You are on page 1of 10

RSC Advances

View Article Online

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

PAPER

Cite this: RSC Advances, 2013, 3,


4381

View Journal | View Issue

Raman scattering and DFT calculations used for


analyzing the structural features of DMSO in water and
methanol3{
Shweta Singh,a Sunil K. Srivastavab and Dheeraj K. Singh*a
We report the concentration dependent Raman spectra of neat dimethyl sulfoxide [(CH3)2SO, DMSO] and
its binary mixtures with water (W)/methanol (M) in both n(SLO) and n(CH) regions. The n(SLO) line profile
of neat DMSO was resolved into four component peaks at 1036, 1044, 1054 and 1064 cm21 and assigned
to different dimeric species of neat DMSO. A careful analysis of the Raman spectra of DMSO with water in
the n(SLO) region at different concentrations reveals that upon dilution, an additional peak is observed at
~1017 cm21 (lower side of the peak ~1036 cm21) which is attributed to the hydrogen bonding of DMSO
with water. For the highest dilution case (x = 0.1 of DMSO), hydrogen bonded species and symmetric
stretching of the dimeric n(SLO) mode were obtained which suggest that DMSO exists in dimer form even
at low concentration of DMSO. The significant blue shifting of the CH frequency due to CHO
hydrogen bonding was also obtained in the case of both solvents. Our experimental results imply the
existence of the dimer form of DMSO in neat as well as at x = 0.1 of DMSO. In order to simulate and

Received 1st November 2012,


Accepted 17th January 2013

validate our experimental findings, detailed ab initio and density functional theory (DFT) calculations were
also performed to obtain the ground state geometries of neat DMSO, self-associated dimer, trimer and
their hydrogen bonded complexes with water and methanol. Our calculated structure of DMSO in dimer

DOI: 10.1039/c3ra22730h

and trimer form reports a more accurate and stable geometry, in comparison to earlier calculation on

www.rsc.org/advances

these structures. Overall in this study nice spectrastructure correlations were obtained.

Introduction:
Non-covalent bonds, in particular, hydrogen bonds are of
fundamental importance in the molecular sciences as they are
known for their critical role in the structure and function of
biological systems.12 Hydrogen bonding between an electron
rich atom, F, N, O or Cl with the H atom of hydrogen donor
solvents has been studied39 recently by our group in different
molecules and attempts have been made to explain the
experimental results using various existing models.1012 A
substantial amount of information about the vibrational
motion and molecular association in molecular liquids can
be obtained by analyzing the experimentally measured Raman
spectra.1315
Here, we present a systematic investigation of the interactions of water and methanol with dimethyl sulfoxide (DMSO)
by means of hydrogen bonding. Binary solutions were studied
a

Laser and Spectroscopy Laboratory, Department of Physics, Banaras Hindu


University, Varanasi-221 005, India. E-mail: dheerajsingh84@gmail.com;
Tel: +919767354320
b
Department of Pure and Applied Physics, Guru Ghasidas Vishwavidyalaya, Main
Campus, Koni, Bilaspur-495009, India
3 Special dedication: our honourable research supervisor Late Prof. B. P. Asthana
{ Electronic supplementary information (ESI) available. See DOI: 10.1039/
c3ra22730h

This journal is The Royal Society of Chemistry 2013

that range from very low water and methanol content to high
content. Both experimental (Raman) and theoretical (DFT)
techniques were employed. By combining the results of both
techniques, a detailed picture of the nature of interactions was
obtained, which in turn increases our understanding of how
these solvents affect the spectral features of DMSO.
Due to the importance and vast application of DMSO in
pharmaceuticals16, the first study of the Raman spectrum of
DMSO was made as early as 1948.17 There are a few quite early
experimental studies,1820 which have dealt with structure
determination of DMSO. Raman study of solutions of DMSO in
carbon tetrachloride CCl4 and water has been also reported,21
for both parallel and perpendicular polarizations in the SLO
stretching region. A few studies on the structure of the DMSO
using different experimental techniques and in different
environments were also carried out in the 909s.2226 In a study
made almost one decade ago,27 a crystallographic and
molecular mechanics study was performed in order to
investigate the influence of hydrogen bonding on coordinated
sulfoxides and the relative role of intra- and inter-molecular
interactions in determining the DMSO orientation in
RRNHOH fac-RuCl3DMSO3 compounds was emphasized. At
the same time28 the structural properties and binding energies
of DMSORu(II) complexes were investigated using density

RSC Adv., 2013, 3, 43814390 | 4381

View Article Online

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Paper
functional theory (DFT). A flexible all-atom model of DMSO for
molecular dynamics simulations was performed by Strader
and Feller et. al.29 and they found a significant change from
previous models of DMSO, incorporating atomic flexibility and
explicit representation of hydrogen atoms. Those enhancements provide a solvent model that is consistent with modern
parameter sets for the simulation of biomolecules. In the same
year30 a quantum chemical study of 1DMSOnwater clusters
was also carried out which investigated how the clusters
interact and attempted to explain which role is played by the
various structures and their inter-cluster interaction modes in
the freezing behaviour of DMSOwater. Recently31 a Raman
and IR study of the effects of DMSO on water structure was
also investigated and the interpretation of the spectra was in
agreement with the capacity of DMSO for breaking the
structure of water in the more concentrated aqueous solutions
and enhancing it in very dilute solutions. In other recent
studies,32,33 vibrational spectroscopic methods such as IR
absorption and Raman scattering along with quantum
chemical calculations were employed to investigate molecular
association in DMSOtrichloromethane TCM and DMSO/water
mixtures respectively.
Thus, from the above brief discussion, it is clear that in
spite of several studies on the structure of DMSO as well as its
association in different solvents, a critical study on spectra
structure correlation in a systematic manner employing both
experimental and theoretical approaches is still lacking. It was,
therefore, considered worthwhile to perform ab initio and DFT
calculations on different DMSO + water (W)/methanol (M)
clusters and to correlate the structures thus obtained to the
spectral features observed in Raman measurements on DMSO
+ W/M binary mixtures with varying molar ratios of the
reference molecule, DMSO and the solvents, W/M in the
n(SLO) and n(CH) region. In the present study our theoretical
calculation provided an accurate structure of DMSO and its
possible dimer and trimer configurations which is relatively
contradictory to the earlier studies.21,23 In addition to this,
dipoledipole interaction energy was also calculated for the
different clusters in order to assess their relative strength of
interaction in the different clusters.

Experimental methods:
Sample preparations:
Liquid dimethyl sulfoxide (DMSO) and methanol obtained
from Fluka were used without any further purification.
Deionised, double distilled water was used for preparing the
mixtures. The samples were stored in an N2-atmosphere after
opening in order to avoid any contamination from the
surrounding atmosphere. Binary mixtures of DMSO + W and
DMSO + M were prepared using different mole fractions x of
DMSO in the solution. The mole fraction xi of a given species
ni
in a mixture is defined as xi ~ P , where ni is the number of
i ni
moles of the species i.

4382 | RSC Adv., 2013, 3, 43814390

RSC Advances
Instrumentation details:
A Spex-1404 double monochromator with 2400 grooves mm21
gratings and equipped with a liquid N2-cooled CCD detector
Photometric model SDS 9000 was employed to record the
spectra in n(SLO) and n(CH) regions. The 514.5 nm line of an
Ar+-laser delivering 20 mW power at the sample was used as
the Raman excitation source in 90u scattering geometry. The
data acquisition time for each spectrum was 60 s per window.
The high spectral resolution of y0.36 cm21, equivalent to an
entrance slit width of 50 mm, allows the determination of peak
positions and linewidth with high precision. For all measurements, the grating position was kept constant, which ensures
reproducibility and a reasonably high precision in the
determination of peak positions.
Raman spectral analysis:
Raman spectra of neat DMSO and six other binary mixtures of
DMSO + W and DMSO + M were fitted taking four Voigt
profiles34 and using the curve-fitting program Spectra Calc,
which is specially suited for the analysis of Raman line profiles
in order to get their multiple components. During the fitting of
the Raman line profiles, each component was assumed to be a
mixture of a Lorentzian and a Gaussian profile, which is
essentially as good as a Voigt profile.34 In order to check the
uniqueness of the fitting parameters thus obtained, each
spectrum was fitted with different reasonable initial guesses
and each guess yielded the same fitted profiles and fitting
parameters.

Theoretical methods:
Computational details:
All the calculations were performed using the Gaussian 03
program package.35 The geometry optimization and harmonic
vibrational frequencies of the different normal modes of neat
DMSO in gas phase and different clusters of DMSO with water
and methanol were calculated using the MllerPlesset many
body perturbation theory truncated at second order MP2 and
DFT with the B3LYP hybrid functional. B3LYP is a hybrid
functional consisting of Beckes exchange functional, the Lee
YangParr correlation functional and a HF exchange term. It is
a well known fact that DFT, using a non-local and gradient
corrected functional, performed equally well as the other
correlated methods, such as calculations at the MP2 level of
theory.36 In the present study the Gaussian type orbitals
(GTOs) basis set, 6-311++G(d,p)37 was employed for all the
calculations. The global minima on the potential energy
surface were confirmed by the real harmonic vibrational
wavenumbers calculated for each calculated molecule/clusters.
In the studies reported recently39 by our group, DFT
calculations with the B3LYP method provided quite nice
results in terms of spectrastructure correlation and several
meaningful conclusions were drawn from the theoretical
results.

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances

Paper

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Simulation protocol:
It is to be noted that the DFT computed calculations produced
Raman activities for the different normal modes, which cant
be in use directly as Raman intensities. The Raman scattering
cross sections, hs/hV, which are proportional to the Raman
intensities, may be calculated from the Raman scattering
amplitude and predicted wavenumber for each normal mode
using the relationship:3840
0
1

4
 4 4


C
n0 {nj
Lsj
2 p B
h
B
C


~
S
(1)
{hcnj A 8p2 cnj j
LV
45 @
1{ exp
kT
where, n0 is the exciting frequency, nj is the vibrational
frequency of the jth normal mode, Sj is the corresponding
Raman scattering amplitude obtained from DFT calculations
and h, c and k are the universal constants. In the present study
the Raman intensities obtained using this relationship match
nicely with the experimentally observed intensities (discussed
in theoretical results and discussion) which were also
performed in some of our recent studies.5,7,9,40,41

Results and discussion:


Raman spectra of DMSO in water and methanol in the n(SLO)
region
The high resolution Raman spectra of neat DMSO and six
other binary mixtures of DMSO + W and DMSO + M with mole
fractions x = 0.9, 0.8, 0.6, 0.5, 0.4 and 0.2 of DMSO, in the
n(SLO) region 9001100 cm21, were recorded and presented in
Fig. 1(a) and (b), respectively. The asymmetric peaks in the
n(SLO) stretching region observed in the Raman spectrum for
neat DMSO as shown in bottom spectra of Fig. 2(a) and (b)
have four components, at ~1035.7, 1044, 1053.8 and 1063.7
cm21, and these are assigned to the SLO symmetric stretching
mode of dimer/trimer, in-phase stretching of a cyclic dimer,
out-of-phase and free DMSO, respectively.21,32 It is to be noted
that the self-association dimer and trimer in neat DMSO are
quite likely due to the high polarity of the SLO bond and this
observation is in conformity with our DFT/ab initio calculations for the self-associated structures of DMSO, in dimer and
trimer form.
Upon adding water to DMSO, an additional peak is observed
at ~1017.5 cm21 (lower side of the peak ~1036 cm21) which is
seen at very low dilution (mole fraction of DMSO, x = 0.9 and
below) and it is confirmed by a proper line shape analysis (see
Fig. 2(a)). This new peak, which appears upon dilution by
water, is obviously due to hydrogen bonding. A closer
examination of the spectra presented in Fig. 2(a), reveals that
upon dilution, the peak corresponding to the n(SLO) mode of
DMSO in the self-associated clusters i.e. 1044, 1053.8 and
1063.7 cm21 gets weaker and almost vanishes at x = 0.1 (see
Table 1). It is also noticed that the hydrogen bonded peak
which started to appear at x = 0.9 of DMSO, goes on increasing
with dilution of water till x = 0.4, after that the intensity of
hydrogen bonded peaks become constant and finally at x = 0.1

This journal is The Royal Society of Chemistry 2013

Fig. 1 Raman spectra in the n(SLO) region for the binary mixture (a) DMSO +
water, (b) DMSO + methanol, as a function of the mole fraction x of DMSO.

of DMSO the contribution of only hydrogen bonded clusters


and the self-associated symmetric vibration of the n(SLO)
mode of dimer/trimer clusters were obtained. Thus, these
results indicate that at dilution with water (x = 0.1 of DMSO),
DMSO exists in hydrogen bonded cluster as well as its selfassociated dimeric forms. However, in propanal4, acrylonitrile6, and acetonitrile,42 the self-associated dimer/trimer were
completely vanished at high dilution and the contribution of

RSC Adv., 2013, 3, 43814390 | 4383

View Article Online

Paper

RSC Advances

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

only hydrogen bonded species and isolated molecular species


were obtained.
Upon adding methanol in DMSO the hydrogen bonded peak
and methanol peak (~1034 cm21, see Fig S1, ESI{)7 get merged
to each other and it is very difficult to differentiate the
contribution of both species separately. However, this merged
peak shows a continuous red shift upon dilution with
methanol which further confirmed the increased interaction
of DMSO with methanol in the form of hydrogen bonds (see
Fig. 2(b) and Table 2). Further at x = 0.2 of DMSO three peaks
at ~1022.7, 1038.3 and 1059.8 cm21 were obtained corresponding to the merged peak [hydrogen bonded, methanol peak and
symmetric stretching n(SLO) mode of dimer/trimer clusters],
in-phase stretching of a cyclic dimer, and out-of-phase
vibration of the dimeric n(SLO) mode, respectively. Once again
at lower concentration of DMSO, the contribution of isolated
n(SLO) mode of DMSO was not obtained which further confirm
the existence of DMSO in dimer/trimer form even at lower
concentration of DMSO (x = 0.1).
In the spectral features of DMSO + W at x = 0.1 of DMSO only
two peaks, hydrogen bonded and dimer form were obtained
whereas in the case of DMSO + M at the same concentration
three peaks were obtained (see Fig. 2). From these facts one
can easily conclude that the water solvent makes stronger
hydrogen bonding with DMSO in comparison to methanol
which is also supported by the dipoledipole interaction
energy calculation and DFT computed strength of hydrogen
bonds discussed in the theoretical parts.
In order to understand the above-mentioned difference
between DMSO/W and DMSO/M systems, as obvious from
their spectra, especially at low mole fraction x = 0.1 of DMSO,
several factors could be responsible. Firstly, concerning the
strength of hydrogen bond interaction, taking into account the
bulk property, one can recall that the boiling point of water
100 uC is higher than the boiling point of methanol 87 uC. 3
This shows that intermolecular hydrogen bond interaction in
water is stronger than that in methanol. However, this feature
depending on a bulk property alone is not enough to describe
the observed spectral difference between DMSO/W and DMSO/
M systems and other molecular-level effects need to be
considered. A second very important aspect, which needs to
be considered, is the fact that water can form two hydrogen
bonds, whereas methanol can form only one hydrogen bond. It
is obvious that at the same molar concentration, a higher
number of DMSO molecules are likely to be consumed in the
case of the DMSO/W system than that in the case of the DMSO/
M system. The steric hindrance in complex formation is
obviously much higher in case of DMSO/M complexes than in
the case of DMSO/W complexes.
Blue shifted CHO hydrogen bond of DMSO in W and M

Fig. 2 Resolved Raman spectra in the n(SLO) region for the binary mixture (a)
DMSO + water, (b) DMSO + methanol, as a function of the mole fraction x of
DMSO.

4384 | RSC Adv., 2013, 3, 43814390

The theoretical calculations revealed the new type of intermolecular binding which resembles the standard hydrogen
bonding. The shift in CH stretching frequency and the
electronic nature of this new type of hydrogen bond differ
from the classical hydrogen bonding and it is given the name
improper, blue-shifting hydrogen bond.43 Standard hydrogen bonding of the type XHY is characterized by weakening
of the XH bond which causes elongation of this bond and a

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances

Paper

Table 1 Wavenumber positions and linewidths of the four Raman peaks ~1036 cm21 (SLO symmetric stretching mode of dimer/trimer), 1044 cm21 (in-phase
stretching of a cyclic dimer), 1054 cm21 (out-of-phase) and 1064 cm21 (free DMSO) obtained by analyzing the observed Raman line profile in the n(SLO) region at
different concentrations in mole fraction, x of DMSO in the DMSO + water binary mixture

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Mole fraction DMSO


0.1
0.2
0.4
0.5
0.7
0.9
1.0

Peak HB

Peak1

Peak2

Peak3

Peak4

n cm21

CFWHM cm21

n cm21

CFWHM cm21

n cm21

CFWHM cm21

n cm21

CFWHM cm21

n cm21

CFWHM cm21

1011.9
1013.3
1015.0
1015.9
1016.4
1017.5

19.3
19.2
19.0
18.5
18.9
18.8

1026.4
1027.4
1028.5
1029.0
1031.8
1033.0
1035.7

25.0
19.6
17.9
19.4
18.9
18.6
30.1

1037.8
1044.5
1045.4
1043.4
1043.8
1044.0

40.0
19.5
20.3
14.6
13.2
12.0

1059.8
1060.0
1055.7
1054.6
1053.8

19.3
17.2
17.6
17.2
13.4

1059.8
1065.5
1063.7

red shift of the XH stretch frequency. In this study the n(SLO)


stretching mode of DMSO shows the red shift upon hydrogen
bonding with water and methanol (see Table 1 and 2) where
significant elongation was calculated in SLO bond (see Table
S1, ESI{). An improper, blue-shifting XHY hydrogen bond
is, on the other hand, characterized by strengthening of the X
H bond which causes contraction of this bond and a blue shift
of the XH stretch frequency and this blue shifted phenomenon in the CH stretch frequency of DMSO was obtained in
the case of both solvents water and methanol. The importance
of these weak and blue shifted CHO hydrogen bonds in
macromolecules, now days, is a very well-established phenomenon.3
Raman spectra of neat DMSO and five different mole
fractions, x = 0.8, 0.6, 0.5, 0.3 and 0.1 of the reference system,
DMSO in the binary mixture DMSO + W/M were recorded in
the n(CH) region 28003100 cm21 and presented in Fig. 3(a)
and (b) respectively. Raman spectra of DMSO (bottom spectra
of Fig. 3) show two peaks situated at ~2916 and ~3001 cm21 in
the 28003100 cm21 region which are due to the symmetric C
H3 stretching and anti-symmetric CH3 stretching modes.
These peaks show a significant ~11 and ~14 cm21 blue shift in
going from neat DMSO to dilution (x = 0.1 of DMSO) with
water (Fig. 3a). Upon dilution with methanol the corresponding peak shows a significant ~5 and ~4 cm21 blue shift in
comparison to the neat DMSO spectra (see Fig. 3b). The larger
shift obtained in the case of water again shows the formation
of a stronger hydrogen bond in comparison to methanol.
The blue shifted hydrogen bonds are mainly characterized
by electron density transfer from the proton acceptor molecule

17.7
17.2
17.5

O to the proton donor CH. The structural reorganization of


the proton donor framework with contraction of the CH bond
directly involved in the CHO contact and a concomitant
blue shift of its stretch frequency (for charge distribution, see
Fig. S2, S3 and S4, ESI{). Formation of the improper, blueshifting hydrogen bonding, is an indirect process where the
strengthening of the CH bond results from structural
reorganization induced by electron density transfer from the
donor to a remote part of the acceptor.43 The improper, blueshifting concept is of a general nature and stabilizes not only
intermolecular complexes but also appears in intra-molecular
systems as obtained in our calculated structures (see Fig. 4,
schemes 1, 2 and 3). Finally it is concluded that the CHO
interaction may be categorized as a true hydrogen bond which
is fully supported and fits in the criteria of the new definition
of hydrogen bonding, recently proposed and accepted by
IUPAC.44
DFT computed results and discussion:
In the binary system DMSO/W, M various different hydrogenbonded complexes can be simultaneously present. The polar
SLO group is a potential hydrogen-acceptor site in DMSO. We
therefore calculated the structures of various small DMSO/W,
M clusters with different possible stoichiometric ratios.
Optimized structures, total energies and hydrogen bonded
clusters
It is well known that, DMSO has C2v and Cs symmetry and the
structures of C2v symmetry are lower in energy than those of Cs
symmetry.45 Moreover, due to its higher symmetry it makes
the computations easier than those for asymmetric sulfoxides.

Table 2 Wavenumber positions and linewidths of the four Raman peaks ~1036, 1044, 1054 and 1064 cm21 (as assigned in Table 1) obtained by analyzing the
observed Raman line profile in the n(SLO) region at different concentrations in mole fraction, x of DMSO in the DMSO + methanol binary mixture

Mole fraction DMSO


0.2
0.4
0.5
0.6
0.8
0.9
1.0

Peak HB
21

Peak1
21

21

Peak2
21

Peak3
21

21

Peak4

n cm

CFWHM cm

n cm

CFWHM cm

n cm

CFWHM cm

n cm

CFWHM cm

n cm21

CFWHM cm21

1017.2

20.0

1022.7
1020.2
1025.9
1027.8
1028.6
1032.2
1035.7

30.2
24.0
28.3
28.8
27.1
20.9
30.1

1038.3
1037.7
1039.0
1040.3
1043.8
1043.9
1044.0

18.9
20.0
18.0
16.7
18.6
14.0
12.0

1059.8
1059.3
1056.3
1057.1
1057.6
1055.4
1053.8

27.8
24.0
26.5
24.3
13.2
16.8
13.4

1066.4
1066.5
1063.7

15.7
16.7
17.5

This journal is The Royal Society of Chemistry 2013

21

21

RSC Adv., 2013, 3, 43814390 | 4385

View Article Online

RSC Advances

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Paper

Fig. 3 Raman spectra in the n(CH) region (28503100 cm21) for the binary
mixture (a) DMSO + water, (b) DMSO + methanol, as a function of the mole
fraction x of DMSO.

Thus, taking the advantage of the work reported by Cubbage


and Jenks,45 we optimized the neat DMSO in gas phase having
C2v symmetry. The ground state optimized geometry of neat
DMSO in C2v symmetry, self-associated clusters and various
hydrogen bonded complexes with W and M along with their

4386 | RSC Adv., 2013, 3, 43814390

Fig. 4 Scheme 1: optimized geometries of (a) neat DMSO in gas phase, (b) selfassociated dimer, and (c) self-associated trimer. Scheme 2: optimized geometries of (a) two DMSO bonded to one water, (b) one DMSO bonded to one
water, (c) one DMSO bonded to two water, (d) one DMSO bonded to three
water, (e) one DMSO bonded to four water, (f) self-associated dimer bonded to
two water molecules, (g) input structure of self-associated trimer bonded to
three water molecules, (h) output structure of self-associated trimer bonded to
three water molecules. Scheme 3: optimized geometries of (a) two DMSO
bonded to one methanol, (b) one DMSO bonded to one methanol, (c) one
DMSO bonded to two methanol, (d) self-associated dimer bonded to two
methanol, (e) self-associated trimer bonded to three methanol.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

RSC Advances

Paper

optimized energy calculated by B3LYP/6-311++G(d,p) level of


theory are presented in the Fig. 4 dissected into scheme 1
(monomer, dimer and trimer), scheme 2 (water complex) and
scheme 3 (methanol complex), respectively. In addition to this,
we have also optimized some selected structures using MP2
methods also employing the same basis set for the sake of
comparison with the experimental results. The numbering
scheme used for describing the structural parameters is
presented in Fig. 4 (scheme 1a). The calculated structural
parameters: bond lengths, bond angles and dihedral angles of
neat DMSO in gas phase using MP2 and DFT methods
employing the B3LYP functional match very well with the
experimental data24 and the results are reported in Table S1,
ESI.{
We have calculated the binding energy (BE) and bond
lengths for different possible clusters of DMSO in W and M
which are shown in Table S2, ESI.{ The Basis-Set
Superposition Error (BSSE) correction (detailed in the supporting information{) was applied in the optimized structure of
the various clusters by the standard counterpoise method46,47
for the calculation of BE and it is defined as:
BE = ESA/HB

complex

2 SEindividual + EBSSE,

(2)

where, ESA/HB SA: self-associated; HB: hydrogen bonded


complex contains the ground state energy of self-associated
or hydrogen bonded complex and SEindividual contains the sum
of ground state energies of all molecules in a particular selfassociated or hydrogen bonded cluster. The binding energies
of the dimer and trimer were calculated with the help of the
above equation and their values were obtained to be 7.02 and
13.49 kcal mol21 respectively. Thus on the basis of calculated
total binding energies per DMSO, the self-associated trimer
form of DMSO appears to be more stable. Further the
hydrogen bond lengths CHO in trimer form are calculated
to be 2.09, 2.10 and 2.27 , whereas in dimer the corresponding value is 2.18 (see Fig. 4, schemes 1b and c). This shows
the strength of the hydrogen bonding interaction is stronger in
trimer form. Thus both the BE as well as the hydrogen bond
length show that the trimer is more stable. Recently our
research group6 and Alia et al.42 have also reported that the
acrylonitrile and acetonitrile molecule exists in trimer form
with larger stability comparison to dimer form. In order to
make a better agreement with the experimental results and
understanding the different aspects of self-association and
hydrogen bonding pattern, the dimer and trimer clusters with
W and M (schemes 2 and 3) were also optimized. It is very
interesting to notice that the molecular structure of trimer
clusters with W/M, almost convert in dimer form (schemes 2h
and 3e). In conclusion, it is obvious from these calculations
that the DMSO exists in dimer and trimer form, however, upon
dilution the structure of the trimer breaks and eventually turns
into the dimeric form.
The effect of solvation is one of the elementary issues in
chemistry, biology and also in certain problems of physics
because of the fact that the structure and reactivity of free
molecules are very different from those in the solvent
environment. In the present study, apart from the dimer,

This journal is The Royal Society of Chemistry 2013

trimer and their hydrogen bonded complexes with W/M


olecules, we have also performed the solvation study of
DMSO using theoretical calculations on the DMSO + W1,2,3,4
and DMSO + M1,2 complexes. For this study, we started our
calculation by bringing the SLO group of DMSO close to one
water (scheme 2a). The BE of the complex, DMSO + W1,2,3,4 is
obtained to be 8.47, 16.69, 28.68 and 38.09 kcal mol21 and for
the complex, DMSO + M1,2, it is obtained to be 8.16 and 16.25
kcal mol21 respectively. It is obvious from these calculations
that the BE of the DMSO + W is more than that of the DMSO +
M and this trend has been observed at both levels (B3LYP and
MP2) of theory used in the present study. Thus the hydrogen
bonded complex, DMSO + W is relatively more stable than
DMSO + M. It is also found that the BE of the hydrogen
bonded complexes, DMSO + W/M increase with the increasing
number of W/M molecules (see Table S2, ESI{).
By successively adding water n = 1, 2, 3, 4, with DMSO, it was
quite interesting to see that, apart from the OHO hydrogen
bond between DMSO and water, the strong hydrogen bonding
among the water themselves also plays a crucial role in
stabilizing a cluster of DMSO and water in the binary mixture.
Apart from the OHO hydrogen bond, the CHO hydrogen
bond between the DMSO and W molecule also plays a
significant role to stabilize the DMSO in a cluster of water
(scheme 2). Similar trends were also obtained in the case of
the complex of DMSO with M (scheme 3).
DipoleDipole (dd) interaction potential
In order to understand the strength of the hydrogen bond
interaction, the dd interaction potential was calculated for
each hydrogen bonded complex using the optimized geometrical parameters, such as hydrogen bond lengths, bond
angels, dipole moments for isolated DMSO, W, M and
hydrogen bonded complexes, DMSO + W and DMSO + M
obtained from DFT calculations. Using the above parameters
the potential energy of interaction V was calculated employing
the following relationship.6,8,41,48
V = msolute 6 msolvent/4pe0 6 f/r3,

(3)

Where the parameter, f = 1 2 3cos2hi, i = 1, 2 and msolute and


msolvent are the dipole moments of the solute DMSO and the
solvents W and M, respectively. The results are presented in
Table 3. It is a well known fact that when the like partial
charges on two freely rotating molecules are close enough, this
leads to repulsion, while the two unlike charges coming close
to each other leads to attraction. Thus there is a likelihood that
the repulsive and attractive interactions mutually cancel,
thereby the average values of the function f over a sphere
should be zero since it changes sign with the change in angle
h. The dd interaction potential is relatively stronger in the
case of complex, DMSO + W as compared to that in case of
DMSO + M and this trend is obtained at both levels of theory
(see in Table 3). The strong dd interaction potential in the
case of the complex, DMSO + W once again shows a relatively
stronger interaction between DMSO and W. The dd interaction potential turns out to be largest in the case of calculation
at the MP2 level and smallest at the B3LYP level in both the
complexes. The potential energy of interaction in the case of

RSC Adv., 2013, 3, 43814390 | 4387

View Article Online

Paper

RSC Advances

Table 3 Potential energies of dipoledipole interactions of DMSO with water W and methanol M calculated at B3LYP and MP2 levels of theory

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Dipole moment Debye

Hydrogen bond length

Method

mDMSO

mW

mM

hW deg

hM deg

rW

rM

Va kcal mole21
DMSO + W

B3LYP
MP2

4.403
4.387

2.159
2.257

1.886
2.025

155.00
154.63

155.78
156.06

1.821
1.809

1.826
1.797

233.24
234.95

complex, DMSO + W is higher by 3.84, and 1.71 kcal mol21 as


compared to the values for the complex, DMSO + M at B3LYP,
and MP2 levels, respectively and this also shows a stronger
interaction of DMSO to W in comparison to M molecules,
which is manifested in our experimental results also by a
larger wavenumber shift in the case of DMSO + W in
comparison to that of DMSO + M.
Correlation of experimental and theoretical results:
The experimentally recorded Raman spectrum of neat DMSO
is compared with the DFT calculated spectrum using the
B3LYP/6-311++G(d,p) method and presented in Fig. 5. It is well
known that the DFT computed harmonic vibrational wavenumbers overestimate the observed anharmonic vibrational
wavenumber.35 In view of this fact we scaled the calculated
wavenumbers with the experimental value. It is to be noted
that a unique scaling factor (0.9697) which is a standard
scaling factor at B3LYP level of theory and triple-f basis set for
the vibrational wavenumbers49 is insufficient (see Fig. 5). As it
can be seen in Fig. 5, the region 28003200 cm21 needs a
weaker scaling factor, whereas the region 2001600 cm21
needs a higher one. For the region 28003200 cm21, we have
taken the most prominent peak ~2924 cm21 as a reference
peak for scaling to the theoretical one and the scaling factor is
calculated to be 0.9615 which is very close to the standard
scaling factor. However, for the region 2001600 cm21, we
have taken the most prominent peak ~673 cm21 as a reference

Fig. 5 Comparison of experimentally observed and calculated room temperature Raman spectra of DMSO in gas phase using a DFT method employing the
B3LYP functional with basis set 6-311++G(d,p).

4388 | RSC Adv., 2013, 3, 43814390

Vb kcal mole21
DMSO + M
229.40
233.24

peak for scaling to theoretical one and the scaling factor is


calculated to be 1.0751.
In order to better understand the dimer and trimer activities
of DMSO, we have performed the comparison between
experimental Raman spectrum of neat DMSO and the
theoretical Raman spectra of dimer and trimer forms of
DMSO in the n(SLO) stretching region from 1010 to 1100 cm21
as depicted in Fig. 6. A closer examination of this spectrum
reveals that the theoretical calculated Raman spectrum of
DMSO in trimer form shows a nice correlation with the
observed Raman spectrum of neat DMSO. The theoretical
Raman spectrum of DMSO in dimer form shows two peaks,
whereas the trimer spectrum shows four peaks and the trimer
form of DMSO is relevant to our observed spectrum of DMSO.
Thus, we can again conclude that in neat form DMSO exists in
trimer form and upon dilution with solvents, it is quite likely

Fig. 6 Comparison of experimentally observed and calculated room temperature Raman spectra of DMSO in dimer and trimer form in gas phase using a DFT
method employing the B3LYP functional with basis set 6-311++G(d,p).

This journal is The Royal Society of Chemistry 2013

View Article Online

RSC Advances
that, the geometry of the trimer breaks and eventually the
complex acquires a minimum energy configuration of a dimer.

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Conclusions:
The present study was undertaken with an objective to
examine the structure of neat DMSO and the intermolecular
interaction between the DMSO and water/methanol using the
Raman spectroscopic technique and quantum chemical
calculations. The n(SLO) line profile of neat DMSO was
resolved into four component peaks at ~1036, 1044, 1054
and 1064 cm21 and assigned in different dimeric species of
neat DMSO. The concentration dependent Raman spectra
revealed that the interactions between DMSO and W/M were
made through the hydrogen bond where hydrogen bond
formation results into a variety of clusters which exhibit their
clear signature in the spectral features.
The most peculiar feature in this study is that at dilution
with water/methanol (x = 0.1 of DMSO), only hydrogen bonded
species and the dimeric n(SLO) mode were obtained which
suggest that DMSO exists in a dimer form even at low
concentrations of DMSO. The larger blue shifting of the C
HO hydrogen bonds obtained in the case of water shows the
formation of strong hydrogen bonds in comparison to
methanol.
Apart from experimental study, detailed DFT calculations
on self-associated dimer and trimer clusters of DMSO and
their hydrogen bonded complexes with W/M were also
performed. Our calculated structure of DMSO in dimer and
trimer form reports the more accurate and stable geometry, in
comparison to earlier studies on these structures.21,23 DFT
computed Raman spectra of DMSO in the n(SLO) stretching
region suggest that the DMSO exists in trimer form in neat
molecules (see Fig. 6), whereas upon hydrogen bond formation
in larger amounts the trimer form of DMSO turns into the
dimer form (as presented in Fig. 4, schemes 2h and 3e). The
hydrogen bond strengths in terms of binding energy and
dipoledipole interaction show that the DMSO + W complex
has a much stronger hydrogen bond than the DMSO + M
complex. All these features are also visible in the measured
Raman spectra.
Overall, in this contribution for the first time, using Raman
spectroscopy and DFT calculation, we demonstrated that (i)
DMSO exists in trimer form and at x = 0.1 of DMSO it remains
either in the form of hydrogen bonded clusters or dimers. (ii)
The CHO interaction plays a significant role to stabilize the
structure of the dimer, trimer and their hydrogen bonded
cluster with water/methanol. (iii) The CHO interaction
further shows the properties of a true hydrogen bond which
supports the new definition of hydrogen bonding, recently
proposed and accepted by IUPAC.44 (iv) Overall in this study
nice spectrastructure correlations were obtained.

This journal is The Royal Society of Chemistry 2013

Paper

Acknowledgements
cker, university of
SS would like to thank Prof. Sebastian Schlu
Osnabruck Germany for providing access to their laboratory
facilities to make the high resolution Raman spectra of DMSO
and its binary complex with water and methanol. Authors also
thank to Dr Ranjan K. Singh for providing the Gaussian 03
software for the theoretical calculations. DKS is thankful to the
Council of Scientific and Industrial Research CSIR, India for
the award of Nehru Post Doctorate Fellowship in Science CSIRNehru PDF. SS would like to thank the U.G.C., New Delhi for a
Research Fellowship in Science for Meritorious Students
RFSMS.

References
1 G. A. Jeffrey, An Introduction to Hydrogen Bonding, Oxford
University Press, New York, 1997.
2 G. R. Desiraju and T. Steiner, the Weak Hydrogen Bond in
Chemistry and Biology, Oxford University Press, Oxford,
1999.
3 D. K. Singh, S. Mishra, A. K. Ojha, S. K. Srivastava,
cker, B. P. Asthana, J. Popp and R. K. Singh, J.
S. Schlu
Raman Spectrosc., 2011, 42, 667675.
cker, R. K. Singh and
4 D. K. Singh, S. K. Srivastava, S. Schlu
B. P. Asthana, J. Raman Spectrosc., 2011, 42, 851858.
5 S. Singh, D. K. Singh, S. K. Srivastava and B. P. Asthana,
Vib. Spectrosc., 2011, 56, 2633.
6 D. K. Singh, S. K. Srivastava and B. P. Asthana, Chem. Phys.,
2011, 380, 2433.
7 S. Singh, D. K. Singh, S. K. Srivastava and B. P. Asthana, Z.
Phys. Chem., 2011, 225, 723740.
8 D. K. Singh, S. K. Srivastava, P. Raghuvansh, R. K. Singh
and B. P. Asthana, Vib. Spectrosc., 2011, 56, 3441.
9 D. K. Singh, S. Singh, B. P. Asthana and S. K. Srivastava, J.
Mol. Model., 2012, 18, 35413552.
10 F. Bondarev and A. I. Mardaeva, Opt. Spectrosc., 1973, 35,
286288.
11 P. C. M. Van Woerkom, J. D. Bleyser, M. de Zwart and J.
C. Leyte, Chem. Phys., 1974, 4, 236248.
12 S. F. Fischer and A. Laubereau, Chem. Phys. Lett., 1975, 35,
612.
13 B. P. Asthana and W. Kiefer, in Vibrational spectra and
structure, ed. J. R. Durig, Elsevier. Amsterdam, 1993, 20, 67
155 and references cited therein..
14 B. P. Asthana, W. Kiefer and E. W. Knapp, J. Chem. Phys.,
1984, 81, 37743778.
15 V. Deckert, B. P. Asthana, W. Kiefer, H.-G. Purrucker and
A. Laubereau, J. Raman Spectrosc., 2000, 31, 805811.
16 Johannes Geiss 2001. The century of space science. Kluwer
Academic. p. 20. ISBN 0-7923-7195-X. Retrieved 2011 Aug 7.
17 R. Vogel-Hogler, Acta Phys. Austriaca, 1948, 1, 311.
18 M. Tamres and S. Searles, J. Am. Chem. Soc., 1959, 81,
21002104.
19 S. D. Russel and S. Meek, J. Phys. Chem., 1961, 65, 1446.
20 A. Selverajan, Proc. Ind. Acad. Sci. A LXIV, 1965, 4451.
21 M. I. S. Sastri and S. Singh, J. Raman Spectrosc., 1984, 15,
8085.
22 S. A. Slivko, M. A. Sarukhanov and N. N. Kulikova,
translated from Z. Struk. Khimmi, 1993, 34, 3135.

RSC Adv., 2013, 3, 43814390 | 4389

View Article Online

Published on 18 January 2013. Downloaded by Fukui Daigaku on 15/11/2016 05:33:46.

Paper
23 S. J. Romanowski, C. M. Kinart and W. J. Kinart, J. Chem.
Soc., Faraday Trans., 1995, 91, 6570.
24 V. Typke, J. Mol. Struct., 1996, 384, 3540.
25 I. S. Perelygin, J. Struct. Chem., 1999, 38, 270.
26 H. C. Allen, D. E. Gragson and G. L. Richmond, J. Phys.
Chem. B, 1999, 103, 660666.
27 S. Geremia, M. Calligaris, Y. N. Kukushkin, A. V. Zinchenk
and V. Yu. Kukushkin, J. Mol. Struct., 2000, 516, 4956.
28 M. Stener and M. Calligaris, THEOCHEM, 2000, 497,
91104.
29 M. L. Strader and S. E. Feller, J. Phys. Chem. A, 2002, 106,
10741080.
30 B. Kirchner and M. Reiher, J. Am. Chem. Soc., 2002, 124,
62066215.
31 A. Bertoluzza, S. Bonora, M. A. Battaglia and P. Monti, J.
Raman Spectrosc., 2005, 8, 231235.
32 M. T. Khatmullina, A. S. Krauze and L. V. Ryabchuk, J.
Struct. Chem., 2007, 48, 569572.
33 O. Shun-Li, W. Nan-Nan, L. Jing-Yao, S. Cheng-Lin, L. ZuoWei and G. Shu-Qin, Chin. Phys. B, 2010, 19,
123101123107.
34 B. P. Asthana and W. Kiefer, Appl. Spectrosc., 1982, 36,
250257.
35 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M.
A. Robb, J. R. Cheeseman, V. G. Zakrzewski, J.
A. Montgomery Jr., R. E. Stratmann, J. C. Burant,
S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, M.
C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi,
R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford,
J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui,
K. Morokuma, D. K. Malick, A. D. Rabuck,
K. Raghavachari, J. B. Foresman, J. Cioslowski, J. V. Ortiz,
A. G. Baboul, B. B. Stefanov, G. Liu, A. Liashenko,
P. Piskorz, I. Komaromi, R. Gomperts, R. L. Martin, D.

4390 | RSC Adv., 2013, 3, 43814390

RSC Advances

36
37
38

39
40

41
42
43
44

45
46
47
48
49

J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng,


A. Nanayakkara, M. Challacombe, P. M. W. Gill,
B. Johnson, W. Chen, M. W. Wong, J. L. Andres,
C. Gonzalez, M. Head-Gordon, E. S. Replogle and J.
A. Pople, Gaussian 03, Gaussian, Inc., Pittsburgh PA, 2003.
M. A. McAllister, THEOCHEM, 1998, 427, 3953.
R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem.
Phys., 1980, 72, 650654.
G. A. Guirgis, P. Klaboe, S. Shen, D. L. Powell, A. Gruodis,
V. Aleksa, C. J. Nielsen, J. Tao, C. Zheng and J. R. Durig, J.
Raman Spectrosc., 2003, 34, 322336 and references cited
therein.
P. L. Polavarapu, J. Phys. Chem., 1990, 94, 81068112.
cker,
S. Singh, S. K. Srivastava, P. Donfack, S. Schlu
A. Materny and B. P. Asthana, Phys. Chem. Chem. Phys.,
2012, 14, 1431514324.
D. K. Singh, S. K. Srivastava, A. K. Ojha and B. P. Asthana, J.
Mol. Struct., 2008, 892, 384391.
J. M. Alia, H. G. M. Edwards, W. R. Fawcett and T.
G. Smagala, J. Phys. Chem. A, 2007, 111, 793804.
P. Hobza and Z. Havlas, Chem. Rev., 2000, 100, 42534264.
E. Arunan, G. R. Desiraju, R. A. Klein, J. Sadlej, S. Scheiner,
I. Alkorta, D. C. Clary, R. H. Crabtree, J. J. Dannenberg,
P. Hobza, H. G. Kjaergaard, A. C. Legon, B. Mennucci and
D. J. Nesbitt, Pure Appl. Chem., 2011, 83, 16371641.
J. W. Cubbage and W. S. Jenks, J. Phys. Chem. A, 2001, 105,
1058810595.
S. Simon, M. Duran and J. J. Dannemberg, J. Chem. Phys.,
1996, 105, 1102411031.
S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553566.
A. J. Stone, the Theory of Intermolecular Forces, Clarandon
Press, Oxford, 1996.
M. P. Andersson and P. Uvdal, J. Phys. Chem. A, 2005, 109,
29372941.

This journal is The Royal Society of Chemistry 2013

You might also like