You are on page 1of 8

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

A combined experimental and theoretical analysis on molecular


structure and vibrational spectra of 2,4-dihydroxybenzoic acid
Yaping Tao, Ligang Han, Yunxia Han, Zhaojun Liu ⇑
College of Physics and Electronic Information, Luoyang Normal University, Luoyang 471022, China

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 FT-IR and FT-Raman spectra of 2,4-


dihydroxybenzoic acid have been
recorded and analyzed.
 Different conformers with their
relative energies of 2,4-
dihydroxybenzoic acid have been
calculated.
 The complete vibrational assignments
of were made by combining
experimental and theoretical data
using PED analysis.
 HOMO, LUMO energies and MEP
distribution of 2,4-dihydroxybenzoic
acid were performed.

a r t i c l e i n f o a b s t r a c t

Article history: The FT-IR and FT-Raman spectra of 2,4-dihydroxybenzoic acid (2,4-DHBA) in the solid phase have been
Received 3 May 2014 recorded in the regions 4000–400 cm1 and 3700–100 cm1, respectively. The total energies of sixteen
Received in revised form 24 July 2014 stable conformers for 2,4-DHBA have been calculated by density functional theory (DFT) using the
Accepted 31 August 2014
B3LYP method with 6-311++G (d, p) basis set, and the C1 conformer with the lowest energy was obtained,
the geometrical parameters between X-ray experiment diffraction and DFT calculation show good consis-
tency. Furthermore, the vibrational frequencies of 2,4-DHBA were computed, and the detailed analysis of
Keywords:
vibrational spectra was made on the basis of the potential energy distribution (PED) by combining exper-
2,4-Dihydroxybenzoic acid
DFT
imental with theoretical data. In addition, frontier molecular orbitals, atomic charge distribution and
FT-IR molecular electrostatic potential (MEP) were also given.
FT-Raman Ó 2014 Elsevier B.V. All rights reserved.
Vibrational analysis

Introduction organic compounds such as resins, polyesters, plasticizers, dyestuff


and rubber chemicals. As an intermediate, 2,4-DHBA plays an
2,4-Dihydroxybenzoic acid, also known as b-resorcylic acid, is important role in the production of cosmetics and dyes [5,6]. It
one of the hydroxybenzoic acid derivatives. Hydroxybenzoic acids can be also used as antimicrobial wash for reducing escherichia
are commonly used in pharmaceutical and perfumery industry coli O157:H7 on apples. In addition, it is very efficient to reduce
because of their activities in the biological system [1]. Dihydroxy- oxidative damage [7] associated with various diseases such as
benzoic acids are the important materials in the production of cancer, cataracts, arthritis, and diabetes. Meanwhile, as a precursor
disinfectants and antipyretic drugs [2–4], and in the synthesis of to the quinones which has an deleterious effect on environment, it
can often cause water pollution [8].
So far, several researches have investigated the structure of 2,4-
⇑ Corresponding author. Tel.: +86 379 65515016. DHBA. In 2007, Parkin et al. [9] made an effort on its crystal struc-
E-mail address: zhaojunliu@gmail.com (Z. Liu). ture and hydrogen bonding by multiple temperature single-crystal

http://dx.doi.org/10.1016/j.saa.2014.08.151
1386-1425/Ó 2014 Elsevier B.V. All rights reserved.
Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085 1079

Fig. 1. Conformers of 2,4-DHBA.

X-ray diffraction study. Recently, Braun and co-workers [10] stud- calculation was used to gain a better understanding on the vibra-
ied its crystallization and compared with 2,5-dihydroxybenzoic tional spectra of 2,4-DHBA. Firstly, the total energy of various con-
acid. To the best of our knowledge, no detailed study on vibrational formers of 2,4-DHBA was calculated, and the most stable
spectra for 2,4-DHBA has been reported up to now. Literature sur- conformer was found at the B3LYP/6-311++G (d, p) level. Secondly,
vey shows that the vibrational assignments for the similar mole- the complete molecular geometry description and the normal
cule like 2,3-dihydroxybenzoic acid [11] and 4-dihydroxybenzoic coordinate analysis were given using the method suggested by
acid [12] have been made. However, the position of hydroxyl Pulay et al. [17–19]. Also, the HOMO–LUMO analysis, atomic
groups and the hydrogen bonds strongly affect the vibrational charge distribution and molecular electrostatic potential (MEP)
spectra, the highly accurate calculations on the structure of 2,4- were studied theoretically on the same basis set.
DHBA and the detailed assignment on its vibrational bands are still
desired.
Density functional theory (DFT) is a reliable and accurate Experimental details
method for predicting molecular structure, vibrational frequencies,
Raman activities, infrared intensities and charge distribution of the The fine solid sample of 2,4-DHBA was purchased from Beijing
organic molecule [13–16]. In this study, quantum chemical Chemical Factory, and used without further purification. The FT-IR
1080 Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085

spectrum was collected from 4000 to 400 cm1 on a Nicolet Nexus


FT-IR spectrometer. Nujol mull method was used in the sample
preparation. The FT-Raman spectrum was recorded in the range
of 3700–100 cm1 on a Raman module of the Nexus spectrometer
with 1064 nm excitation from an Nd:YAG laser. The laser power on
the sample was about 400 mw. Both IR and Raman spectra were
accumulated with a 2 cm1 resolution at room temperature.

Computational details

Geometry optimization, energy and vibrational wavenumber


computation were performed using Gaussian 09 w [20] program
for 2,4-DHBA at B3LYP [21,22] level with the standard 6-311++G
(d, p) basis set. B3LYP is a popular and relatively efficient density
functional method for the determination of energies and geome-
tries and has been proven to be useful to investigate interactions
and conformational preferences for the various molecules [23–
25]. A comparison between the calculated and the measured fre-
quencies usually shows that the calculated data are slightly greater
than the measured frequencies. To improve the agreement, the
computed harmonic frequencies are usually scaled for comparison.
Many studies [26,27] show that the scaled quantum mechanical
Fig. 2. Molecular model of 2,4-DHBA along with numbering of atoms.
(SQM) method provides calculated frequencies well matched to
the experimental. In this paper, the scaling calculation and poten-
tial energy distribution (PED) for each normal mode were done
with the MOLVIB program (Version 7.0-G77) by Sundius [28,29].
frequency, and shown in Fig. 1. It was clear from Table S1 that
Raman activities (Si) obtained from Gaussian 09 output were
the C1 conformer of 2,4-DHBA has produced the global energy
converted to relative Raman intensity (Ii) using the following
minimum. The total energies and energy differences with respect
relation from the basic theory of Raman scattering [30,31]:
to C1 conformer were also listed in Table S1. The discussion below
f ðm0  mi Þ4 Si refers only to this C1 conformer. The molecular models along with
Ii ¼ numbering of atoms on the monomer and dimer were shown in
mi ½1  exp ðhcmi =kT Þ
Figs. 2 and 3, respectively. The total energies of the monomer
In the above formula, m0 is the laser excitation wavenumber (in and dimer were about 1500007.06 and 3000071.92 kJ/mol,
cm1), mi is the vibrational wavenumber (in cm1) of the ith normal respectively. The molecules in dimer are bound together via doubly
mode; c, h, k and T are the speed of light, Planck’s constant, hydrogen bonded, so the energy of dimer was not twice of its
Boltzmann’s constant and temperature (in K), respectively. f is a monomer structure owing to the effect of hydrogen bonds. The
suitably chosen scale factor for all the band intensities. Simulation interaction energy of the formation of the intermolecular hydrogen
of calculated IR and Raman spectra have been plotted using pure bond dimer (DE = Edimer  2  Emonomer) [32] was 57.7980619 kJ/
Lorentzian band shapes with a bandwidth (FWHM) of 10 cm1. mol. This is very close to the energy of hydrogen bond.
Also, the geometrical parameters for monomer and dimmer
Results and discussion were depicted in Table 1. For the purpose of comparison, the cor-
responding experimental parameters obtained from X-ray diffrac-
Molecular structure description tion were provided. Root mean square deviation (RMSD), as
showed in Table 1, has been calculated separately for bond lengths
In order to obtain the most stable geometry of 2,4-DHBA, the and bond angles. From the structural data given in Table 1, we can
total energy calculations were carried out using B3LYP/6-311++G find that most of the optimized bond lengths of dimer were
(d, p) level for various possible conformers without imaginary matched to the experimental values than monomer values, the

Fig. 3. Dimer of 2,4-DHBA.


Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085 1081

Table 1
The selected experimental and theoretical geometric parameters for 2,4-DHBA.

Bond length (Å) Bond angle (°) Dihedral angle (°)


a a
Parameters Cal. Exp. Parameters Cal. Exp. Parameters Cal.
Monomer Dimer Monomer Dimer Monomer Dimer
C6AC1 1.409 1.407 1.413 C5AC6AC1 118.70 118.58 118.43 C4AC5AC6AC1 0.00 0.00
C1AC2 1.378 1.379 1.382 C6AC1AC2 121.61 121.49 121.44 C5AC6AC1AC2 0.00 0.00
C2AC3 1.406 1.407 1.411 C1AC2AC3 118.91 119.06 119.10 C6AC1AC2AC3 0.00 0.00
C3AC4 1.389 1.388 1.397 C2AC3AC4 121.04 121.00 120.91 C1AC2AC3AC4 0.00 0.01
C4AC5 1.397 1.395 1.396 C3AC4AC5 120.03 119.94 119.50 C2AC3AC4AC5 0.00 0.02
C5AC6 1.419 1.422 1.416 C4AC5AC6 119.71 119.94 120.65 C3AC4AC5AC6 0.00 0.01
C6AC7 1.456 1.457 1.458 C1AC6AC7 122.30 121.05 120.59 C4AC5AC6AC7 180.00 179.93
C5AO8 1.341 1.342 1.363 C5AC6AC7 119.00 120.37 120.96 C3AC4AC5AO8 180.00 179.98
C3AO9 1.360 1.360 1.349 C6AC5AO8 122.75 122.69 122.00 C1AC2AC3AO9 180.00 179.99
C7AO10 1.229 1.249 1.253 C4AC5AO8 117.54 117.37 117.41 C1AC6AC7AO10 180.00 179.97
C7AO11 1.350 1.320 1.320 C2AC3AO9 116.77 121.87 117.00 C5AC6AC7AO11 180.00 179.93
C1AH12 1.082 1.082 1.082 C4AC3AO9 122.18 117.13 122.10 C3AC2AC1AH12 180.00 180.00
C2AH13 1.082 1.085 1.082 C6AC7AO10 124.54 122.23 122.60 C6AC1AC2AH13 180.00 179.98
C4AH14 1.085 1.082 1.084 C6AC7AO11 114.95 116.07 115.51 O9AC3AC4AH14 0.00 0.00
O8AH15 0.984 0.981 0.985 O10AC7AO11 120.51 121.70 122.00 C6AC5AO8AH15 0.00 0.04
O9AH16 0.964 0.963 0.976 C6AC1AH12 118.50 118.44 118.80 C2AC3AO9AH16 180.00 0.01
O11AH17 0.969 0.998 1.003 C2AC1AH12 119.88 120.08 119.82 C6AC7AO11AH17 180.00 179.99
O10AH15 1.747 1.751 C3AC2AH13 119.27 120.33 119.30 O10AC7AO11AH17 0.00 0.05
C1AC2AH13 121.82 120.61 121.60
C5AC4AH14 118.43 119.61 119.70
C3AC4AH14 121.54 120.45 120.90
C5AO8AH15 107.73 107.76 107.90
C3AO9AH16 110.12 110.01 111.50
C7AO11AH17 106.79 110.50 111.60
RMSD 0.0148 0.0078 RMSD 1.35 1.57
Inter-molecular bond, angles and dihedral angles
Bond Value(Å) Bond angle Value (°) Dihedral angle value
length (°)
O10AH34 1.673 C7AO10AH34 126.88 C7AC11AO27AC24 0.03 C7AO10AO28AC24 0.03
H17AO27 1.673 C24AO27AH17 126.88 C23AC24AO27AH17 179.98 O11AC7AO10AH34 0.02
O28AC24AO27AH17 0.02 C6AC7AO10AH34 179.98
a
Ref. [9].

biggest bond difference between the experimental and calculated 2,4-DHBA were defined in Table S2. Based on these definitions, a
bond length was 0.0341 Å of O11AH17 in monomer, while it was non-redundant set of local symmetry coordinates were summa-
0.005 Å in dimer structure. The CAC bond lengths of the ring are rized in Table S3, which were constructed by a suitable linear com-
varying in the range 1.389–1.419 Å in monomer and 1.379– bination of internal coordinates following the recommendations of
1.422 Å in dimer. The six-member ring in benzene is a perfect hexa- Pulay et al. [17–19]. The 45 vibrational modes of 2,4-DHBA were
gon, the phenyl ring of 2,4-DHBA appears little distorted, due to the dispersed among the symmetry species as Cvib = 31A0 (in-
substitutions of the hydroxyl group and carboxyl group. The com- plane) + 14A00 (out-of-plane) in agreement with Cs point group
putational result showed that the torsional angles of symmetry. All vibrations were both Raman and IR active. The
C5AC6AC1AC2, C3AC4AC5AC6, O9AC3AC4AH14 and O10AC7A detailed vibrational assignment of fundamental modes of 2,4-
O11AH17 are nearly 0.0°or 180.0°, indicating the optimized DHBA along with the calculated IR and Raman intensities and nor-
structure of 2,4-DHBA is planar. mal mode descriptions (characterized by PED) was summarized in
Intermolecular hydrogen bonds are responsible for the stability Table 2. In order to obtain reliable data, refinement of scaling fac-
and geometry of a predominant conformation [33]. Depending on tors was applied and optimized via least square refinement algo-
our calculation, it showed that some bonds are affected by the rithm which resulted in RMSD of 9.95 cm1 for 2,4-DHBA. The
dimerization. The shortening of the single C7AO11 bond upon observed and calculated wavenumbers were in close agreement
dimerization is due to the redistribution of partial charges on the with each other. For visual comparison, the observed and simu-
O11 atom as the unpaired electron is significantly delocalized lated FT-IR and FT-Raman spectra of 2,4-DHBA were presented in
and thereby the C7AO11 bond shows considerable double bond Figs. 4 and 5 respectively, which help to understand the observed
character. Similar effect can also be seen in bond angle spectral features.
C7AO11AH17 with an increase of 4.81°. Intermolecular hydrogen
bonds O15AH34, H17AO32 are predicted as 1.673 Å, which are
well within the range <2.6 Å for hydrogen bond interaction. OAH vibrations
The hydroxyl group vibrations are likely to be the most sensi-
tive to the environment, so hydrogen bond will have a large effect
Analysis of vibrational spectra on the OAH stretching vibrations. A free hydroxyl group gives
peaks at the range of 3600–3550 cm1, whereas the existence of
The title molecule consists of 17 atoms and thus exhibits 45 hydrogen bond can lower the OAH stretching wavenumber to
normal modes of vibration. Detailed description of vibrational the 3550–3200 cm1 region [34], and its intensity was medium
modes can be designated by means of normal coordinate analysis. to strong in the IR spectra but was generally weak in the FT-Raman
For this purpose, the full set of standard internal coordinates of spectra. The title compound in this study showed a very strong
1082 Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085

Table 2
a,b
Comparison of the experimental and theoretical wavenumbers (cm1) of 2,4-DHBA at B3LYP/6-311++G (d, p) level .

No.(i) Experimental Theoretical IInfrared IRaman Assignment with PED (%)c


IR Raman Unscaled Scaled
1 3826 3588 15.12 13.63 vOH(100)
2 3771 3506 21.63 13.00 vOH(100)
3 3373s 3433 3373 57.65 10.03 vOH(100)
4 3098vw 3214 3098 0.25 23.48 vCH(99)
5 3199 3085 0.04 9.76 vCH(99)
6 3063vw 3177 3063 0.68 18.76 vCH(99)
7 1653vs 1645vs 1718 1660 100.00 99.17 vC@O(27) vCC(17)
8 1630vs 1610 m 1662 1631 42.57 23.88 vCC(40) vC@O(21)
9 1590w 1586w 1623 1602 14.92 9.11 vCC(47) bOH(10)
10 1518 m 1537 1518 21.95 0.38 vCC(29) bCH(20) bOH(13)
11 1450vs 1480vw 1486 1460 12.03 27.70 vCC(32) bCH(21)
12 1410 m 1428w 1436 1428 13.17 7.42 bOH(42)
13 1346 m 1338s 1401 1368 37.28 39.85 vCOa(12) bC@O(12)
14 1357 1335 0.82 23.12 vCO(26) vCC(18) Rtrigd(14)
15 1330 1303 11.65 9.29 vCC(42) bOHa(13)
16 1262s 1265w 1256 1245 7.54 18.65 bCH(26) vCC(20) bOH(12)
17 1222vs 1219 m 1221 1207 47.28 22.26 bOHa(22)bOH(21) bCH(15) vCCa(13)
18 1211 1181 19.44 8.25 vCO(43) bCH(24)
19 1185s 1187w 1184 1170 46.92 20.09 bOH(22) bCH(11) vCCa(11) bOHa(10)
20 1152s 1155vw 1153 1134 35.46 3.81 bCH(27) vCO(16)
21 1082 m 1087vw 1091 1075 17.07 2.88 Rtrigd(34) vCOa(26)
22 982 m 981 m 991 994 4.37 17.22 vCC(28) vCO(19) Rtrigd(10) Rasymd0 (10)
23 970 977 0.10 0.01 xCH(85)
24 848 m 850vw 834 845 5.95 0.21 xCH(56) sRpucking(19)
25 832vw 823 835 0.32 0.00 xCH(76)
26 773 m 780 783 22.92 0.27 sCO(50) sRpucking(17) xCOa(13)
27 782vs 764 762 1.57 100.00 Rasymd(34) vCC(29) vCO(16)
28 748vw 756vs 758 753 3.84 0.75 xCOa(41) sCO(33) xCCa(12) sRpucking(12)
29 751 744 1.66 4.29 bC@O(19) vCOa(17) vCCa(15)
30 687 695 4.01 1.45 sRpucking(63) xCO(20)
31 642vw 645 636 0.01 2.14 xCO(58) sRasymd0 (34)
32 624 m 610 618 10.18 7.39 bC@O(41) Rasymd0 (17)
33 605w 608w 599 609 1.82 14.82 bCCa(22)bCO(29) vCC(12)Rasymd(10) bCOa(10)
34 548 551 12.80 8.79 sCOa(79)
35 533w 531 m 531 547 0.88 17.31 bCO(27) Rasymd(20) Rasymd0 (15) bCOa(11)
36 451 458 4.19 3.24 sRasymd(53) xCO(19)
37 473w 460w 438 442 1.55 11.32 bCOa(35) bCO(25) bC@O(12)
38 399w 377 380 19.90 12.08 sCO(91)
39 373 377 1.18 38.84 bCO(47) Rasymd0 (16) vCC (11) vCCa (10)
40 347 353 1.24 4.06 bCO(39) vCCa(17) Rasymd0 (13) Rasymd(12)
41 274 274 0.04 5.45 xCCa(44) sRasymd0 (40)
42 245w 226 230 0.56 6.63 bCCa(59) bCOa(20)
43 225 229 0.00 11.94 sRasymd(27)sRpucking(25)xCH(16)
44 110vw 105 106 0.02 5.95 sCO(31) sCCa(29) xCCa(16)
45 95 95 0.14 35.74 sCCa(56) sRasymd0 (23) xCCa(13)
a
Frequency (cm1), IR intensities: IInfrared (km mol1), Raman intensities: IRaman (a.u.).
b
Abbreviations: vs, very strong; s, strong; m, medium; w, weak; vw, very weak.
c
Primitives with more than 10% weight in normal mode are given.

Fig. 4. Experimental and calculated FT-IR spectra of 2,4-DHBA. Fig. 5. Experimental and calculated FT-Raman spectra of 2,4-DHBA.
Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085 1083

absorption peak at 3373 cm1 in the FT-IR spectrum which was scaled theoretical wavenumbers of ring CAC stretching modes
due to the OAH stretching vibration. The calculated value of this were matched well with the experimental values.
mode was found to be in consistent with the experimental data, The bands ascribed at 1082 cm1 in the FT-IR spectrum and
and this mode was pure stretching mode as it was evident from 1087, 782 cm1 in the FT-Raman spectrum have been designated
PED. to ring in-plane bending modes by careful consideration of their
The OAH in-plane bending vibrations usually occur in 1440– quantitative descriptions. The ring out-of-plane bending modes
1260 cm1 [35,36] and the OAH out-of-plane deformation vibra- of 2,4-DHBA were also listed in Table 2.
tion lie in 570–710 cm1 [36]. In the present study, three IR bands
observed at 1410, 1222, 1185 cm1 which were corresponding to HOMO–LUMO energy gap
Raman bands at 1428, 1219, 1187 cm1, have been designated as
OAH in-plane bending modes. The bands observed at 773 cm1 The frontier molecular orbital plays a critical role in the electric
in FT-IR spectrum and 399 cm1 in FT-Raman spectrum were asso- and optical properties. Both the highest occupied molecular orbital
ciated with OAH out-of-plane bending modes, which were in com- (HOMO) and lowest unoccupied molecular orbital (LUMO) are the
bination with other modes such as ring deformation mode. main orbitals that take part in chemical stability [38]. Atomic orbital
compositions and energy levels of the HOMO and LUMO orbital
CAH vibrations computed at the B3LYP/6-311++G (d, p) level for the title compound
The aromatic structure shows the presence of CAH stretching were shown in Fig. 6 and the positive and negative phases were rep-
vibrations in the range from 3100 to 3000 cm1 [34,36], which is resented in red and green color, respectively. The HOMO–LUMO
the characteristic region for the identification of the CAH stretch- energy calculation revealed that there were 40 occupied and 243
ing vibrations. In this region, the bands are not affected appreciably unoccupied molecular orbital associated with 2,4-DHBA. The ener-
by the nature and position of the substitutes. Hence in the present gies corresponding to the highest occupied and lowest unoccupied
investigation, FT-Raman bands at 3098 and 3063 cm1 have been molecular orbital were found to be 9.26820 and 5.37425 eV,
designated to CAH stretching vibrations, which were pure modes respectively. LUMO as an electron acceptor represents the ability
and were supported by calculations. of electron accepting, HOMO represents the ability of electronic giv-
The CAH in-plane bending vibrations are usually occurring in the ing. Energy difference between HOMO and LUMO orbital is called
region 1290–990 cm1 and the CAH out-of-plane bending vibra- energy gap that is an essential factor for structural stability [39].
tions were well identified in the recorded spectrum within its char- The frontier orbital energy gaps (ELUMO–EHOMO) in the case of 2,4-
acteristic region 900–650 cm1 [36]. The Raman and infrared bands DHBA was found to be 3.89395 eV. According to Fig. 6, it shows that,
observed at 1265, 1155 cm1 and 1262, 1152 cm1, respectively, the HOMO is located mainly over the O atoms (O8, O9, O10 and O11)
have been designated to CAH in-plane bending modes. The FT-IR and C atoms (C2, C3, C5 and C6) of ring, and the LUMO is
peak was identified at 848 cm1 and the FT-Raman peak was mainly located over the ring and carbonyl, hydroxyl group sketch.
located at 850 and 832 cm1 which were occurring due to the effect Consequently, the HOMO–LUMO transition implies an electron
of CAH out-of-plane bending vibrations. density transfer to COOH and OH group on the aromatic ring.

Molecular electrostatic potential (MEP)


CAO vibrations
In general, the position of C@O stretching vibration is also very
The molecular electrostatic potential (MEP) is a very convenient
sensitive to hydrogen bond. The intensity of the carbonyl group can
descriptor in understanding sites of electrophilic attack and nucle-
increase by the conjugation or formation of hydrogen bonds. Based
ophilic reaction for the study of biological recognition process and
on this, the C@O stretching vibration for carboxylic acids gives rise
hydrogen bonding interactions [40]. In order to visualize the most
to a band which is stronger than that for ketones or aldehydes. As
probable sites of 2,4-DHBA, MEP has been plotted for the title
expected, in the present study, the most intense band at 1653 cm1
in the FT-IR spectrum and 1645 cm1 in the FT-Raman spectrum
was related to C@O stretching mode. The match between theory
and experiment was excellent, for calculation, it showed at
1660 cm1. The above assignment agrees well with the value avail-
able in the literature [36].
The CAO stretching band was observed at 1346 cm1 in FT-IR
spectrum and 1338 cm1 in the FT-Raman spectrum, which was also
strongly couple with other modes. The in-plane and out-of-plane
bending vibrations of C@O group have been identified and presented
in Table 2, both of them were coupled with other modes, too.

Ring vibrations
Most of the ring vibrational modes are influenced by the substi-
tutions in the ring. The characteristic ring stretching vibrations of
variable intensity were appeared at between 1625 and
1430 cm1 [36,37]. In the present investigation, the observed infra-
red bands at 1630, 1590, 1450 cm1 and three Raman bands at
1610, 1586, 1480 cm1 were considered as ring CAC stretching
vibrations. Among them, the bands 1630 and 1610 cm1 were
sharp and strong, and other bands were weak in intensity which
were overlapped by the CAH or OAH deformation vibrations, sim-
ilar to previously reported [36]. In addition, we also observed a
weak Raman band 981 cm1 and infrared band 982 cm1 due to Fig. 6. Frontiers molecular orbitals of 2,4-DHBA by B3LYP method using 6-311++G
ring CAC stretching vibration, just as Socrates predicted [36]. The (d, p) basis set.
1084 Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085

natural charge (0.79448 e) when compared to other atoms and


become more acidic. These results show that substitution of COOH
and OH group in 2,4-DHBA leads to a redistribution of electron
density. In the title molecule, both Mulliken and natural atomic
charges show that all the hydrogen atoms have positive charges
which behave as electron donor and all the oxygen atoms have
negative charges which behave as electron acceptor. It suggests
the presence of both inter-molecular bonding in the gas phase.

Conclusion

In this work, quantum chemical calculations were applied to


investigate the conformation, stability and vibrational spectra of
2,4-DHBA by using density functional theory (DFT) method with
the B3LYP/6-311++G (d, p) basis set combination. The theoretical
calculations indicate the presence of sixteen stable conformers,
based on calculated energy differences, the most stable conformer
was found, and optimized geometrical parameters for the most
stable were agreed well with literature value. The harmonic vibra-
tional frequencies of the compound were analyzed and compared
with experimental data, and shown to have an excellent correla-
tion with each other. Furthermore, the HOMO–LUMO band gap
of 2,4-DHBA molecule was determined to be about 3.89395 eV,
which directs the molecule becomes less stable and more reactive.
Mulliken atomic charges show that the substitution of hydroxyl
Fig. 7. Molecular electrostatic potential of 2,4-DHBA. group and carboxyl group leads to redistribution of electron den-
sity. In particular, the MEP map reveals the negative potential sites
are on oxygen atoms for electrophilic attack as well as the positive
molecule in B3LYP/6-311++G (d, p) basis set using the computer potential sites are around the hydrogen atoms for nucleophilic
software GaussView 5.0 [41]. The 3D electrostatic potential contour attack. This study demonstrates that the usefulness of quantum
map was shown in Fig. 7. The different values of the electrostatic mechanical calculations for understanding the structural proper-
potential at the surface were represented by different colors. Poten- ties and vibrational spectra of organic compounds.
tial increases in the order red < orange < yellow < green < blue1, and
the color code of these maps is in the range between 0.030 a. u. Acknowledgement
(deepest red) to 0.030 a. u. (deepest blue), in other words, the electro-
static potential ranges from 78.765 to +78.765 kJ/mol. Blue indi- This work was supported by National Natural Science Founda-
cates the molecule’s the positive region which implies strongest tion of China ( No. 11304140), Science and Technology Project of
nucleophilic reactivity and red demonstrates the molecule’s the neg- Luoyang (No. 1101041A), the Henan Joint Funds of the National
ative region which displays electrophilic reactivity. From the MEP, it Natural Science Foundation of China (No. U1204109) and Science
was evident most surface of negative charge covers the ring and oxy- and Technology Development Foundation of Henan Province (No.
gen (O8, O9, O10, O11) atom, showing the most favorable site for elec- 142102210109).
trophilic attack. When focused on positive regions of the electrostatic
potential, we found that the hydrogen atoms of the phenyl ring are
Appendix A. Supplementary material
surrounded by blue color, indicating that these sites are probably
involved in nucleophilic processes. Taking into account these results,
Supplementary data associated with this article can be found, in
these sites confirm the existence of intermolecular interactions
the online version, at http://dx.doi.org/10.1016/j.saa.2014.08.151.
observed in the solid state.

References
Atomic charge distribution
[1] R. Świsłocka, E. Regulska, M. Samsonowicz, W. Lewandowski, J. Mol. Struct.
1044 (2013) 181–187.
The charge distribution has been proved to be an efficient [2] M. Nsangou, Z. Dhaouadi, N. Jaidane, Z. Ben Lakhdar, J. Mol. Struct. 850 (2008)
method in the quantum chemical calculation of molecular systems 135–143 (Theochem.).
[1,15,35]. Atomic charges are directly related to the molecular [3] S. Vecchio, B. Brunetti, Thermochim. Acta 515 (2011) 84–90.
[4] R. Mathammal, N. Jayamani, N. Geetha, J. Spectrosc. 2013 (2013) 1–19.
orbital characteristics, especially the chemical bonds present in [5] E. Ritzer, R. Sundermann, Hydroxycarboxylic Acids, Aromatic, Berlin, 2000.
the molecule. The charge distributions have been calculated on [6] C.R.L. Martin, Crystal Engineering Approaches to Controlling the Formation of
the various atoms of 2,4-DHBA by the Mulliken and NBO methods, Molecular Complexes and their Polymorphs, School of Chemistry, University of
Glasgow, 2012.
and the results were presented in Table S4. In general, the relative
[7] B. Velika, I. Kron, Free Radicals Antioxidants 2 (2012) 62–67.
values of the atomic charges can offer valuable information. For [8] K. Azrague, V. Pradines, E. Bonnefille, C. Claparols, M.T. Maurette, F. Benoit-
example, the charge of an atom of C6 is the highest value for posi- Marquie, J. Hazard Mater. 237–238 (2012) 71–78.
[9] A. Parkin, M. Adam, R.I. Cooper, D.S. Middlemiss, C.C. Wilson, Acta Crystallogr.
tive Mulliken charge (1.046404 e), however this charge is less
B 63 (2007) 303–308.
value for negative natural charge (0.27906 e). Meanwhile, the [10] D.E. Braun, P.G. Karamertzanis, S.L. Price, Chem. Commun. 47 (2011) 5443–
carbon atom C7 in the COOH group accommodates higher positive 5445.
[11] D. Hatzipanayioti, K. Kontotheodorou, Spectrochim. Acta A 78 (2011) 949–960.
[12] S.A. Brandán, F. Márquez López, M. Montejo, J.J. López González, A. Ben Altabef,
1
For interpretation of color in Fig. 7, the reader is referred to the web version of Spectrochim. Acta A 75 (2010) 1422–1434.
this article. [13] X. Xuan, M. Guo, Y. Pei, Y. Zheng, Spectrochim. Acta A 78 (2011) 1492–1499.
Y. Tao et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 137 (2015) 1078–1085 1085

[14] X. Xuan, N. Wang, Z. Xue, Spectrochim. Acta A 96 (2012) 436–443. [25] J.B. Du, Y.L. Tang, Zh.W. Long, Sh.H. Hu, T. Li, Rus. J. Phys. Chem. A 88 (2014)
[15] E. Yabalak, F. Gunay, V.T. Kasumov, H. Arslan, Spectrochim. Acta A 110 (2013) 819–822.
291–303. [26] X. Xuan, C. Zhai, Spectrochim. Acta A 79 (2011) 1663–1668.
[16] R. Świsłocka, E. Regulska, M. Samsonowicz, W. Lewandowski, Polyhedron 28 [27] R. Swislocka, Spectrochim. Acta A 100 (2013) 21–30.
(2009) 3556–3564. [28] T. Sundius, J. Mol. Struct. 218 (1990) 321–326.
[17] P. Pulay, G. Fogarasi, G. Pongor, J.E. Boggs, A. Vargha, J. Am. Chem. Soc. 105 [29] T. Sundius, Vib. Spec. 29 (2002) 89–95.
(1983) 7037–7047. [30] G. Keresztury, S. Holly, G. Besenyel, J. Varga, Spectrochim. Acta A 49 (1993)
[18] P. Pulay, G. Fogarasi, F. Pang, J.E. Boggs, J. Am. Chem. Soc. 101 (1979) 2550– 2007–2026.
2560. [31] G. Keresztury, J.M. Chalmers, P.R. Griffith (Eds.), Raman Spectroscopy: Theory,
[19] G. Rauhut, P. Pulay, J. Phys. Chem. 99 (1995) 3093–3100. John Wiley & Sons Ltd., New York, 2002.
[20] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, [32] H. Ghalla, N. Issaoui, M. Govindarajan, H.T. Flakus, M.H. Jamroz, B. Oujia, J. Mol.
G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, Struct. 1059 (2014) 132–143.
X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M.Hada, [33] P. Muniappan, R. Meenakshib, G. Rajavel, M. Arivazhagan, Spectrochim. Acta A
M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, 117 (2014) 739–753.
O. Kitao, H. Nakai, T. Vreven, J.A., Montgomery Jr., J.E. Peralta, F. Ogliaro, M. [34] N.B. Colthup, L.H. Daly, S.E. Wiberley, Introduction to Infrared and Raman
Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, T. Keith, R. Spectroscopy, Academic Press, New York, 1990.
Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. [35] V. KrishnaKumar, D. Barathi, R. Mathammal, J. Balamani, N. Jayamani,
Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, Spectrochim. Acta A 121 (2014) 245–253.
C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. [36] G. Socrates, Infrared and Raman characteristic group frequencies tables and
Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, charts, third ed., John Wiley & Sons Ltd., Chichester, 2001.
G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, O. Farkas, J.B. [37] K. Govindarasu, E. Kavitha, Spectrochim. Acta A 122 (2014) 130–141.
Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Revision D.01, in, [38] M. Srivastava, P. Rani, N.P. Singh, R.A. Yadav, Spectrochim. Acta A 120 (2014)
Gaussian Inc., Wallingford CT, 2013. 274–286.
[21] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789. [39] S. Muthu, A. Prabakaran, Spectrochim. Acta A 121 (2014) 420–429.
[22] A.D. Becke, J. Chem. Phys. 104 (1996) 1040–1046. [40] S. Saravanan, V. Balachandran, Spectrochim. Acta A 120 (2014) 351–364.
[23] R. Świsłocka, Spectrochim. Acta A 111 (2013) 290–298. [41] R. Dennington, T. Keith, J. Millam, GaussView 5.0.9, in: K.S. Shawnee Mission,
[24] M. Xie, Y. Qi, Y. Hu, J. Phys. Chem. A 115 (2011) 3060–3067. Semichem. Inc., 2009.

You might also like