Sunteți pe pagina 1din 16

Available online at www.sciencedirect.

com

Journal of Molecular Structure 877 (2008) 20–35


www.elsevier.com/locate/molstruc

Non-bonded interactions and its contribution to the NLO activity


of Glycine Sodium Nitrate – A vibrational approach
T. Vijayakumar a, I. Hubert Joe a, C.P. Reghunadhan Nair b, V.S. Jayakumar a,*

a
Centre for Molecular and Biophysics Research, Department of Physics, Mar Ivanios College, Thiruvananthapuram 695 015, Kerala, India
b
Polymers and Special Chemicals Division, Vikram Sarabhai Space Centre, Thiruvananthapuram 695 022, Kerala, India

Received 10 November 2006; received in revised form 9 July 2007; accepted 11 July 2007
Available online 24 July 2007

Abstract

Vibrational spectral analysis of the novel nonlinear optical (NLO) material, Glycine Sodium Nitrate (GSN) is carried out using NIR
FT-Raman and FT-IR spectroscopy, supported by Density Functional Theoretical (DFT) computations to derive equilibrium geometry,
vibrational wave numbers and first hyperpolarizability. The reasonable NLO efficiency, predicted for the first time in this novel com-
pound, has been confirmed by Kurtz–Perry powder SHG experiments. The influence of Twisted Intramolecular Charge Transfer (TICT)
caused by the strong ionic ground state hydrogen bonding between charged species making GSN crystal to have the non-centrosymmet-
ric structure has been discussed. The shortening of CAH bond lengths, blue-shifting of the stretching frequencies and intensity variation
indicating the existence of ‘blue-shift or improper’ CAH  O hydrogen bonding. The intense low wavenumber H-bond Raman vibrations
due to electron–phonon coupling and non-bonded interactions in making the molecule NLO active have been analyzed based on the
vibrational spectral features. The Natural Bond Orbital (NBO) analysis confirms the occurrence of a strong intra- and intermolecular
NAH  O and CAH  O hydrogen bonds.
 2007 Elsevier B.V. All rights reserved.

Keywords: NIR FT-Raman; FT-IR; Ionic hydrogen bonds; Nonlinear optics; SHG; First hyperpolarizability; DFT; Ab initio computations; Twisted
Intramolecular Charge Transfer (TICT); Glycine conformers; Natural Bond Orbital analysis

1. Introduction crystal structure and its optical nonlinearities [6–8]. Chiral-


ity [9] and hydrogen bonding [10] are important factors in
Nonlinear optical (NLO) materials are active elements the design of NLO chromophores since the measurement
for optical communications, optical switching data storage of bulk NLO properties in the solid state is dependent on
technology, optical mixing and electro-optic application a non-centrosymmetric packing environment within the
[1–4]. The development of photonic and optoelectronic crystal lattice. Molecules in pure organic crystals are often
technologies rely heavily on the growth of NLO materials coupled by relatively weak van der Waals forces or hydro-
with high nonlinear optical responses and the development gen bonding, resulting in rather poor mechanical proper-
of novel and more efficient materials [5]. To design and fab- ties. In such organic crystals, two requirements to be
ricate the NLO materials, much effort is being devoted to satisfied are: (i) they are made of highly polarizable mole-
understand the origin of nonlinearity in large systems and cules, the so-called conjugated molecules, where the asym-
to relate NLO responses to electronic structure and molec- metric p electron system of aromatic molecules can easily
ular geometry. The molecular engineering approach has led move between a donor and an acceptor substituent groups
to better understanding of the relationship between the that induce a molecular charge transfer and (ii) the mole-
cules are adequately packed to build up a non-centrosym-
*
Corresponding author. Tel.: +91 471 2530887. metric crystal structure that provides non-vanishing
E-mail address: vsjk@vsnl.net (V.S. Jayakumar). second order nonlinear coefficients [11,12]. Metal-organic

0022-2860/$ - see front matter  2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.molstruc.2007.07.021
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 21

hybrids also offer interesting advantages with respect to ous identification of vibrational modes and provide deeper
both pure organics and pure inorganic materials in that insight into the bonding and structural features of complex
they may be reliably designed by integrating highly predict- organic molecular systems. Glycine Sodium Nitrate (GSN)
able structural features, such as hydrogen bonds and coor- crystals, being the first complex of glycine reported with
dination bonds, in each of these cases which are used NLO property, the vibrational spectral studies of this novel
jointly to achieve spatial and dimensional control in NLO system is taken up, based on NIR FT-Raman and IR
organic–inorganic hybrid [13]. spectra along with DFT and MP2 theoretical support to
Presently, semi organics are being explored, which can elucidate the relationship between the molecular structural
share the properties of both organic and inorganic materi- features and NLO properties.
als, such as high laser damage threshold, optical transpa-
renc and high efficiency. Complexes of amino acids with 2. Experimental
inorganic salts have been of interest as materials for optical
second harmonic generation (SHG), and all amino acids 2.1. Preparation
except glycine contain chiral carbon atoms and perhaps
crystallize in the non-centrosymmetric space group [8]. Glycine Sodium Nitrate (GSN) crystals grown by slow
Dipolar molecules possessing an electron donor group evaporation [40] were subjected to repeated recrystalliza-
and an electron acceptor group contribute to large second tion and good quality single crystals with size of around
order optical nonlinearity arising from the intramolecular 0.6 mm were obtained.
charge transfer between the two groups opposite nature.
Due to this dipolar nature, amino acids have been consid- 2.2. Crystal structure
ered potential candidates for NLO applications [6,8].
Although the salts of amino acids like L-Arginine [14], L- GSN (Fig. 1) crystallizes in monoclinic space group Cc
Histidine [15] and L-Proline [7,16] are reported to have with four formula units in unit cell (Z = 4) [40]. The cell
NLO properties; the complexes of glycine with inorganic dimensions are: a = 14.329(3) Å, b = 5.2662(11) Å,
salts are not explored for optical SHG so far, since glycine, c = 9.1129(18) Å, b = 119.10(3). The glycine molecules
the simplest amino acid, does not possess the asymmetric are seen ‘sandwiched’ between layers of Na (NO3) as
carbon, it is NLO inactive. Out of the number of semi shown in Fig. 2. Both carboxyl O atoms participate in
organic single crystals of glycine that have been already the hydrogen bonds as acceptors forming head-to-tail
reported, most of them are not NLO active [17–19]. hydrogen bonds. Almost linear OANaAO chains involving
Glycine has three polymorphic crystalline forms a, b, c carbonyl O atoms run along the (20 2) plane in the [1 0 1]
[20,21]. Both a and b forms crystallize in centrosymmetric direction. X-ray powder diffraction was used for the identi-
space groups ruling out the possibility of optical second fication of the grown crystals of GSN. Usually, the charac-
harmonic generation. But c-glycine crystallizes in non-cen- teristic strong peaks of a-glycine and c-glycine are expected
trosymmetric space groups P31 making it a possible near 2h values of 29 and 24, respectively [41].
candidate for NLO applications and it is difficult to grow
the c-glycine crystals [22,23]. The thermodynamic stabili- 2.3. Raman and IR measurements
ties of the three polymorphs of glycine at room tempera-
ture are in the order c > a > b [24]. It has recently been The NIR FT-Raman spectrum (Fig. 3) of GSN was
reported that complexes of the c-glycine can be efficient obtained on a Bruker RFS 100/S FT-Raman Spectrometer
in optical SHG with inorganic salt sodium nitrate [8].
Due to their potential applications in photonic devices,
bulk NLO properties of materials as well as their depen-
dence on the first hyperpolarizabilities of molecules have
evoked a lot of experimental efforts [25–27] and theoretical
research [28–30]. The Natural Bond Orbital (NBO) analy-
sis can be employed to identify and substantiate the possi-
ble intra- and intermolecular interactions between the units
that would form the H-bonded network [31]. Vibrational
spectral studies of the molecules can be used to provide
deeper knowledge about the relationships between molecu-
lar architecture, nonlinear response and hyperpolarizabil-
ity. NIR FT-Raman spectra combined with quantum
chemical computations have recently been effectively
applied in the vibrational analysis of drug molecules [32],
biological compounds [33,34], natural products [35,36]
and NLO active compounds [7,37–39], since fluorescence
free Raman spectra and computed results help unambigu- Fig. 1. Optimized molecular structure of GSN.
22 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

2.4. Second harmonic generation (SHG) efficiency


measurements

Second harmonic generation from microcrystalline pow-


ders of GSN was examined using Kurtz–Perry powder
SHG method [42]. Particle sizes, ranging from 100 to
300 lm, graded using standard sieves were used for the
study. Samples were loaded in glass capillaries having an
inner diameter of 600 lm. The fundamental beam
(1064 nm) of a Q-switched ns-pulsed (6 ns, 10 Hz) Nd:
YAG laser (Spectra Physics model INDI-40) was used.
The second harmonic signal was collected using appropri-
ate optics and detected using a monochromator, Photo
Multiplier Tube (PMT) with a boxcar integrator and oscil-
loscope (Tektronix model TDS 210, 60 MHz). Filters were
Fig. 2. Intermolecular interactions (shown by dotted lines) in GSN
used to bring the signals for all the sample size in the same
crystal.
range. Urea with particle size of P150 lm was used as the
reference and the calibration measurements were carried
with a liquid nitrogen-cooled Ge-diode detector using Nd: out using N-(4-Nitrophenyl)-(L)-Prolinol (NPP). The
YAG laser at 1064 nm of 300 mw output as the excitation SHG efficiency of GSN was evaluated (Table 1) to be an
source with the powder sample in a capillary tube. Thou- average of 0.3 times that of urea and the Fig. 5 describes
sand scans were accumulated with a total registration time that GSN crystal is reasonably phase matchable.
of about 30 min. The spectral resolution after apodization
was 4 cm1. A correction according to the fourth-power 3. Computational
scattering factor was performed, but no instrumental cor-
rection was made. The upper limit for the wave numbers The complete geometry optimizations and normal-
is 3500 cm1 owing to the detector sensitivity and the lower mode analysis were performed employing the Becke–
limit is around 50 cm1 owing to the Rayleigh line cut-off Lee–Yang–Parr hybrid exchange-correlation three-parame-
by the notch filter. The IR spectrum of GSN in the region ter functional (B3LYP) [43], Moller–Plesset second order
600–4000 cm1 (Fig. 4) was recorded using a TENSOR-27 perturbation (MP2) and ab initio computations to com-
MICRO ATR accessory MIRACLE, Pike; ZnSe crystal pare the accuracy of different kinds of computational
FT-IR spectrometer with the sample in KBr matrix. The methods. Molecular geometries were fully optimized by
lower region available with the spectrum is up to Berny’s optimization algorithm using redundant internal
600 cm1.The resolution is about 2 cm1 and 300 scans coordinates and confirmed to be minimum energy con-
were used. formations. The special basis set LANL2DZ has been

Fig. 3. NIR FT-Raman spectrum of GSN.


T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 23

Fig. 4. FT-IR spectrum of GSN.

Table 1 surface (PES) and the analytic second derivative of


Dependence of SHG efficiencies with particle size energy leads to the vibrational frequencies. The calcu-
S.No Average particle size (lm) SHG efficiency (·Urea) lated harmonic vibrational frequencies were uniformly
1 125 0.35 scaled down [44], to account for systematic errors caused
2 175 0.36 by basis set incompleteness, neglect of electron correla-
3 225 0.25 tion and vibrational anharmonicity. All theoretical calcu-
4 275 0.29
lations were carried out using GAUSSIAN ’98 program
package [45] and the theoretical Raman and IR spectra
adopted with Hatree–Fock (HF), Density Functional are shown in Figs. 6 and 7.
Theory (DFT) and Moller–Plesset second order perturba- The first hyperpolarizability (b0) of this of novel molec-
tion (MP2) methods due to the existence of metal atom ular system and related properties (b, l) of GSN are calcu-
Na in GSN. The self-consistent field equation has been lated using standard basis set. In the presence of an applied
solved iteratively to reach the equilibrium geometry cor- electric field, the energy of a system is a function of the elec-
responding to the saddle point on the potential energy tric field. First hyperpolarizability is a third rank tensor

0.38

0.36

0.34
SHG Intensity (X Urea)

0.32

0.30

0.28

0.26

0.24
120 140 160 180 200 220 240 260 280
Particle size (μm)

Fig. 5. Dependence of the second harmonic intensity on the particle size.


24 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

Fig. 6. Theoretical Raman spectrum from HF/LANL2DZ.

Fig. 7. Theoretical IR spectrum from HF/LANL2DZ.

that can be described by a 3 · 3 · 3 matrix. The 27 compo- E ¼ Eo  li F i  1=2aij F i F j  1=6bijk F i F j F k


nents of the 3D matrix can be reduced to 10 components
 1=24cijkl F i F j F k F l þ . . . :
due to the Kleinman symmetry [46]. The components of
b are defined as the coefficients in the Taylor series expan- where Eo is the energy of the unperturbed molecules, Fi is
sion of the energy in the external electric field. When the the field at the origin li,aij, bijk and cijkl are the components
external electric field is weak and homogeneous, this of dipole moment, polarizability, the first hyperpolarizabil-
expansion becomes. ities and the second hyperpolarizabilities, respectively.
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 25

4. Optimized Geometries with LANL2DZ basis set show the corresponding bond
lengths are 1.518 and 1.564 Å whereas with 6-311G (d, p)
The optimized structural parameters of GSN for differ- basis set, the bond lengths are 1.502 and 1.556 Å. In
ent schemes of computations are given in Tables 2–4. The GSN, the bond lengths N1AC2 and C2AC3 are measured
corresponding values from X-ray diffraction are also given to be 1.480 and 1.52 Å. In a-glycine, c-glycine and GSN,
for comparison. The optimized molecular structure of the the experimental bond lengths are rather lower than the
compound with atom numbering scheme adopted in the data measured by the different theoretical methods includ-
computations is shown in Fig. 1. From Tables 2–4, it can ing the special basis set LANL2DZ. All experimental and
be seen that there are some deviation in the computed geo- theoretical bond lengths and bond angles of the GSN crys-
metric parameters from the corresponding XRD data [40] tal are reasonably comparable with the a-glycine rather
and these differences are probably due to the intermolecu- than c-glycine. All dimensions of the c-glycine are very
lar interactions in the crystalline state. All the bond lengths close to that of other forms, a and b, with the exception
of H atoms determined experimentally by the XRD of the rather short C3AO4 bond lengths of 1.237 Å. This
method are 0.94 Å. The correct value determined by the may be interpreted partly due to the shortening caused
neutron diffraction method would be 1.08 Å [47]. by the angular oscillations of the oxygen atom around a
The bond lengths N1AC2 and C2AC3 corresponding to centre near the C3 atom [23]. The C3AO4 and C3AO10
the a- and c-glycine are measured to be 1.474, 1.523 Å bond lengths are measured to be 1.242 and 1.247 Å in
and 1.460, 1.53 Å, respectively [22]. DFT computations GSN and the corresponding bond angles for the a-glycine

Table 2
Optimized bond lengths (Å) in GSN
Bond Experimental B3LYP/6-311G(d, p) BLYP/6-31G(d) CBS-4 HF/LANL2DZ B3LYP/LANL2DZ MP2/LANL2DZ
N1AC2 1.480 1.502 1.520 1.520 1.510 1.518 1.543
N1AC5 0.890 1.036 1.050 1.016 1.008 1.055 1.042
N1AH6 0.890 1.017 1.030 1.010 1.006 1.022 1.027
N1AH7 0.890 1.060 1.082 1.052 1.027 1.053 1.062
C2AC3 1.520 1.556 1.572 1.563 1.547 1.564 1.568
C2AH8 0.970 1.087 1.096 1.079 1.078 1.091 1.098
C2AH9 0.970 1.091 1.100 1.078 1.078 1.094 1.101
C3AO4 1.242 1.241 1.260 1.215 1.232 1.258 1.296
C3AO10 1.247 1.261 1.283 1.281 1.282 1.318 1.320
N5AO11 2.062 1.664 1.659 1.621 1.713 1.677 1.677
O11AN12 1.241 1.301 1.336 1.345 1.306 1.347 1.368
N12AO13 1.235 1.217 1.242 1.220 1.225 1.266 1.297
N12AO14 1.247 1.256 1.280 1.288 1.283 1.314 1.326
O14ANa15 2.615 2.355 2.370 2.193 2.334 2.391 2.450

Table 3
Optimized bond angles () in GSN
Bond angle Experimental B3LYP/6-311G(d, p) BLYP/6-31G(d) CBS-4 HF/LANL2DZ B3LYP/LANL2DZ MP2/LANL2DZ
N1AC2AC3 111.92 104.28 103.57 107.88 109.85 105.85 104.06
H5AN1AH6 109.41 112.94 113.32 111.76 109.94 114.49 112.68
H5AN1AH7 109.43 103.92 103.35 103.88 104.18 104.09 105.33
H6AN1AH7 109.46 112.47 112.70 110.63 108.77 112.22 112.00
H5AN1AC2 109.55 103.32 102.61 106.36 109.24 101.24 104.71
H6AN1AC2 109.49 115.31 115.49 113.32 112.28 114.63 114.00
H7AN1AC2 109.48 107.87 108.22 110.41 112.11 109.15 107.47
C2AC3AO4 117.76 117.62 117.76 116.27 116.19 118.67 118.57
C2AC3AO10 116.19 113.70 113.653 112.83 113.60 112.55 114.06
H8AC2AC3 109.21 111.26 111.78 110.52 109.88 110.16 110.96
H8AC2AN1 109.18 109.74 109.45 108.64 108.60 110.39 109.36
H9AC2AC3 109.25 111.66 112.57 110.14 110.12 110.25 112.08
H9AC2AN1 109.22 109.11 108.61 109.37 109.19 110.30 108.69
O4AC3AO10 126.03 128.30 128.13 130.89 130.19 128.75 126.96
N1AH7AO11 154.90 146.35 149.59 164.94 155.83 143.73 145.64
H7AO11AN12 131.55 132.52 119.41 115.74 148.81 147.65 142.39
O11AN12AO13 128.46 120.41 120.07 120.25 121.12 121.21 120.80
O11AN12AO14 119.07 115.94 116.06 115.99 115.75 114.96 115.58
O13AN12AO14 120.47 123.65 123.86 123.75 123.14 123.84 123.626
N12AO14ANa15 97.07 94.65 92.348 93.07 98.55 95.68 94.685
26 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

Table 4
Optimized torsion angles () in GSN
Dihedral angle Experimental B3LYP/6-311 G(d, p) BLYP/6-31G(d) CBS-4 HF/LAN L2DZ B3LYP/LANL2DZ MP2/LANL2DZ
N1AC2AC3AO4 6.64 31.83 32.663 6.661 8.300 15.427 33.270
N1AC2AC3AO10 175.23 141.71 140.17 174.01 172.61 162.56 139.91
H5AN1AC2AC3 60.03 40.62 41.799 38.845 51.957 28.482 44.027
H6AN1AC2AC3 179.97 164.33 165.585 162.03 174.19 152.26 167.61
H7AN1AC2AC3 59.98 69.03 67.033 73.235 63.009 80.894 67.636
H8AC2AC3AO4 114.38 150.05 150.40 111.98 111.11 134.76 150.78
H8AC2AC3AO10 68.75 23.478 22.438 67.350 67.981 43.223 22.405
H9AC2AC3AO4 127.73 85.855 84.461 125.98 128.60 103.85 83.995
H9AC2AC3AO10 54.13 100.61 102.71 54.694 52.313 78.159 102.82
H8AC2AN1AH7 61.06 50.234 52.307 46.596 57.180 38.296 50.973
H9AC2AN1AH7 178.91 171.55 173.10 166.96 176.14 159.86 172.78
C2AN1AH7AO11 32.29 60.784 65.338 96.348 138.46 78.237 57.552
N1AH7AO11AN12 4.42 135.12 125.51 126.55 175.24 151.49 152.44
H7AO11AN12AO13 76.80 37.407 42.322 65.838 74.083 32.521 22.291
H7AO11AN12AO14 103.93 142.44 137.22 113.12 106.00 147.23 157.90
O11AN12AO14ANa15 4.03 13.095 21.009 16.278 5.640 5.128 6.897
O13AN12AO14ANa15 176.69 166.75 158.51 162.64 174.45 174.61 173.30

and c-glycine are reported to be 1.252, 1.255 Å and 1.254, angles are computed to be 138.46 and 96.35 from HF/
1.237 Å, respectively. Furthermore, the bond angle corre- LANL2DZ and CBS-4 calculations. It has been observed
sponding to the glycine skeleton molecule in a-form and from the computed and measured dihedral angles that
c-form is found to be 111.8 for and 107, respectively, the deviations of few dihedral angles are mainly due to
and it is evident that the twist about CAN bond in c-gly- the influence of metal atom Na coordination with oxygen
cine makes the c-form asymmetric. It is interesting to note atoms of both nitrate and carbonyl groups which distorted
that the experimental and theoretical bond lengths and or twisted the geometry of GSN molecule in the crystal.
bond angles of the glycine molecules in GSN are quite
comparable to that of the a-glycine than c -glycine, though 5. Natural Bond Orbital analysis
the c-form preferably crystallizes in non-centrosymmetric
structure resulting to be NLO active [41]. NBO analysis is proved to be an effective tool for chem-
The GSN molecule was primarily optimized using stan- ical interpretation of hyperconjugative interaction and elec-
dard B3LYP, BLYP and HF levels at different split valence tron density transfer (EDT) from filled lone electron pairs
basis sets 6-31G(d, p), 6-31G(d). The dihedral angles of the n (Y) of the ‘‘Lewis base’’ Y into the unfilled anti-
N1AC2AC3AO4 and N1AC2AC3AO10 of the glycine skele- bond r* (XAH) of the ‘‘Lewis acid’’ XAH in XAH  Y
ton in GSN are calculated to be 32.6 and 140.3 for the hydrogen bonding systems [48]. To elucidate intermolecu-
DFT (B3LYP and BLYP) calculations with 6-31G(d) basis lar hydrogen bonding, intermolecular charge transfer
set and 6.6 and 174.01 for the CBS-4 calculation. With (ICT), rehybridization, delocalization of electron density
the special basis set LANL2DZ, the same dihedral angles and cooperative effect due to n (O) fi r* (NAH), the
are found to be 8.3 and 172.6, 15.4 and 162.6, NBO analysis has been performed on GSN and glycine,
32.2 and 139.9 corresponding to the HF, B3LYP and and the corresponding results are presented in Tables 5
MP2 computations, respectively, while the experimentally and 6. The intermolecular NAH  O hydrogen bonding
measured values are 6.6 and 175.2. It is also inferred is formed by the orbital overlap between the n (O) and r*
that the glycine skeleton in GSN is reasonably twisted by (NAH) which results ICT causing stabilization of the H-
around 6.6 than the glycine zwitterion where the same bonded systems. Hence hydrogen bonding interaction leads
dihedral angle is found to be 0.4 that reveals the glycine to an increase in electron density (ED) of NAH anti-bond-
molecule in GSN is twisted. The optimized structures of ing orbital. The increase of population in NAH anti-bond-
GSN based on HF/LANL2DZ and CBS-4 calculations ing orbital weakens the NAH bond. Thus the nature and
are observed to be quite comparable with the experimental strength of the intra- and intermolecular hydrogen bonding
measured values than other DFT and MP2 calculations. can be explored by studying the changes in electron densi-
Although the dihedral angles of the GSN molecule in ties in vicinity of N  H hydrogen bonds. The NBO analy-
HF/LANL2DZ calculation are more similar than the sis clearly shows the existence of strong NAH  O
CBS-4 calculation with the experimental values, slight devi- intermolecular hydrogen bonding in GSN. This investiga-
ations of few dihedral angles are found mainly between the tion obviously clarifies the formation of H-bonded interac-
glycine cation and nitrate anion. The dihedral angle tion between n (O11) and r* (N1AH7) anti-bonding
C2AN1AH7AO11 between glycine cation and nitrate anion orbitals. The difference in stabilization energy E(2) associ-
are measured to be 26.33 whereas the same torsional ated with the hyperconjugative interaction n1 (O11) fi r*
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 27

Table 5 boxylate groups, skeletal modes, and low frequency


Second order perturbation theory analysis of Fock matrix in NBO basis hydrogen bonds.
Donor NBO Acceptor E(2) kcal mol1 E(j)  E(i) F(i, j)
(i) NBO (j) a.u. a.u. 6.1.1. Amino vibrations
LP1O11 r*N1AH7 16.31 1.19 0.125 In saturated amines, the asymmetric NH2 stretch and
LP2O11 r*N1AH7 1.64 0.69 0.030 their symmetric counterpart are usually expected in the
LP3O11 r*N1AH7 7.24 0.68 0.067
LP2O10 r*N1AH6 0.88 0.63 0.022
region 3380–3350 cm1 and 3310–3280 cm1, respectively
(GSN) [50]. However, the protonation of NH2 group can shift in
LP2O10 r*N1AH6 1.14 0.56 0.023 band position towards the range 3300–3100 cm1 and
(Glycine) 3100–2600 cm1 for asymmetric and symmetric stretching
E(2), energy of hyperconjugative interactions (stabilization energy); modes, respectively, as observed in glycine derivatives
E(j)  E(i), energy difference between donor i and acceptor j NBO orbi- [51]. The NHþ 3 stretching bands are broader and weaker
tals; F(i, j), Fock matrix element between i and j NBO orbitals. in IR than those arising from the uncharged NH2 groups.
The asymmetric stretching mode of the NHþ 3 group
(N1AH7), n2 (O11) fi r* (N1AH7) and n3(O11) fi r* appears at 3241 and 3244 cm1, respectively, in the IR
(N1AH7) are 16.31, 1.64 and 7.24 kcal mol1, respectively, and Raman spectra as weak bands. Furthermore, the posi-
which is due to the accumulation of electron density in the tion and broadness of this mode NHþ 3 asymmetric stretch-
NAH bond drawn not only from n (O) of the hydrogen ing frequency indicates the formation of both intra- and
acceptor but from the entire molecule leading to its elonga- intermolecular strong NAH  O hydrogen bonding of the
tion and concomitant red shift of the NAH stretching NHþ 3 group with oxygen of both the carbonyl group and
wavenumber [49]. the inorganic nitrates. The presence of strong NAH  O
intra- and intermolecular hydrogen bonding is also evident
6. Vibrational Spectral Analysis from the lowering of the NHþ 3 symmetric stretching fre-
quency to 2886 cm1 in IR spectrum. Table 8 explains
The vibrational spectral analysis is performed based on the hydrogen bonding geometry of the GSN crystal which
the characteristic vibrations of the glycine molecule and clearly explains the existence of strong intra- and intermo-
inorganic nitrate separately. The bands observed between lecular NAH  O hydrogen bonding. The bond distance
3500–300 cm1 arise from the internal modes of the glycine and bond angle between the N1 atom of the glycine mole-
molecules and the internal modes of inorganic nitrates cule and the O atoms of nitrate group and of carboxylate
while the bands below 300 cm1 occur due to the external group are around 2.724, 3.003 Å and 146, 113, respec-
modes of glycine molecule, the liberational and transla- tively, resulting in strong NAH  O intramolecular hydro-
tional modes of Na and the vibrations of the low frequency gen bonding between the ionic spices, which might provide
hydrogen bonds. The computed vibrational wave numbers, the non-centrosymmetric structure for the GSN crystal.
their IR and Raman activities and the atomic displace- The crystal structure [40] and optimized geometries of
ments corresponding to the different normal modes are GSN show that the organic molecular units are located
used to identify the vibrational modes unambiguously. between layers of Na (NO3) chains and linked to sodium
The calculated vibrational wave numbers, measured infra- nitrate by strong intramolecular hydrogen bonds of the
red and Raman band positions and their tentative assign- N+AH  O type. This structural organization of infinite
ments are presented in Table 7. chains of highly polarizable entities connected in a head-
to-tail arrangement in GSN is favourable in contributing
6.1. Glycine vibrations to the NLO properties of the crystal. Another structural
feature of interest in this material is the role of Sodium
The internal vibrations of the glycine molecule are dis- atom in the inorganic sodium nitrate to make GSN crystal
cussed based on the vibrations of amino, methylene, car- NLO active, though the inorganic salts are not forming

Table 6
NBO results showing the formation of Lewis and non-Lewis orbitals
Bond (AAB) ED/Energy (a.u.) EDA% EDB% NBO S% P%
rN1AH7 1.98985 76.92 23.08 2.75
0.8770(sp ) N + 0.4804 (s) H 26.66 73.27
rN1AH6 1.99393 72.34 27.66 0.8505(sp3.38) N + 0.5259 (s) H 22.79 77.11
rN1AH5 1.99327 73.88 26.62 0.8566(sp3.48) N + 0.5160 (s) H 22.32 77.59
r*N1AH7 0.06090 23.08 76.92 0.4804(sp2.75) N  0.8770 (s) H 26.66 73.27
r*N1AH6 0.01209 27.66 72.34 0.5259(sp3.38) N  0.8505 (s) H 22.79 77.11
r*N1AH5 0.01288 26.62 73.38 05160(sp3.48) N  0.8566 (s) H 22.32 77.59
LP2O35 1.94784 sp0.27 78.43 21.56
LP2O35 1.92749 sp99.99 0.68 99.25
LP2O35 1.70094 sp99.99 0.27 99.63
28 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

Table 7
Calculated vibrational wavenumbers, measured infrared and Raman band positions (cm1) and assignments for GSN
HF/LANL2DZ mIR/cm1 mRaman/cm1 IR intensity Raman activity Force constant Depol. Assignment
mcal/cm1 (Absolute) (Absolute) (mdyne/Å) ratio
3392 3241 wbr 3244 vwbr 48.12 129.98 7.92 0.31 NHþ3 asym stretch
3016 3026 wsh 3023 w 124.8 43.890 6.64 0.12 CH2 asym stretch
2947 2959 wbr 2969 m 4.460 103.66 5.96 0.13 CH2 sym stretch
2810 2880 wbr 2886 vw 1110.8 154.06 5.19 0.27 NAH  O sym stretch
2802 wsh Overtones/combinations
2714 vw 2723 vw
2625 vw 2615 vw
2440 vw
2405 vw
2009 vwbr Combination of NHþ 3 asym
bend and torsion
1768 vw Combination m1 and m4 of NO 3
1631 1615 m 1616 vw 322.82 1.83 7.56 0.72 NHþ 3 asym bend
1623 1579 s 199.62 3.07 2.08 0.75 COO asym stretch
1517 1506 m 1510 vw 137.87 4.57 2.02 0.72 NHþ 3 sym bend
1460 1449 m 1449 w 120.73 7.60 1.43 0.68 CH2 scissoring
1433 1414 ssh 1408 wsh 143.53 8.74 1.56 0.74 COO sym stretch
1389 1353 vvs 287.63 9.69 5.10 0.72 NO 3 asym stretch
1320 1328 w 146.48 3.61 5.40 0.42 CH2 wagging
1310 1307 vs sh 1309 vw 146.48 3.61 5.40 0.42 CH2 wagging
1151 1137 m 1143 vw 263.59 4.97 2.18 0.75 CH2 twisting
1120 1116 s 1118 vw 46.710 4.44 1.09 0.50 NHþ 3 rocking
1084 1068 vw 4.840 3.31 0.91 0.67 CH2 rocking
1072 1052 vvs 114.34 23.11 9.13 0.10 NO 3 sym stretch
1016 1039 m 1039 wsh 13.74 3.64 1.68 0.63 CAN stretch
940 935 vs 935 vw 29.96 1.62 0.80 0.46 CH2 rocking
902 890 s 895 m 44.45 6.92 2.69 0.05 CAC stretch
829 829 s 16.03 0.27 5.66 0.60 cNO 3 out of plane deform
741 721 m 4.61 1.60 4.57 0.68 COO deform
686 676 s 677 vw 3.641 1.55 2.03 0.69 dNO 3 in-plane deform
598 586 vw 16.161 2.40 1.69 0.70 COO deform
515 509 w 34.651 2.42 0.43 0.40 COO rocking
418 399 vw 19.291 0.35 0.18 0.45 NHþ 3 torsion
323 331 vw 48.151 1.22 0.12 0.22 CCN bending
213 178 msh 42.011 0.69 0.43 0.24 Na+ translation
143 138 ssh 5.451 0.65 0.16 0.72 COO torsion
107 109 vs 3.261 0.58 0.10 0.69 N    O vibrations=
NO 3 torsion

Table 8 It has been recently explored that the ground state


Hydrogen bonding geometry hydrogen bonding between the electron donor and the elec-
DAH  A DAH H  A D  A DAH  A tron acceptor as the major cause of the Twisted Intramo-
N1AH5  O10 1.036 1.967 3.003 112.7 lecular Charge Transfer (TICT) state formation by means
N1AH7  O11 1.060 1.664 2.724 146.3 of ‘promotion effect’ [52]. The analogues of the material
N1AH7  O11a 0.890 2.062 2.952 154.9 contain an electron donor in the form of amino groups,
N1AH5  O4a 0.890 2.030 2.789 142.0 and therefore, already in their ground states they are apt
N1AH6  O10a 0.890 1.970 2.780 151.0
C3AH14  O16a 0.970 2.540 3.256 131.0
in the form of hydrogen bond with excellent electron accep-
a
tors. In the excited ICT state, the H-bond to the positively
Indicates intermolecular hydrogen bonding taken from ref. [40].
charged amino group is expected to break, and new H-
bonds are expected to form at the sites of high electron
non-centrosymmetric crystals with glycine. The amino density, on breaking the H-bond in the excited state, a pre-
group of the glycine can be act as a donor and both the condition for the twist [52]. As discussed earlier in the pre-
oxygen atoms of the carboxylate group and the nitrate vious section that the twist about CAN bond in c-glycine
group can be act as acceptors, which are connected by makes the c-form asymmetric unlike the a-form of the gly-
the strong intramolecular ionic hydrogen bonds, are clearly cine which normally does not have any twist about the
substantiated by the DFT computations. The presence of skeleton of glycine molecule. XRD and DFT investigations
strong NAH  O intra- and intermolecular hydrogen are quite favourable to the existence of the a-form rather
bonding is also substantiated by the NBO analysis. than c-form in GSN. However, the twist of the glycine skel-
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 29

eton of a-form, which is pre-requisite to make the glycine [58,59]. It has been shown that ECCF from infrared spec-
molecule NLO active, may be introduced by the Twisted trum account quantitatively for several intramolecular elec-
Intramolecular Charge Transfer (TICT) due to the pres- tronic effects including induction and backdonation. For
ence of strong ionic intra- and intermolecular N+AH  O the molecules in which induction produces stronger polar-
hydrogen bonding. ization of CAH bonds, along with the increase of both
The NHþ 3 asymmetric and symmetric bending vibrations CAH force constant and charge of the hydrogen atom
are generally expected near 1660–1610 cm1 and 1550– and decrease of CAH stretching intensity and CAH bond
1485 cm1, respectively [51]. The NHþ 3 asymmetric defor- length, it can cause the enhancement of vibrational wave-
mation mode identified in IR as medium band at number of CAH stretching modes. The methylene group
1615 cm1 and as weak Raman band at 1616 cm1, unam- hydrogen atoms in GSN are subjected to the electronic
biguously. Further, the medium IR band at 1506 cm1 and effect induction leading to the enhancement of stretching
weak Raman band at 1510 cm1 correspond to the NHþ 3 wavenumbers and decrease of infrared intensities.
symmetric deformation mode. Generally, in the amino
acids, multiple combination bands of weak intensity occur 6.1.2.1. Improper, blue-shifting hydrogen bonding. Hydro-
in the 2222–2000 cm1 region, the most prominent being gen bonding (H-bond) originates from an attractive inter-
the band near 2000 cm1. In GSN, an additional band is action between the electron-deficient hydrogen donor
identified at 2009 cm1 can be unambiguously correlated group (AAH) and a region of high electron density accep-
to the combination of the torsional oscillation and degen- tor atom (B), leading to the variation of H  B distance
erate deformation of the NHþ 3 groups, which can be used than the van der Waals radii of the isolated H and B atoms.
a ‘finger print’ for the identification of NHþ 3 group in The attractive interaction between the hydrogen donor
GSN crystal. This band is the characteristic marker of group and the acceptor atom modifies the molecular poten-
the NHþ 3 groups present in the organic compounds partic- tial energy surface and has dramatic consequences for the
ularly in the amino acids, as this mode disappears in the vibrational spectra. Most frequently, an H-bond is of the
spectrum of the deuterated derivatives [53]. The rocking XAH  Y type, where X and Y are electronegative ele-
and torsion vibrations of NHþ 3 group has been identified ments and Y possesses one or two electron lone pairs.
and assigned with the aid of DFT calculations. Although the interaction energy of a CAH  O hydrogen
bond is less than those of typical NAH  O and OAH  O
6.1.2. Methylene vibrations type bonds, the CAH type hydrogen bond plays an impor-
The internal vibrational modes of the CH2 group are tant role in determining higher order structure in proteins,
well separated from the vibrations of the other part of molecular structure and conformation and crystal packing.
the glycine molecule. This is particularly expected for the In CAH  O type hydrogen bond where the proton donor
stretching CH2 vibrations. Absorption arising from CAH is sp3-hybridized, the hydrogen bonded CAH undergoes
stretching in the alkanes occurs in the region 3000– contraction due to interaction with a proton acceptor.
2840 cm1. The qualitative interpretation of intensities The shortening of CAH bond contrasts sharply with the
must rely upon the understanding of some basic aspects elongation of OAH and NAH in OAH  O and NAH  O
of intramolecular charge distribution and on their effects type hydrogen bonds. It has been concluded that the differ-
on infrared intensities [54]. The molecules that show well- ences in charge transfer exist between CAH  O type and
recognized effects of induction from electronegative atoms typical OAH  O, NAH  O type hydrogen bonds [60,61]
in the surrounding or of backdonation of negative charge and also the mechanism of OAH  O and NAH  O type
from lone pairs in the same molecules or of hyperconjuga- hydrogen bond formations is a direct process where the pri-
tion with systems of an aromatic rings or multiple bonds mary effect is the charge transfer from the proton acceptor
can give rise to specific signals in the spectrum, especially to the OAH and NAH anti-bonding orbital of the proton
in intensity. These signals are so strong that they can lead donor, and thus an increase in electron density in this orbi-
to a quantitative diagnosis of charge distribution directly tal leads to weakening of the OAH and NAH bonds,
from a careful analysis of the vibrational spectrum. Differ- accompanied by their elongation. In CAH  O hydrogen
ent theoretical methods have been proposed for infrared bond, alternatively, a charge transfer from the lone pairs
intensities, which can reveal a depiction of the charge dis- of the electron donor is directed mainly to the anti-bonding
tribution and charge mobility in molecules [55,56]. The orbitals in the remote part of the complex, thus causing
parameters charge (q0H ) and charge flux (oqH/orCH) are very elongation in that part of a complex. This primary effect
important markers of the charge distribution inside the of elongation is accompanied by a secondary effect of struc-
molecule and are directly related to the molecular struc- tural reorganization of the proton donor, leading to con-
ture, the vibrational potential and the molecular confirma- traction of the CAH bond. Therefore, a CAH  O
tion [57]. Particularly, a large charge in general implies a hydrogen bond is considered an ‘improper hydrogen bond
strong bond, with a short interatomic distance r0CH and a or blue-shifting H-bond’ [62]. Blue- shifting hydrogen
large stretching force constant kCH. It is important to be bonds are characterized by a contraction of the CAH dis-
aware that the infrared intensity in the CAH stretching tance, a blue shift of the CAH stretching vibrational mode,
region is a genuine marker of the charge distribution and a reduction of its infrared intensity, features which are
30 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

in sharp contrast to those rooted to the conventional culated at 1623 cm1) and the COO symmetric stretching
hydrogen bonds [63]. mode observed at 1414 cm1 (calculated at 1433 cm1) in
The asymmetric CH2 stretching (mas CH2) mode is the Raman spectrum. The lowering of this mode from
expected in the region near 2926 cm1 and the symmetric the computed values can be due to the interaction of the
mode (ms CH2) around 2853 cm1 [51]. But, for amino lone pair oxygen atoms with the nucleophilic atoms of
acids, mas CH2 vibrations generally appear in the region GSN molecules through the strong intra- and intermolecu-
3100–3000 cm1 while ms CH2 vibrations appear around lar NAH  O and CAH  O hydrogen bonding. Further-
3000–2900 cm1 [33]. The CH2 asymmetric stretching more, as discussed in the preceding section optimized
appears at 3026 cm1 in IR and in the Raman spectrum geometry, the considerable lowering of both asymmetric
at 3023 cm1 as weak bands. The corresponding stretching and symmetric carboxylate wavenumbers by 44 and
mode is computed at 3016 cm1 by the DFT calculations, 19 cm1, respectively, in GSN is mainly due to the intermo-
which explains that the observed vibrational wavenumbers lecular non-bonded interaction of metal atom Na with
are larger by around 10 cm1 than the computed value eight neighbouring oxygen atoms of the carboxylate and
from their normal coordinates, as shown in Fig. 8. The nitrate groups. The COO deformation, wagging and rock-
symmetric CH2 stretching mode band is observed at ing modes have been identified and assigned, which con-
2959 cm1 as weak IR band and in the Raman spectrum firms that the glycine molecule in GSN exists in
at 2969 cm1 as medium band, which is calculated at zwitterionic form with deprotonated carbonyl groups and
2947 cm1. As observed in the asymmetric stretching protonated amino groups.
mode, the calculated symmetric stretching vibration is low-
ered from the observed wavenumbers by around 20 cm1. 6.2. Sodium nitrate vibrations
The bond lengths of C2AH9 and C2AH8 bonds are calcu-
lated to be 1.09 and 1.10 Å from the distance matrix of The nitrate anion in sodium nitrate is a potential acceptor
the optimized geometry of GSN while the experimentally of six hydrogen atoms and thus six NAH  O hydrogen
observed bond lengths for the corresponding bonds are bonds may be formed. The crystal structure [40] shows that
found to be 0.97 and 0.97 Å, respectively. The experimental there are four such hydrogen bonds in a two dimensional
CAH bond lengths are shortened about 0.12 and 0.13 Å network. As such, one O atom is an acceptor of two protons
over the calculated values, which can be attributed to the and the other two O atoms accept a proton each. The nitro-
increase in wavenumbers of the stretching modes of the gen/oxygen stretch of inorganic nitrate appears as an intense
CH2 group in GSN. Shifting of the stretching frequencies band between 1400 and 1340 cm1. Planar XY3 molecules
towards higher wavenumbers, intensity variation and the have D3h symmetry and their normal vibrations m1 mode
CAH bonds contraction indicating the existence of (symmetric stretching), m2 mode (out-of-plane deformation),
CAH  O hydrogen bonding throughout the GSN crystal. m3 mode (asymmetric stretching) and m4 mode (out-of-plane
This fact is further substantiated by XRD analysis of GSN deformation) are concurrently active, except one mode, in
crystal that explains the CAH  O bond distance and the IR and Raman spectra. The inorganic nitrate anion, NO 3,
corresponding bond angle between molecules of GSN are is planar and their m1 mode, m sym NO 3 , is IR inactive and
to be 3.2 Å and 131, respectively. the m2 mode, cNO 3 , is Raman inactive. Thus, the IR active
The bending vibrations of the CAH bonds in the meth- modes are m2 through m4 and all but m2 are Raman active.
ylene group are identified in their respective positions. The Generally, the m1 mode appears in the region 1034–
scissoring mode of the CH2 group gives rise to a character- 1067 cm1 while the m2 mode is expected in the region 801–
istic band near 1465 cm1 in IR and Raman spectra. This 834 cm1. Moreover, the m3 (m asym NO 
3 ), and m4 (dNO3 ),
mode appears as intense Raman band at 1449 cm1, which modes are doubly degenerate, which usually occur in the
is calculated at 1421 cm1. The twisting, wagging and rock- region 1300–1550 cm1 and 700–751 cm1, respectively [65].
ing vibrations appear in the region of 1422–719 cm1. The The most intense Raman band of the GSN crystal
frequencies corresponding to IR bands at 1307 cm1 (very observed at 1052 cm1 can be correlated to the symmetric
strong), 1137 cm1 (medium) and 935 cm1 (very strong) stretching m1 mode of the nitrate anion and the nitrate
are correlated to the CH2 wagging, twisting and rocking asymmetric stretching m3 mode appears at 1353 cm1 as
modes, respectively, which are supported by computational most intense doublet IR bands, which is overlapping with
wave numbers. the wagging mode of the methylene group. The NO3 asym-
metric and symmetric stretching vibrations are calculated
6.1.3. Carboxylate vibrations at 1389 and 1072 cm1, respectively, and observed frequen-
The carboxylate ion gives rise to two modes; asymmetric cies are lowered from the computed values by around
stretching near 1650–1550 cm1 and symmetric stretching 25 cm1 which reveals the non-bonded interactions of oxy-
mode near 1400 cm1 while the C@O stretching wave num- gen atoms of the nitrate groups with neighbouring metal
bers of the un-ionized carboxylic group of the glycine mol- atom Na and amino groups of the glycine molecules. Fur-
ecule is usually found near 1740 cm1 as an intense IR thermore, these modes are broader and stronger than those
band [51,64]. The asymmetric stretching mode of COO of all other modes in the GSN crystal suggesting that being
vibration appears in IR as intense band at 1579 cm1 (cal- a strong acceptor these nitrate groups are actively partici-
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 31

Fig. 8. Atomic displacements of selected modes.

pated in both intra- and intermolecular NAH  O ionic 1768 cm1 in IR spectrum, as expected, as an additional
hydrogen bonding with the other electron-deficient atoms. weak band, which is not computed in the DFT calculations.
The m2 and m4 modes of the inorganic nitrate anion vibra-
tions have been identified and assigned. 6.3. Skeletal modes
Inorganic nitrates exhibit one or two weak IR bands in
the region 1734–1790 cm1, which are correlated to the com- The absorption bands arising from CAN and CAC
bination tone m1 + m4. This combination band appears at stretching vibrations are generally measured in the region
32 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

Fig. 8 (continued)

1150–850 cm1 [49,66]. The C2AC3 stretching mode appears in Raman spectrum at 178 cm1 as medium
appears as medium IR band at 895 cm1 and in Raman band.
spectrum at 890 cm1 (medium). Likewise, the C2AN1
stretching mode observed in IR at 1039 cm1 (medium) 6.4. Low wavenumber hydrogen bond vibrations
and at 1039 cm1 in Raman spectrum. The very weak
Raman band at 331 cm1 is attributed to the C3AC2AN1 The attractive interaction between the hydrogen donor
deformation mode. The translation mode of the Na+ ion group and the acceptor moiety leads to the occurrence of
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 33

new vibrational degrees of freedom, the so-called hydrogen tions, highly polarizable cations, responsible for NLO
bond modes. Such modes are connected with elongations properties, are linked to anions through hydrogen bond
changing the A  B distance and/or the relative orientation networks, which generate a non-centrosymmetric struc-
of the hydrogen bonded groups. Thus, they provide direct tural organization. In the case of molecules with lower
insight into the structure of hydrogen bonds and into pro- symmetry, it is interesting to note that an alternated stack
cesses of bond formation and cleavage. As such modes are would necessarily result in a non-centrosymmetric chain.
characterized by a high reduced mass of the oscillator and a Recently the method derived from the dielectric theory of
small force constant determined by the comparably weak complete crystals and the Levine bond charge model to
attractive interaction along the hydrogen bond, hydrogen the hydrogen bonded solids has been successfully applied
bond modes occur at low frequencies in the range between and the results obtained show that the hydrogen bond con-
about 50 and 300 cm1 [67]. An interesting feature of these tributes to the second order NLO tensor coefficient (dijk) of
vibrations is the occurrence of the intense Raman band the crystals [73]. Certainly hydrogen bonds create and sta-
observed at 109 cm1 is correlated to the N  O stretching bilize the crystal structures but more evidently that they
H-bonds vibrations [7,68]. Such Raman bands are also contribute considerably to the enhancement of hyper-
observed in the amino acid based NLO active crystals that polarizability of hydrogen bonded molecular systems [74]
stabilize the crystal structure through the strong hydrogen or to the enhancement of the second order susceptibility
bonded networks. However, these bands appear in certain of the crystals [75].
NLO inactive crystals like glycine lithium nitrate (GLiN) in The low wavenumber wave packet motions have been
which two Raman bands appear at 132 and 104 cm1 with recently observed along moderately strong intramolecular
the same intensity as the Raman band 109 cm1 of GSN hydrogen bonds with well-defined geometries by the
crystal [69]. It is interesting to note that, in GLiN, the high pump-probe spectroscopy [67]. It would be useful to men-
wavenumber band 132 cm1 disappears both on deutera- tion that with the extent of charge transfer is expected to
tion or when the lithium atom is replaced by a sodium provide stronger H-bonding interaction between the accep-
atom. These spectral features in such crystals are significant tor group and the H atom of the donor group [76]. The cal-
area of research to be explored. The analysis of the eigen- culated the first hyperpolarizability (btot) and the ground
vectors reveals that the vibration corresponding to the state dipole moment (lg) is 14.5487 · 1031 esu and 6.18
119 cm1 band is associated with the in-phase vibrations Debye, respectively. Table 9 shows that btot has the largest
of atoms of the molecule connected through the ionic calculated value for GSN crystal and the zwitterionic gly-
NAH  O hydrogen bonding. The twisted intramolecular cine molecule while btot is considerably decreased for the
charge transfer from the donor to the acceptor through constituents of the GSN crystal. As can be seen that the
the H-bond results in electron–phonon coupling in this compounds having the higher dipole moment results in
hydrogen bonded system, which provokes the band to be the higher btot value and the corresponding HOMO–
very intense in the Raman spectrum. This reveals the vibra- LUMO energy gap is quite low. This clearly indicates that
tional contribution to the hyperpolarizability of such non- the strong hydrogen bonding between the charged species
linear optical crystals [12,70] that provide the non- reduces the energy gap considerably with the formation
centrosymmetric structure contributing to render GSN of the charge transfer axis [77]. In acid–base hybrid crystals
crystal NLO active. hydrogen bonds play an important role not only in the cre-
ation of crystal structure and its stability, but also in the
6.5. Effect of ionic hydrogen bonds on NLO properties enhancement of second order susceptibility of the crystal
due to the perturbation of the electronic structure of the
It is generally recognized that the first electronic hyper- organic partner and also due to the strong electron–pho-
polarizability value (b) is enhanced when there is a low- non coupling [72,78].
lying ground-to-excited state transition with large change
in the dipole moment [71,72]. However, much less progress 7. Conclusions
has been made in understanding the microscopic optical
nonlinearity in inorganic materials. Arising from the com- Single crystals of GSN grown by slow evaporation tech-
plexation of organic molecules based on acid–base interac- nique and the second harmonic generation efficiency was

Table 9
Comparison of static first hyperpolarizability, dipole moment, HOMO–LUMO energy gap and SCF energy for the constituents of GSN
Compounds Dipole moment lg (Debye) Hyperpolarizability btot (·1031 esu) Energy gap a.u. SCF energy a.u.
Glycine (Molecular) 6.2602 0.1536 0.6038 282.825
Glycine (Zwitterion) 10.945 12.537 0.5243 282.786
Glycine Sodium Nitrate 6.1877 14.543 0.2216 727.293
Sodium nitrate 8.7763 0.2365 0.3943 440.765
34 T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35

measured by Kurtz–Perry powder SHG experiments, University of Hyderabad for their support in NLO mea-
which is around 0.3 times that of urea. The calculated first surements. The assistance from BRUKER OPTICS in
hyperpolarizability of GSN is found to be 14.54 · 1031 esu, recording the NIR FT-Raman and FT-IR spectra is grate-
which is 7 times that of urea. The equilibrium geometry fully acknowledged.
optimizations of GSN were carried out using DFT, MP2
and ab initio computations at different special basis set References
including LANL2DZ and analyzed. In a-glycine, c-glycine
and GSN, the experimentally measured bond lengths are [1] P.N. Prasad, D.J. Williams, Introduction to Nonlinear Optical Effects
rather lower than the data measured by the different theo- in Organic Molecules and Polymers, John-Wiley & Sons Inc., New
York, 1991.
retical methods including the special basis set LANL2DZ. [2] H.S. Nalwa, S. Miyata, Nonlinear Optics of Organic Molecules and
In GSN, XRD shows the same angle is 111.9 and it is Polymers, CRC Press, Boca Raton, 1997.
interesting to note that the experimental and theoretical [3] D.S. Chemla, J. Zyss, Nonlinear Optical Properties of Organic
bond lengths and bond angles of the glycine molecules in Molecules and Crystals, Academic Press, New York, 1987.
[4] Ch. Bosshard, K. Sutter, P.h. Pretre, J. Hulliger, M. Florsheimer, P.
GSN are quite comparable to that of the a-glycine than
Kaatz, P. Gunter, Organic nonlinear optical materials, Advances in
c-glycine. However, the twist of the glycine skeleton of Nonlinear Optics, vol. 1, Gordon and Breach, Amsterdam, 1995.
a-form, which is pre-requisite to make the glycine molecule [5] Ying Li, Zhi-Ru Li, Di Wu, Rui-Yan Li, Xi-Yun Hao, C.C. Sun, J.
NLO active, may be introduced by the Twisted Intramolec- Phys. Chem. B 108 (2004) 3145.
ular Charge Transfer (TICT) due to the presence of strong [6] M.N. Bhat, S.M. Dharma prakash, J. Cryst. Growth 236 (2002) 376.
ionic intra- and intermolecular N+AH  O hydrogen [7] L. Padmaja, T. Vijayakumar, I. Hubert Joe, C.P. Reghunadhan Nair,
V.S. Jayakumar, J. Raman Spectrosc. 37 (2006) 1427.
bonding. [8] M.N. Bhat, S.M. Dharma prakash, J. Cryst. Growth 235 (2002) 511.
The presence of strong NAH  O intra- and intermolec- [9] J. Zyss, J.F. Nicoud, M. Coquillay, J. Chem. Phys. 81 (1984) 4160.
ular hydrogen bonding is also evident from the lowering of [10] C. Razzetti, M. Ardoino, L. Zaotti, M. Zha, C. Parorici, Cryst. Res.
1
the NHþ 3 symmetric stretching frequency to 2886 cm in Technol. 37 (2002) 456.
IR spectrum, which is further substantiated by the NBO [11] H. Ratajczak, J. Baran, J. Barycki, S. Debrus, M. May, A. Pietraszko,
H.M. Ratajczak, A. Tramer, J. Venturini, J. Mol. Struct. 555 (2000)
analysis and computations. The experimental CAH bond 149.
lengths are shortened about 0.12 and 0.13 Å over the calcu- [12] S. Debrus, H. Ratajczak, J. Venturini, N. Pincon, J. Baran, J.
lated values, which can be attributed to the increase in Barycki, T. Glowiak, A. Pietraszko, Syn. Metals 127 (2002) 99.
wavenumbers of the stretching modes of the CH2 group [13] G.R. Desiraju, J. Mol. Struct. 656 (2003) 5.
[14] D. Eimert, S. Velsko, L. Davis, F. Wang, G. Loiaceono, G. Kennedy,
in GSN. Shifting of the stretching frequencies towards
IEEE J. Quantum Electron 25 (1989) 179.
higher wavenumbers, intensity variation and the CAH [15] M.D. Aggarwal, J. Choi, W.S. Wang, K. Bhat, R.B. Lal, A.D. Shield,
bonds contraction indicating the existence of ‘blue-shift B.G. Penn, D.O. Frazier, J. Cryst. Growth 204 (1999) 179.
or improper’ CAH  O hydrogen bonding throughout [16] R.D.A. Hudson, A.R. Manning, J.F. Gallagher, M.H. Garcia,
the GSN crystal. The occurrence of the intense Raman Tetrahedron Lett. 43 (2002) 8375.
band observed at 109 cm1 is correlated to the N  O [17] B. Brezina, Mat. Res. Bull. 6 (1971) 401.
[18] J.K.M. Rao, M.A. Vishwamitra, Acta Crystallogr. B 28 (1972) 1484.
stretching H-bonds vibrations. This vibration favours the [19] S. Natarajan, J.K.M. Rao, Z. Kristallogr. 152 (1984) 179.
twisted intramolecular charge transfer from the donor to [20] G. Albrecht, R.B. Corey, J. Am. Chem. Soc. 61 (1939) 1087.
the acceptor and carries out the phenomenon of the elec- [21] E. Fischer, Ber. Dtsch. Chem. Ges. 38 (1905) 2917.
tron–phonon coupling in this hydrogen bonded material [22] Y. Iitaka, Acta Crystallogr. 11 (1958) 225.
what provokes to be very intense in the Raman spectrum. [23] Y. Iitaka, Acta Crystallogr. 14 (1961) 1.
[24] I. Weissbuch, V. Yu. Torbeev, L. Leiserowitz, M. Lahav, Angew.
These vibrations provide the non-centrosymmetry struc- Chem. Int. Ed. 44 (2005) 3226.
ture may be contributing to make the GSN crystal NLO [25] S. Debrusa, M. Maya, J. Baryckib, T. Glowiakc, A.J. Barnesd, H.
active, which may be considered as a diagnostics tool for Ratajczakc, D. Xuef, J. Mol. Struct. 661-662 (2003) 595.
identifying the crystals to be NLO active. In acid–base [26] A.A. Sukhorukov, S. Kivshar Yu, J. Opt. Soc. Am. B 19 (2002)
hybrid crystals such ionic hydrogen bonds play an impor- 772.
[27] G. Maroulis, J. Chem. Phys. 113 (2000) 1813.
tant role not only in the creation of crystal structure and [28] S.F. Mingaleev, S. Kivshar Yu, Opt. Photon. News 13 (2002) 48.
its stability, but also in the enhancement of second order [29] B.A.S. Mendis, K.M.N. de Silva, J. Mol. Struct. (Theochem) 678
susceptibility of the crystal due to the perturbation of the (2004) 31.
electronic structure of the organic partner and also due [30] S.P. Liyanage, R.M. de Silva, K.M.N. de Silva, J. Mol. Struct.
to the strong electron–phonon coupling. (Theochem) 639 (2003) 195.
[31] M. Montejo, A. Navarro, G.J. Kealey, J. Vazquez, J.J.L. Gonzalez, J.
Am. Chem. Soc. 126 (2004) 15087.
Acknowledgements [32] J. Binoy, Jose P. Abraham, I. Hubert Joe, V.S. Jayakumar, GR.
Pettit, OF. Nielsen, J. Raman Spectrosc. 35 (2004) 939.
The authors are grateful to the Department of Space, [33] D. Sajan, J. Binoy, B. Pradeep, K. Venkata Krishna, V.B. Kartha, I.
Hubert Joe, V.S. Jayakumar, Spectrochim. Acta A 60 (2004) 173.
Government of India, and Vikram Sarabhai Space Centre,
[34] D. Sajan, K. P Laladhas, I. Hubert Joe, V.S. Jayakumar, J. Raman
Trivandrum, for financial support through the RESPOND Spectrosc. 36 (2005) 1001.
project. The authors thank Professor T.P.Radhakrishnan [35] Jose P. Abraham, I. Hubert Joe, V. George, O.F. Nielsen, V.S.
and Dr. Philip Anthony of the School of Chemistry, Jayakumar, Spectrochim. Acta A 59 (2003) 193.
T. Vijayakumar et al. / Journal of Molecular Structure 877 (2008) 20–35 35

[36] J. Binoy, Jose P. Abraham, I. Hubert Joe, V.S. Jayakumar, J. [53] G. Dhanaraj, M.R. Srinivasan, H.L. Bhat, J. Raman Spectrosc. 22
Aubard, O.F. Nielsen, J. Raman Spectrosc. 36 (2005) 63. (1991) 177.
[37] D. Sajan, J. Binoy, I. Hubert Joe, V.S. Jayakumar, J. Zaleski, J. [54] M. Gussoni, C. Castiglioni, J. Mol. Struct. 521 (2000) 1.
Raman Spectrosc. 36 (2005) 221. [55] W.T. King, G.B. Mast, P.B. Blanchette, J. Chem. Phys. 53 (1972)
[38] J. Binoy, I. Hubert Joe, V.S. Jayakumar, J. Raman Spectrosc. 36 4440.
(2005) 1091. [56] R.E. Bruns, R.E. Brown, J. Chem. Phys. 68 (1978) 880;
[39] D. Sajan, I. Hubert Joe, V.S. Jayakumar, J. Raman Spectrosc. 37 M. Gussoni, J. Mol. Struct. 113 (1984) 323.
(2006) 508. [57] M. Gussoni, C. Castiglioni, G. Zerbi, J. Phys. Chem. 88 (1984)
[40] R.V. Krishnakumar, M. Subha Nandhini, S. Natarajan, K. Sivaku- 600.
mar, Babu Varghese, Acta Crystallogr. C57 (2001) 1149. [58] M. Gussoni, J. Mol. Struct. 141 (1986) 63.
[41] Y. Iitaka, Proc. Jpn. Acad. 30 (1954) 109. [59] C. Castiglioni, M. Gussoni, G. Zerbi, J. Mol. Struct. 141 (1986) 341.
[42] S.K. Kurtz, T.T. Perry, J. Appl. Phys. 39 (1968) 3798. [60] P. Hobza, Chem. Rev. 100 (2000) 4253.
[43] J.B. Foresman, A. Frisch, Exploring Chemistry with Electronic [61] A. Kovacs, A. Szabo, D. Nemcsok, I. Hargittai, J. Phys. Chem. A 104
Structure Methods, second ed., Gaussian Inc., Pittsburgh, USA, 1996. (2000) 6286.
[44] A.P. Scott, L. Radom. J. Phys. Chem. 100 (1996) 16503. [62] P. Hobza, J. Sponer, W. Cubero, M. Orozco, F.J. Luque, J. Phys.
[45] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, Chem. A 106 (2002) 5671.
J.R. Cheeseman, V.G. Zakrzewski, J.A. Montgomery, R.E. Strat- [63] P. Hobza, Phys. Chem. Chem. Phys. 7 (2005) 3027.
mann, J.C. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N. [64] B. Smith, Infrared Spectral Interpretation, a Systematic Approach,
Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. CRC press, Washington, DC, 1999.
Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, [65] R.A. Nyquist, C.L. Putzig, M.A. Leugers, Infrared and Raman
G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma, D.K. Malick, Spectral Atlas of Inorganic Compounds and Organic Salts, Academic
A.D. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. press, New York, 1995.
Ortiz, A.G. Baboul, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, [66] L.J. Bellamy, The Infra-red Spectra of Complex Molecules, vols. 1–2,
I. Komaromi, R. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A Chapman and Hall, London, 1975.
Al-Laham, C.Y Peng, A. Nanayakkara, M. Challacombe, P.M.W. [67] E.T.J. Nibbering, T. Elsaesser, Chem. Rev. 104 (2004) 1887.
Gill, B. Johnson, W. Chen, M.W. Wong, J.L. Andres, C. Gonzalez, [68] M. Meot-Ner (Mautner), Chem. Rev. 105 (2005) 213.
M. Head-Gordon, E.S. Replogle, J.A. Pople, Gaussian 98, revision [69] J. Baran, M. Drozd, A. Pietraszko, M. Trzebiatowska, H. Ratajczak,
A.9, Pittsburgh, PA, 1998. Polish J. Chem. 77 (2003) 1561.
[46] D.A. Kleinman, Phys. Rev. 126 (1962) 1977. [70] D.M. Bishop, B. Kirtman, B. Champagne, Phys. Rev. B 56 (1997)
[47] G.A. Jefrrey, Accurate crystal structure analysis by neutron diffrac- 2273.
tion in accurate molecular structures, in: A. Domenicano, I. Hargittai [71] W.H. Thompson, M. Blanchard-Desce, A.J. Muller, A. Fort, M.
(Eds.), Oxford University Press, Oxford, 1992, pp. 270–298. Barzoukas, J.T. Hynes, J. Phys. Chem. A 103 (1999) 3766 (and
[48] A.E. Reed, L.A. Curtiss, F. Weinhold, Chem. Rev. 88 (1988) 899. references therein).
[49] V. Alabugin, M. Manoharan, S. Peabody, F. Weinhold, J. Am. [72] Y. Shen, The Principles of Nonlinear Optics, Wiley, New York, 1984.
Chem. Soc. 125 (2003) 5973. [73] B.F. Levine, Phys. Rev. B. 7 (1973) 2600.
[50] D.L. Vein, N.B. Colthup, W.G. Fateley, J.G. Grasselli, The Hand- [74] F.L. Huyskens, P.L. Huyskens, A.P. Persoon, J. Chem. Phys. 108
book of Infrared and Raman Characteristic Frequencies of Organic (1998) 8161.
Molecules, Academic Press, New York, 1991. [75] D. Xue, S. Zhang, J. Phys. Chem. A 101 (1997) 5547.
[51] R.M. Silverstein, F.X. Webster, Spectroscopic Identification of [76] P.K. Nandi, K. Mandal, T. Kar, Chem. Phys. Lett. 381 (2003) 230.
Organic Compounds, sixth ed., John Wiley & Sons Inc., New York, [77] Ch. Bosshard, R. Spreiter, L. Degiorgi, P. Gunter, Phys. Rev. B 66
2003. (2002) 205107.
[52] Z.R. Grabowski, K. Rotkiewicz, W. Rettig, Chem. Rev. 103 (2003) [78] D.M. Bishop, B. Kirtman, B. Champagne, J. Phys. Chem. A 101
3907. (1997) 5780.

S-ar putea să vă placă și