The np* and pp* excitations at 9.13 eV
(136 nm) and 9.93 eV (125 nm) would also contribute to the
broadest and most intense experimental band, although in a
minor extent since their oscillator strengths are 2 orders of
magnitude smaller than for the 11Bu state.
The theoretical work by Lourderaj et al.37 focuses on the
low energy region of the absorption spectrum of the dimer,
and provides the vertical transition energies for the rst three
singlets and triplets at dierent levels of theory. From the
comparison of these results with our MS-CASPT2 excitation
energies and the experimental absorption spectrum, we
conclude that TD-B3LYP is able to accurately predict the
position of the np* transition responsible for the weak absorption
band at low energies. The next bright transition in the
TD-B3LYP spectrum at 6.92 eV (192 nm) involves one
electron excitation from the 10ag to the 3au orbital. Considering
the symmetry of the orbitals involved, we assume that this
excitation corresponds to the CT S5 state at MS-CASPT2 level
(9.13 eV/136 nm). Not surprisingly, TD-B3LYP underestimates
by 2.21 eV the excitation energy for this np* transition, which
can be attributed to the poor description of CT states achieved

in general by standard DFT functionals. As expected, the


CASSCF excitation energies are signicantly overestimated
compared to MS-CASPT2 values, due to the lack of dynamical
correlation of the former method. Moreover, CASSCF(8,8)
energies reported in ref. 37 deviate between ca. 13 eV from
our CASSCF(12,14), which we ascribe to the reduced size
and/or composition of the active space used by the authors.
In order to get some insight into the nature of the rst
excited state of the dimer, of np* nature, the minimum
connected to the S1 was located at the CASSCF(12,8)/
6-311++G(3df,2p) level of theory. Note that due to the
symmetry of the molecule, an equivalent geometry is expected
for the minimum associated to the FranckCondon S2 excited
state. The most relevant CASSCF geometrical parameters for
both minima are shown in Fig. 1.
From the comparison of the S0 and S1 geometries (see
Fig. 1), it is clear to see the loss of planarity upon intramonomer one electron excitation from the lone pair to the p*
orbital. More precisely, the carbonyl group involved in the
excitation moves out from the plane of the molecule, due to
the population of the p* orbital of the carbonyl CQO double
bond. Also connected to this bond weakening is the change in
the CQO and COH distances within this monomer
that lengthen by 0.185 A and 0.026 A, respectively. These
lengthenings reect a signicant decrease in the electron
densities at the corresponding CQO and COH BCPs on
going from the ground state minimum (0.4449 and 0.3386
electrons per volume unit (e a.u.3)) to the minimum of the
rst excited state (0.2827 and 0.3151 e a.u.3) (see Fig. 4).
Concomitantly, the population of the corresponding COH
and CQO disynaptic basins also decreases dramatically (from
1.76 and 2.48 e to 1.50 and 1.62 e, respectively) (see Fig. 5).

Fig. 4 Molecular graphs of HCOOHHCOOH, HCOSHHCOOH and HCSOHHCOOH ground state minima (left) and the minima of rst
excited state PES (right). Electron densities at the BCPs and RCPs are in a.u.

This journal is

the Owner Societies 2010

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

13041

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

Fig. 5 Three-dimensional representations of ELF isosurfaces with ELF = 0.80 for the S0 ground state (left column) and the S1 rst excited state
(right column) of formic acid dimer and its thio-derivatives. Yellow lobes correspond to disynaptic basins involving H atoms, green
lobes correspond to disynaptic basins between two bonded atoms and red lobes to electron lone-pairs. The population of the dierent basins is
given in e.

Less important are, however, the changes observed in the bond


distances and angles of the opposite monomer, whose geometry
remains almost unaltered compared to that of the ground
state. The np* excitation also inuences the double hydrogen
bond. Coincidentally, both H  O distances increase by ca.
0.15 A whilst the charge density at these BCPs decreases
almost by a factor of two at the CASSCF level of theory.
One important factor aecting the strength of the CQO  H
hydrogen bond in S1 is the change in the orientation of the
oxygen lone pairs, which in the ground state lie on the plane
of the molecule, favoring the interaction with the OH
donor of the second monomer, whereas in the excited
state they appear rotated and almost perpendicular to the
molecular plane, which aects signicantly the strength of the
hydrogen bond.
Considering the poor performance of the CASSCF method
concerning hydrogen bond description, the S1 excited state
was reoptimized at the RI-CC2/def-TZVP and TD-BP86 levels
of theory, in order to correct the deciencies of the former
method. The most relevant bond angles and distances at these
levels of theory are also summarized in Fig. 1.
From the comparison of the S0 and S1 RI-CC2 geometries,
we observe in general the same trends described above for the
CASSCF method. Upon np* excitation, CQO and COH
bond distances for the monomer whose orbitals are involved in
the transition increase, respectively, by 0.224 A and 0.024 A.
13042

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

The changes in H  O bond distances, however, are predicted


to be subtler at these levels of theory.
3.2 HCOSHHCOOH
This section analyses the changes observed in the electronic
absorption spectrum, and in the S0 and S1 geometries of formic
acid dimer upon hydroxyl-by-thiohydroxyl substitution.
From the comparison of HCOSHHCOOH ground state
geometries optimized at dierent levels of theory, we nd
similar trends for the performance of the theoretical methods
as the ones observed for formic acid dimer. Due to the lack of
experimental structures for this dimer, B3LYP geometrical
parameters were taken as a reference. Compared to B3LYP,
CASSCF tends to overestimate the strength of covalent bonds,
giving rise to shorter bond distances than B3LYP, except for
the CSH bond, which is predicted to be 0.01 A longer at
CASSCF level of theory. Also in agreement with our ndings
for formic acid dimer and with the observations by Song
et al.46 and Fores et al.,47 CASSCF was found to signicantly
overestimate hydrogen bond distances in comparison with
B3LYP, e.g. CASSCF predicts SH  O and O  HO distances
to be 0.359 A and 0.252 A longer than those provided
by B3LYP, respectively. Interestingly other than SH  O
hydrogen bond interaction, OH-by-SH substitution within
one of the monomers was not found to specially inuence
the ground state geometry of the dimer.
This journal is

the Owner Societies 2010

View Article Online

Table 2 MS-CASPT2 excitation energiesa D,E (eV, nm), main one-electron excitations with the conguration interaction (CI) coecients and
oscillator strengths f, for the low lying excited states of the HCOSHHCOOH dimer. (S) and (O) indicate the monomer, SH- or OH-substituted, to
which the referred orbital belongs to. (+) and () denote the positive or negative linear combination of the nCQO orbitals. See Fig. 2
MS-CASPT2//CASSCF(12,12)/6-311++G(3df,2p)
State symmetry

Main conguration

11A00 (S1)

nCQO
nCQO
pCQO
nCQO
nCQO
pCQO
nCQO
nCQO
pCQO
nCQO
nCQO
nCQO
nCQO
nCQO
pCQO
pCQO
nCQO
nCQO
nCQO
nCQO

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

21A 0 (S2)
21A00 (S3)
31A 0 (S4)
31A00 (S5)
41A 0 (S6)
41A00 (S7)
51A00 (S8)
51A 0 (S9)
61A 0 (S10)
61A00 (S11)

CI coecient

(S) - p*CQO (S)


(O) - p*CQO (S)
(S) - p*CQO (S)
(S) - p*CQO (O)
(O) - p*CQO (O)
(O) - p*CQO (O)
(O) - p*CQO (S)
(S) - p*CQO (S)
00
(S) - RydpA00
A
(S) - Rydp 00
(O) - RydpA00
(S) - RyddA 00
(O) - RyddA00
(O) - RyddA
(O) - p*CQO (S)
(S) - p*CQO (O)
(S) - p*CQO (O)
(O) - p*CQO (O)
00
(S) - RyddA 00
A
(O) - Rydd

a
Ground state total energy: 701.451554 hartrees.
transition should be around 9.1 eV (see Table 1).

0.69
0.65
0.89
0.63
0.60
0.86
0.67
0.64
0.89
0.73
0.54
0.53
0.42
0.49
0.82
0.82
0.47
0.47
0.45
0.38

the Owner Societies 2010

DE/nm

4.94

251

0.0000

5.98
6.17

207
201

0.3964
0.0011

8.31
8.71

149
142

0.4555
0.0003

8.76
9.26

141
134

0.0079
0.0031

9.58

129

0.0044

9.59
9.71
10.63b

129
128
117

0.0042
0.0108
0.0000

Overestimated value due to the mixing with Rydberg states. The expected value for this

The MS-CASPT2/SA-CASSCF FranckCondon excitation


energies and oscillator strengths for the HCOSHHCOOH
dimer, calculated on the B3LYP/6-311++G(3df,2p) ground
state geometry, are collected in Table 2.
Similarly to the formic acid dimer spectrum, there is a clear
separation between intramonomer excitations below 8.31 eV
and CT excitations, which peak at energies above 8.71 eV. In
this case, however, the degeneracy between pairs of excitations
of the same nature disappears and the ordering rst np*
then pp* within each group of excitations observed for
HCOOHHCOOH does not hold any longer. The OH-by-SH
substitution in one of the monomers not only breaks the C2h
symmetry of formic acid dimer but also aects p and p* orbital
energies of the sulfur substituted monomer leading to shifts in
the excitation energies involving these orbitals (see Table 2).
Curiously, the two pairs of np* and pp* transitions, 21A00 ,
31A 0 , 61A00 and 61A 0 , which promote an electron to the p*CQO
orbital localized on the formic acid moiety, peak at the same
energies as those calculated for formic acid dimer (compare
Tables 1 and 2). However, MS-CASPT2 protocol predicts
excitations into the p*CQO orbital localized on the SH
substituted unit at much lower energies. More precisely,
intramonomer np* and pp* excitations shift by 1 and 2 eV
to lower energies, while the shifts experienced by CT excitations
are much smaller and amount to ca. 0.20.4 eV. The latter
shifts are small enough as to preserve the pattern observed in
formic acid dimer for the CT region. However, since the shifts
are much larger for the intramonomer excitations the pattern
changes to np*pp*np*pp*. Also of note is the analysis of
the changes connected to the oscillator strengths upon
sulfur substitution. While np* excitations exhibit very weak
absorptions in both dimers, the pp* excitations which were
dark, for symmetry reasons, in the formic acid dimer become
This journal is

DE/eV

among the most intense absorptions in the UV spectrum of


HCOSHHCOOH (compare the black solid and the red
dashed spectra in Fig. 3). Another important dierence
between the spectra of formic acid dimer and HCOSH
HCOOH are excitations into Rydberg orbitals. While our
calculations do not predict the existence of low lying Rydberg
states for formic acid dimer, these states govern the region
of the absorption spectrum lying above 8.76 eV in
HCOSHHCOOH.
In order to evaluate the eect of OH-by-SH substitution on
the geometry of the minimum connected to the rst excited
state, the np* S1 excited state was optimized at the
CASSCF(12,8)/6-311++G(3df,2p) level of theory, as a rst
approach. The most relevant geometrical parameters are
shown in Fig. 1. Similarly to formic acid dimer, the equilibrium
geometry of the S1 minimum is not planar and the carbonyl
group of the monomer containing the S atom deviates ca. 651
from the plane of the molecule, as a consequence of the
population of the p*CQO orbital. The population of the
p*CQO orbital leads to a signicant lengthening (by 0.156 A)
of the CQO bond distance in the S1 excited minimum. Also in
this case, CASSCF theory predicts signicant changes in the
nature of the intramolecular hydrogen bonds of the heterodimer. The out-of-plane displacement of the carbonyl group of
one of the two subunits of the dimer leads to a reorganization
of the hydrogen bond interface, resulting in an increase by ca.
0.1 and 0.2 A of the SH  O and O  HO bond distances.
This rearrangement also aects the hydrogen-bond angle
SH  O, which reduces from 167.71 to 146.61. The structural
reorganization undergone by the heterodimer upon np*
promotion also aects the topology of the charge density.
Consequently, the activation of the CQO double bond and the
SH  O and O  HO hydrogen bonds is translated into a
Phys. Chem. Chem. Phys., 2010, 12, 1303713046

13043

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

decrease of the charge density of the corresponding BCPs


(see Fig. 4), and into a signicant decrease of the population of
the CQO and CS disynaptic basins (see Fig. 5).
In order to correct the well-known deciencies of CASSCF
theory for the description of ground and excited state
hydrogen bonds, we tried to reoptimize this minimum at the
RI-CC2/def-TZVP level of theory. Unfortunately, CC2 fails to
optimize this minimum, showing a similar behavior as
TD-BP86. In fact, along the optimization path a higher lying
CT pp* state crosses the title np* state, preventing the
optimization of the latter.
In principle, from the conclusions extracted from the
comparison of the geometries obtained at dierent levels of
theory considered in this work we expect the geometry for the
S1 excited state to be in qualitative agreement with the
CASSCF structure. Nevertheless, we presume that introducing
dynamical correlation in our calculations would result in
stronger intramolecular H-bond distances, a smaller HSCO
dihedral angle and a more stretched CQO bond distance for
the monomer where the excitation occurs.
3.3

HCSOHHCOOH

Due to the superior performance of the B3LYP functional


indicated above, excited state calculations for CQS substituted
formic acid dimer were carried out on the B3LYP/
6-311++G(3df,2p) optimized geometry, shown in Fig. 1.
Table
3
collects
the
MS-CASPT2/SA-CASSCF
FranckCondon excitation energies in the energy range
below 9.9 eV, together with their calculated intensity. Similarly
to the vertical absorption spectra calculated for formic acid
and HCOSHHCOOH dimers, all np* and pp* excitations
arrange into two groups of transitions depending on whether
the excitation takes place within a single monomer or if it
involves the two subunits of the dimer. Thus, the rst two
pairs of np* and pp* excitations occurring in one of the
monomers peak below 8.33 eV, whilst their CT analogues appear

in the range of 8.629.84 eV. Also for this system Rydberg states
absorbing at low energies were found above 7.98 eV.
Since the three dimers considered contain, at least, one
formic acid molecule it is reasonable to think that part of
the spectrum should be common to all of them. In fact, the S3
and S5 excited states in HCSOHHCOOH taking place in the
formic acid subunit absorb at 6.27 and 8.33 eV very close to
the values calculated for these transitions in the other two
dimers. Notice from the examination of the spectrum in
Table 3 that O-by-S substitution in the carbonyl group also
alters the relative ordering of np* and pp* excitations.
Intramonomer one electron promotion into p* orbitals of
the sulfur-substituted monomer involves signicantly lower
excitation energies compared to transitions into p* orbitals of
the formic acid monomer, whereas the opposite is observed in
the CT region. This leads to a np*pp*np*pp*. . . pattern in
the vertical spectrum. It is worthwhile mentioning that sulfur
by oxygen substitution in the carbonyl group leads in general
to a red-shift in the excitation energies. This shift is especially
important in the low energy region of the spectrum where
energy displacements amount to 2.5 eV.
Finally, it is important to stress that CQO-by-CQS
substitution has also an eect on the oscillator strength of
pp* transitions. In particular, the very intense band in formic
acid vertical spectrum arising from one of the two intramonomer pp* excitations (recall that the other pp* excitation
is dark for symmetry reasons) splits into two very intense
absorptions, the rst at 5.58 eV and the second at 8.33 eV (see
Fig. 3). This is also the case of the second pp* CT excitation,
dark in the unsubstituted dimer, which becomes a bright
transition upon sulfur substitution.
The most relevant geometrical values for HCSOHHCOOH
upon nSp*CQS excitation are presented in Fig. 1. As already
pointed for the other two dimers, the intramonomer promotion
of an electron from the lone pair (nS) into the p* orbital
(p*CQS) seems, also in this case, not to aect the geometry of

Table 3 MS-CASPT2 excitation energiesa DE (eV, nm), main one-electron excitations with the conguration interaction (CI) coecients and
oscillator strengths f, for the low lying excited states of the HCSOHHCOOH dimer
MS-CASPT2//CASSCF(12,11)/6-311++G(3df,2p)
State symmetry
1

00

1 A (S1)
2 A (S2)
21A00 (S3)
31A00 (S4)
31A 0 (S5)
41A00 (S6)
41A 0 (S7)
51A 0 (S8)
51A00 (S9)
1

6 A 0 (S10)
a

Main conguration

CI coecient

DE/eV

DE/nm

nCQS - p*CQS
nCQO - p*CQS
pCQS - p*CQS
nCQS - p*CQO
nCQO - p*CQO
00
nCQS - RydpA00
A
nCQS - Rydd
pCQO - p*CQO
nCQO - p*CQO
nCQS - p*CQO
pCQS - p*CQO
00
pCQS - RydpA00
A
pCQS - Rydd
pCQS - p*CQO
00
pCQS - RydpA
nCQO - p*CQS
nCQS - p*CQS
pCQO - p*CQS
nCOH (S) - p*CQS

0.72
0.60
0.92
0.68
0.60
0.55
0.50
0.85
0.49
0.48
0.71
0.55
0.65
0.52
0.43
0.64
0.57
0.76
0.36

3.82

324

0.0000

5.58
6.27

222
198

0.4771
0.0012

7.98

155

0.0030

8.33
8.62

149
144

0.4711
0.0017

8.73

142

0.0031

8.82

141

0.0087

8.92

139

0.0010

9.84

126

0.0345

Ground state total energy: 701.458382 hartrees.

13044

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

This journal is

the Owner Societies 2010

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

the formic acid subunit. Within the thioformic acid monomer,


however, both CASSCF and CC2 methods agree to predict:
(i) an increase by 0.16 A and 0.04 A in the CQS and SCOH
bond distances, respectively, (ii) the weakening of the OH  OC
hydrogen bond, which increases by ca. 0.13 A and (iii) a
slight reinforcement of the OH bond, as a consequence of
the signicant activation of the OH  OC hydrogen bond.
Also in this case, the lengthening and shortening of the bond
lengths are consistent with the decrease/increase of the
electron density at the corresponding BCPs (see Fig. 4) and
with the decrease/increase of the electron population of the
corresponding disynaptic basins (see Fig. 5). The obvious
out-of-plane displacement of the CQS group as well as the
rearrangement of the hydrogen bond interface connected to it
in the S1 geometry are also predicted by CASSCF and CC2
protocols, though, for these quantities bigger discrepancies
between the two methods are found. In this respect, it is worth
noting that the reorientation of the sulfur lone pairs of the
thiocarbonyl group is apparent in the corresponding ELF
(see Fig. 5).

4. Summary and conclusions


This paper reports on the MS-CASPT2 analysis of gas-phase
singlet valence vertical transition energies for formic acid
dimer and their comparison to those predicted for SH- and
CQS-substituted formic acid heterodimers at the same level of
theory. Two clear regions can be dierentiated in the spectra
of the three dimers, the rst which consists of intramonomer
np* and pp* transitions and a second, at higher energies,
composed of CT np* and pp* excitations which involve the
two monomers at the same time. The electronic absorption
spectrum of formic acid dimer is strongly inuenced by the
high symmetry of the molecule. In this sense, the spectrum is
composed by a series of pairs of degenerate np* and pp*
transitions, one of them dark within each pair. Hydroxylby-thiohydroxyl or carbonyl-by-thiocarbonyl substitutions
were found to aect both the position and oscillator strength
of molecular valence transitions. In fact, oxygen by sulfur
exchange translates, in general, into a shift to the red of the
vertical excitations energies and into an increase of the
absorption intensities, especially of pp* excitations, which
breaks the characteristic np*np*pp*pp* pattern of formic
acid electronic spectrum.
Finally, the nature of the minima associated with the
lowest bright state in formic acid dimer and its sulfurmonosubstituted heterodimers, which show an np* character,
was investigated at the CASSCF, TD-BP86 and CC2 levels of
theory. Intramonomer np* promotion results in the out-ofplane displacement of the carbonyl/thiocarbonyl group
involved in the excitation and in the activation of the CQX
and CXH bonds within the excited monomer, as a consequence of the population of the p* orbital.

Acknowledgements
This work has been supported by the DGI Project No.
CTQ2009-13129-C02, by the Project MADRISOLAR2, Ref.:
S2009PPQ/1533 of the Comunidad Autonoma de Madrid, by
This journal is

the Owner Societies 2010

Consolider on Molecular Nanoscience CSD2007-00010 and


by the COST Action COST CM0702. A generous allocation of
computing time at the CCC of the UAM is also acknowledged.
AC gratefully acknowledges nancial support from the DGI
project CTQ2006-08558/BQU, and IC from the Programa
Juan de la Cierva.

References
1 P. J. F. Derege, S. A. Williams and M. J. Therien, Science, 1995,
269, 14091413.
2 L. Sanchez, M. Sierra, N. Mart n, A. J. Myles, T. J. Dale, J. Rebek,
W. Seitz and D. M. Guldi, Angew. Chem., Int. Ed., 2006, 45,
46374641.
3 J. L. Sessler, B. Wang and A. Harriman, J. Am. Chem. Soc., 1993,
115, 1041810419.
4 F. Wessendorf, J. F. Gnichwitz, G. H. Sarova, K. Hager,
U. Hartnagel, D. M. Guldi and A. Hirsch, J. Am. Chem. Soc.,
2007, 129, 1605716071.
5 S. Aloisio, P. E. Hintze and V. Vaida, J. Phys. Chem. A, 2002, 106,
363370.
6 P. O. Astrand, G. Karlstrom, A. Engdahl and B. Nelander,
J. Chem. Phys., 1995, 102, 35343554.
7 J. Chocholousova, V. Spirko and P. Hobza, Phys. Chem. Chem.
Phys., 2004, 6, 3741.
8 R. Kumaresan and P. Z. Kolandaivel, Z. Phys. Chem. (Munich),
1995, 192, 191.
9 D. Preim, T. K. Ha and A. Bauder, J. Chem. Phys., 2000, 113,
169175.
10 P. R. Rablen, J. W. Lockman and W. L. Jorgensen, J. Phys. Chem.
A, 1998, 102, 37823797.
11 M. J. Fernandez-Berridi, J. J. Iruin, L. Irusta, J. M. Mercero and
J. M. Ugalde, J. Phys. Chem. A, 2002, 106, 41874191.
12 M. Torrent-Sucarrat and J. M. Anglada, J. Phys. Chem. A, 2006,
110, 97189726.
13 L. F. Pacios, J. Comput. Chem., 2006, 27, 16411649.
14 V. V. Matylitsky, C. Riehn, M. F. Gelin and B. Brutschy, J. Chem.
Phys., 2003, 119, 1055310562.
15 J. Chocholousova, J. Vacek and P. Hobza, Phys. Chem. Chem.
Phys., 2002, 4, 21192122.
16 R. W. Gora, S. J. Grabowski and J. Leszczynski, J. Phys. Chem. A,
2005, 109, 63976405.
17 W. Qain and S. Krimm, J. Phys. Chem. A, 2002, 106, 1166311671.
18 Y. Zhao and D. G. Truhlar, J. Phys. Chem. A, 2005, 109,
66246627.
19 A. D. H. Clague and H. J. Bernstein, Spectrochim. Acta, Part A,
1969, 25, 593596.
20 M. Halupka and W. Sander, Spectrochim. Acta, Part A, 1998, 54,
495500.
21 T. Ari and M. H. Guven, J. Electron Spectrosc. Relat. Phenom.,
2000, 106, 2935.
22 E. E. Barnes and W. T. Simpson, J. Chem. Phys., 1963, 39,
670675.
23 H. Basch, M. B. Robin and N. A. Kuebler, J. Chem. Phys., 1968,
49, 50075018.
24 S. Bell, T. L. Ng and A. D. Walsh, J. Chem. Soc., Faraday Trans. 2,
1975, 71, 393401.
25 D. Demoulin, Chem. Phys., 1976, 17, 471478.
26 C. Fridh, J. Chem. Soc., Faraday Trans. 2, 1978, 74, 190193.
27 F. Ioannoni, D. C. Moule and D. J. Clouthier, J. Phys. Chem.,
1990, 94, 22902294.
28 S. Iwata and K. Morokuma, Theor. Chim. Acta, 1977, 44, 323339.
29 S. Leach, M. Schwell, F. Dulieu, J. L. Chotin, H.-W. Jochims and
H. Baumgartel, Phys. Chem. Chem. Phys., 2002, 4, 50255039.
30 S. Leach, M. Schwell, D. Talbi, G. Berthier, K. Hottmann,
H.-W. Jochims and H. Baumgartel, Chem. Phys., 2003, 286, 1543.
31 S. Nagakura, K. Kaya and H. Tsubomura, J. Mol. Spectrosc.,
1964, 13, 14.
32 T. L. Ng and S. Bell, J. Mol. Spectrosc., 1974, 50, 166181.
33 S. D. Peyerimho and R. J. Buenker, J. Chem. Phys., 1969, 50,
18461861.
34 W. C. Price and W. M. Evans, Proc. R. Soc. London, Ser. A, 1937,
162, 110120.

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

13045

Published on 31 August 2010. Downloaded by Temple University on 28/10/2014 02:30:28.

View Article Online

35 R. Mualem, E. Sominska, V. Kelner and A. Gedanken, J. Chem.


Phys., 1992, 97, 88138814.
36 D. L. Singleton, G. Paraskevopoulos and R. S. Irwin,
J. Photochem., 1987, 37, 209216.
37 U. Lourderaj, K. Giri and N. Sathyamurthy, J. Phys. Chem. A,
2006, 110, 27092717.
38 A. D. Becke, J. Chem. Phys., 1993, 98, 13721377.
39 C. T. Lee, W. T. Yang and R. G. Parr, Phys. Rev. B: Condens.
Matter, 1988, 37, 785789.
40 V. Jonas and G. Frenking, Chem. Phys. Lett., 1991, 177, 175183.
41 A. Timoshkin and G. Frenking, J. Chem. Phys., 2000, 113,
84308433.
42 L. Gonzalez, O. Mo and M. Yanez, Chem. Phys. Lett., 1996, 263,
407413.
43 B. O. Roos, in Ab initio Methods in Quantum Chemistry-II,
ed. K. P. Lawyley, Wiley, Chichester, 1987.
44 P. O. Widmark, P. A. Malmqvist and B. O. Roos, Theor. Chim.
Acta, 1990, 77, 291306.
45 B. O. Roos and K. Andersson, Chem. Phys. Lett., 1995, 245, 215223.
46 X. L. Song, J. C. Li, H. Hou and B. S. Wang, J. Chem. Phys., 2006,
125, 09430110943016.

13046

Phys. Chem. Chem. Phys., 2010, 12, 1303713046

47 M. Fores, M. Duran, M. Sola and L. Adamowicz, J. Comput.


Chem., 2000, 21, 257269.
48 O. Christiansen, H. Koch and P. Jorgensen, Chem. Phys. Lett.,
1995, 243, 409418.
49 R. F. W. Bader, Atoms in Molecules: A Quantum Theory,
Clarendon Press, Oxford, 1990.
50 A. D. Becke and K. E. Edgecombe, J. Chem. Phys., 1990, 92,
53975403.
51 A. Savin, R. Nesper, S. Wengert and T. F. Fassler, Angew. Chem.
Int. Ed., 1997, 36, 18091832.
52 B. Silvi and A. Savin, Nature, 1994, 371, 683686.
53 S. Noury, X. Krokidis, F. Fuster and B. Silvi, Comput. Chem.,
1999, 23, 597604.
54 G. Karlstrom, R. Lindh, P. A. Malmqvist, B. O. Roos, U. Ryde,
V. Veryazov, P. O. Widmark, M. Cossi, B. Schimmelpfennig,
P. Neogrady and L. Seijo, Comput. Mater. Sci., 2003, 28,
222239.
55 M. J. Frisch, et al., GAUSSIAN 03 (Revision C.02), Gaussian Inc.,
Wallingford, CT, 2004.
56 R. Ahlrichs, M. Bar, M. Haser, H. Horn and C. Kolmel, Chem.
Phys. Lett., 1989, 162, 165169.

This journal is

the Owner Societies 2010

También podría gustarte