0% found this document useful (0 votes)
18 views5 pages

Martins Et Al 2019 Raman

martins et al 2019 raman
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
18 views5 pages

Martins Et Al 2019 Raman

martins et al 2019 raman
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 5

Vibrational Spectroscopy 104 (2019) 102945

Contents lists available at ScienceDirect

Vibrational Spectroscopy
journal homepage: www.elsevier.com/locate/vibspec

Investigation of ostrich oil via Raman and infrared spectroscopy and T


predictions using the DFT method

Q.S. Martinsa,b, , P.V. Almeidaa, Q.S. Ferreiraa, A. Oliveiraa, C. Aguirreb, J.L.B. Fariab
a
Departamento de Física, Universidade Federal de Rondônia, UNIR, Ji-Paraná, Brazil
b
Instituto de Física, Universidade Federal de Mato-Grosso, Cuiabá, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: Ostrich oil (OA) containing omega-3, 6, 9, was used to obtain vibrational signatures via Raman and Fourier
Ostrich oil transform infrared spectroscopy, comparing the results generated by ab initio calculations, following a quali-
Fatty acids tative transversal approach. The density functional theory (DFT) model, whose molecular geometries were based
Raman spectroscopy on the B3LYP functions with the 6-311G basis set, performed by GAUSSIAN 09 packages, was chosen due it is
FTIR
widely used for calculations involving organic compounds. The symmetries, vibrational assignments and cal-
DFT-calculation
culations of the potential energy distribution (PED) were done using vibrational energy distribution analysis
(VEDA) software. A bibliographical approach, based on typical samples, such as oleic, palmitic, linoleic, pal-
mitoleic, stearic and linolenic acids (fatty acids), was submitted to the theoretical stage. The results, experi-
mental (FTIR/Raman) and calculated spectra (CS) showed modes in the regions between 1400 and 1800 cm−1,
directly related to C=O and C=C belonging to carbon chains and in the range 3300–3500 cm−1, related to the
OH stretching modes. The modes showed a good approximation of the composition of the material, for both
FTIR/Raman and CS, therefore justifying to the use of the methodology to certify the vibrational signature of the
sample.

1. Introduction ingredients to transport them below the skin barrier [2,3], among other
benefits [4–7]. Majewsk [8] and others authors [9–13] have cited oleic,
The physical and chemical properties of oil are mainly related to its palmitic, and linoleic acids as the most abundant in concentrations of
composition of fatty acids (FA) and degree of unsaturation. FA are or- ostrich oil, and in smaller quantities palmitoleic, stearic, and linolenic
ganic substances found in solid or semi-solid (fats) and liquid phases acids. In this work, FTIR and Raman spectroscopy was used for the
belonging to the carboxylic acid group, compounds that have carboxyl, characterization of ostrich oil. Both are spectroscopic vibrational
carbon double bonded to oxygen and hydroxyl groups. The difference techniques for studying vibrational transitions in molecules. First, in
between fat and oil is the physical state at room temperature, where fats FTIR spectroscopy, the samples are irradiated with IR light. The vi-
contain more highly saturated fatty acids, and liquids unsaturated ones brational state is reached when part of the IR light is absorbed by the
[1,2]. FA represent the predominant component of ostrich oil, with a molecules, but only vibrations resulting in changes in the dipole mo-
lipid content of 98.8% from subcutaneous adipose tissue and 98.0% ment of a molecule can absorb the IR radiation. A dipole is the product
from retroperitoneal adipose tissue. Also described as polyunsaturated of charge (positive and negative) and distance. The last is a physi-
fatty acids (PUFA's), several studies correlate their therapeutic effects cal–chemical technique that shows the effects of the molecular polar-
(anti-inflammatory, antibacterial and antifungal) with their ability to ization tensor in order to present its respective vibrational patterns. The
modify cell membrane phospholipids, modify cellular functions, and technique has been widely applied for presenting fast and safe results,
play a protective role for normal tissues [3]. Recently, in commercial as it does not require any sample preparation and is easily applied to
research, oleic (omega-9), linoleic (omega-6) and linolenic (omega-3) non-crystalline materials [14–16]. The Raman spectra were compared
acids were recognized as essential fatty acids. They represent products with ab initio calculations using DFT for the components presented.
of commercial interest, containing certain vitamins and amino acids Thus Raman, as well as others [17–21], can be combined with che-
that help; for example, in the salubrity of the skin membranes, used as a mometric methods, providing a good approach to the analysis of several
mediator in combination with various medicinal or cosmetic materials [22–27]. The results obtained allowed presenting a spectral


Corresponding author at: Departamento de Física, Universidade Federal de Rondônia, UNIR, Ji-Paraná, Brazil.
E-mail address: quesle@fisica.ufmt.br (Q.S. Martins).

https://doi.org/10.1016/j.vibspec.2019.102945
Received 1 April 2019; Received in revised form 22 July 2019; Accepted 23 July 2019
Available online 29 July 2019
0924-2031/ © 2019 Elsevier B.V. All rights reserved.
Q.S. Martins, et al. Vibrational Spectroscopy 104 (2019) 102945

pattern or vibrational signatures for the ostrich oil and could be com-
bined with an investigation about the general nature of FA. In the DFT
calculations, the base was 6311G for the B3LYP functional, contributing
to the processes of identification of the chemical nature, without pre-
judice to other identification methods or other bases. The results ob-
tained showed a spectral pattern of vibrational signatures for the os-
trich oil, exhibiting affinities with the general nature of the FA. Modes
related to the O–H, C–H, and C=C bonds are well defined and may
correspond to the molecules of flavons, polyphenols, and heptadeca-
noic, stearic, ecosanoic, palmitic, and palmitoleic acids, as well as
omegas 3, 6 and 9.

2. Material and methods

The ostrich oil, sample came from Rondônia-Brazil in the micro-


region of Ji-Paraná was provided by Amazon Struthio, considering the
necessary and essential procedures for preserving the quality of the
material. For sample analysis, Raman and FTIR spectroscopy, and DFT Fig. 1. Experimental FTIR of pure ostrich oil.
as an analytical method were used.

2.1. Raman and FTIR spectroscopy

Confocal laser scanning microscopy (CLSM) was used on Horiba


Xplora Series Raman equipment, spectral range from 0 to 4000 cm−1,
single laser 532 nm, confocal imaging 0.5 μm XY at high resolution.
FTIR was carried out on an IRPrestige-21 Shimadzu Fourier transform
infrared spectrophotometer with a resolution of 4 cm−1, accumulation
of 1 min, in the region of 2100 cm−1 with a reading range of
0–4000 cm−1.

2.2. DFT calculations

The molecular calculations were performed according to the density


functional theory (DFT) model, whose molecular geometries were
based on the B3LYP function with set 6-311G [26–28] in GAUSSIAN 09
packages [29]. PED calculations were carried out with the aid of the
VEDA 4 program [32]. DFT was used to obtain the optimized structure
of the compound, and its harmonic energy and vibrational frequencies Fig. 2. Pure ostrich oil Raman spectrum.
were calculated with a hybrid three-parameter function (B3) used for
the exchange part and Lee-Yang-Parr (LYP) functional correlation (a addition to the previously mentioned omegas-3, 6, and 9. The peaks at
functional correlation that has local and nonlocal terms). The LYP 1378 and 1467 cm−1 (Fig. 1) correspond to the bending of the methyl
correlation function is accepted as the most cost-effective approach for and methylene groups (CH groups), and the peaks at 1650 cm−1 to
calculating the molecular structure, vibrational frequencies, and energy C=C and 1744 cm−1 (1742 cm−1) (Fig. 1) to the stretching of C=O.
of optimized structures [25–31]. The symmetries, vibrational assign- Since C=O is a strong absorbent, we see that C=C absorbs much less,
ments, and calculations of the potential energy distribution (PED) were but with extra-close bands, that overlapping such a band mainly in the
done with a high degree of accuracy. VEDA optimizes the set of internal IR spectrum, due to bond symmetry. In this particular case, the C=C
coordinates for the development of IR/Raman experimental/theoretical mode appears at the base of the band of C=O, practically hidden in the
systems. The results are represented by internal coordinates, giving analysis. The spectra show a strong C–H vibrational mode between
users an overview of the use of multiple directions [33–38]. 3000 and 2850 cm−1. Fig. 1 clearly shows separate bands corre-
sponding to asymmetric C–H stretching at 2929 cm−1 and symmetrical
3. Results and discussion C–H stretching at 2856 cm−1. Typically, OH absorption bands can be
observed in the 3800–3200 cm−1 stretching region. On the other hand,
3.1. FTIR analysis we can see that this region has no absorbance pattern and remains
unchanged, due to oxidative conditions, which strongly depend on the
After the experimental readings, it was observed in Fig. 1 that the nature of the sample. In both Figs. 1 and 2, this expected mode is not
722 cm−1 band can correspond to the bending of the –(CH2)n–HC=CH observed. On the other hand, in theoretical terms (Fig. 5), this mode is
group, as well as the C–O stretching [39–41]. The presence of the peak visible, suggesting its existence. The peaks at 2853, 2922, and
in Fig. 2 correspond to the C–C stretching [42]. The existence of this 3005 cm−1 correspond to the C–H, CH2, and –CH3 stretches, respec-
peak indicates the presence of long carbon chains and point out a tively. Finally, there is at peak at 3003 cm−1 in Fig. 1 is related to a cis
sample characteristics. The peaks at 1095 and 1118 cm−1 (Fig. 1) double bond and stretching of =C–H [39,40,44]. The vibrational pat-
correspond to the stretching of CO, while the peak at 1745 cm−1 is the terns observed in Fig. 1 can be compared with Fig. 2. As complementary
stretching of C=O [43]. The peaks previously mentioned at 722, 1095, techniques, the adopted procedure can validate the mode set in each
and 1118 cm−1 represent the vibrations of omega-3 (linolenic acid), technique. Important bands can be detected, for example C=O
omega-6 (linoleic acid), and omega-9 (oleic acid). The 1160 cm−1 peak (1650–1745 cm−1), OH (2800–3200 cm−1) and long-chain carbon
represents the flavone, polyphenol, heptadecanoic, stearic, ecosanoic, bands, which break down below 750 cm−1 in IR.
pentadecanoic, myristic, palmitic, and palmitoleic acid vibrations in

2
Q.S. Martins, et al. Vibrational Spectroscopy 104 (2019) 102945

3.2. Raman analysis presenting the base formation fragment of the FA molecules is also
shown. The importance of presenting the calculation for glycerol is
Fig. 2 presents a characteristic spectrum of pure ostrich oil in the precisely to understand the formation of vibrational modes when they
region between 867 and 1745 cm−1, since that it is composed of long contain the hydroxyls present in the chemical formation for the cal-
carbonic chains. Linolenic, linoleic, and oleic acids are some of the culated components. CS comprise the region from 0 to 4000 cm−1 with
acids present in the sample (in Fig. 5. At 2721, 2857, 2896, and the main modes appearing between 1400 and 2000 cm−1 and between
3011 cm−1 the modes stand out in very typical form and intensity for 2800 and 3250 cm−1. It can be seen that the mode at 3641 cm−1 and
vegetable oils. This way, with intense and overlapped, fine and medium 3637 cm−1 is a symmetrical stretch-type vibration of the OH, a char-
peaks, 2857 and 2896 cm−1 exhibits the convolution of the CH2, CH3 acteristic linkage of the FA. For glycerol, this vibration occurs at
and OH groups. They are very close to those observed in Fig. 2, but 3740 cm−1 with symmetrical stretching and at 3700 cm−1, asymmetric
differ in intensity, possibly related to the means of observation, one stretching for the OH at the opposite ends of the molecule, and between
considered an ideal medium, which is not true. Thus, because of the 1500 and 2000 cm−1, the absence of some modes may be related to
amorphous form of the material, the form at the base of the modes and C=C bonds, mainly in bands around 1650 cm−1. For A, B, C, and D,
the intensities may take different forms due to other forces or interac- mean intensities of C=O bonds are 1720 cm−1, which can be observed
tions with the environment. at very low intensity for E and F. At 1525 cm−1, we see the angular CH
Peaks at 867, 1079, 1300, 1433, 1655, and 1745 cm−1 (Fig. 2) are deformations within the molecules. Still, for the main components ex-
present in common Raman spectra and can be found in almost all edible cept G, we see the active modes at 1340, 1325, 1155, 1105, and
vegetable oils. Huang (2016), studying olive, peanut, and corn cooking 1075 cm−1. In the region of 2900–3050 cm−1, asymmetric stretch type
oil, points out eight of these peaks and suggests that in Raman spec- vibrations for CH are exhibited for all components. Table 2 presents
troscopy different oils and fats have only a small difference in relative data from the theoretical process, but emphasizes the molecules of
peak intensity at 1654, 1265, and 970 cm−1. Reference values (circles) oleic, linoleic and linolenic acid (omegas 9, 6 and 3), whose structures
were indicated to identify a linear correlation, which is shown in Fig. 3. are shown in Fig. 5 since they are the major constituents of ostrich oil.
The results presented in the Raman spectra, which in a simple way In assessing these results, it should be borne in mind that the modes
show us the nature of the ostrich oil in a relative pattern and analog of of C=O and C–O (carboxylic acids) are well defined, and are in the
other oils. An analogy is can be seen for the 1745 and 1742 cm−1 peaks, region of 1720 cm−1 and 1087 cm−1, except for omega 6, at 1713 cm−1
exhibiting stretching vibrations of the ester bond carbonyl, and still at and 1146 cm−1. The modes caused by the C=C conjugate stretches
1655–1300 cm−1 there are cis (C=C) and cis (CH) groups of un- generally appear as a small intensity shoulder alongside the carboxyl,
saturated fatty acid and the peak at 970 cm−1, which is related to trans- between 1700 and 1715 cm−1, which is true for the calculated cases of
bending vibration (C=C) [28]. In this condition, Raman and IR bands omega 9 and 6. In Figs. 1 and 2, C=C appears quite discreetly at the
show a clear change in intensity, and due to this we can indicate the base of the peak at 1742 cm−1. It can be seen that in Fig. 5 this in-
vibrational selection factor for the type of technique. One is based di- formation is hidden at the base of the spectra but is shown at its re-
rectly on the variation of the induced dipole, while the other selects the spective frequencies after DFT analysis. Among the aspects presented in
vibrational modes based on the variation of molecular polarizability. Of the region 1320–1420 cm−1, the lower intensity modes are the re-
course we may have similar, absent or displaced bands for a respective presentations for the vibrations of the –CH bonds along the molecule,
sample. usually of the twisting type. In Table 2, DFT data suggest vibrational
modes in two general aspects: bending (δ) and stretching (ν). The
contributions such as these vibrations are made are from the values
3.3. Calculated spectra calculated under the same potential energy distribution (PED), which
appears in the signatures column. The minimum difference for these
In Fig. 4, the set of calculated spectra (CS) have similar vibrational frequencies, or the absence of them, shows that the method is also valid
modes. This is observed due to the similar nature of the molecules, for understanding the signatures of the components. It can be seen that
attributed to long carbon chains (Table 1), differentiated precisely by for the calculated components, the optimization given by the maximum
the difference in the presence of C=C. energy parameter (EPM), for oleic (EPM = 25.721), linoleic
For the CS results, the standard for the glycerol (G) molecule (EPM = 30.888), and linolenic acid (EPM = 31.078), is related to the
size of the molecule [38]. The types of vibrations may vary for the cases
mentioned, and this is closely linked to the energy received. However,
these vibrational variations are minimized in the scope of this study. A
good study on the subject can be found in other work [45]. Still, in
Table 2, other vibrational contributions appear and are indicated by S
+number. For the band at 1530 cm−1 (omega 9), we have S59, S69, and
S83, respectively, contributions of lower percentage of PED appearing
for the same molecular group, vibration harmonic. Table 2 presents a
summary format for presentation. A more complete view can be ac-
cessed in the supplementary material, Table S1.

4. Conclusions

The results of Raman/FTIR spectroscopy combined with ab initio


calculations show ostrich oil vibrational bands. Differences found in the
spectra may be related to the presence of C=C for the different com-
ponents of the oil. Different intensities for carboxylic groups are ver-
Fig. 3. Fitting for our results (α = experimental/theoretical – stars) and other ified, an indication of the form of activation of these vibrational modes
experimental results (β = reference values – circles) oils of different nature. according to the experimental technique used. In the experimental
Y = a + bx, with a = 322.294, b = 0.64. In this case, the adjustments are spectra, for example, the significant displacement of C=O
compared with more expressive results for analysis of vegetable oils observed in (1720 cm−1), may be due to the morphological relationships of the
the literature with ostrich oil. See the comparison between two lines. sample itself. As calculated, C=O appears at 1721 cm−1 for oleic and

3
Q.S. Martins, et al. Vibrational Spectroscopy 104 (2019) 102945

Fig. 4. DFT theoretical results B3LYP/6311G; Raman (left) and IR (right). (A) Linoleic; (B) Oleic; (C) Palmitoleic; (D) Linolenic; (E) Palmitic; (F) Stearic and (G)
Glycerol.

linoleic acid. Between 750 and 1750 cm−1 and between 2896 and Table 1
3011 cm−1 the present modes are not distinguishable as to the origin of Fatty acids and of ostrutch oil.
the sample, whether vegetable or animal. The DFT results helped to FA's Carbon numbers Symbol
understand the behavior of the main components of the sample. Raman
spectroscopy and FTIR were satisfactory and complementary to the Palmitic C16:0 E
characterization. Palmitoleic C16:1 C
Stearic C18:0 F
Oleic (omega 9) C18:1 B
Linoleic (omega 6) C18:2 A
Authors contributions Linolenic (omega 3) C18:3 D

All the authors were involved in the preparation of the manuscript. The symbolism given links the results of Fig. 4.
All the authors have read and approved the final manuscript.

Fig. 5. Structure fatty acids. (a) Oleic acid (b) Linoleic acid (c) Linolenic acid for all structure (green-carbon), (red oxygen) and white hydrogen.

4
Q.S. Martins, et al. Vibrational Spectroscopy 104 (2019) 102945

Table 2 H. Khalifa, W. Chyliński, A. Wójcik, M. Kawka, Anim. Sci. Pap. Rep. 22 (2004)
General patterns of vibrational behavior for calculated values for omega 9, 6 247–251.
and 3 molecules. [6] V. Baeten, P. Hourant, M.T. Morales, R. Aparicio, Agric. Food Chem. 44 (1996)
2225–2230.
−1 [7] U.D. Palanisamy, M. Sivanathan, A.K. Radhakrishnan, N. Haleagrahara,
Component ω/cm Assignment of the molecular Other PED
T. Subramaniam, G.S. Chiew, Molecules 16 (2011) 5709–5719.
Raman FTIR [8] D. Majewska, D. Szczerbińska, M. Ligocki, M. Buclaw, A. Sammel, Z. Tarasewicz,
K. Romaniszyn, J. Majewski, J. Brit. Poult. Sci. 57 (2016) 193–200.
1344 δ(HCC)(21 20 23) [14] [9] X. Liu, F. Wang, X. Liu, Y. Chen, L. Wang, Eur. J. Lipid Sci. Technol. 113 (2011)
775–779.
1464 1464 δ(HCC)(27 26 28) [41] S73
[10] D. Majewska, J. Brit. Poult. Sci. (2016) 57.
1527 δ(HCH)(38 36 37) [42] S63, S77, S81
[11] D. Belichovska, Z. Hajrulai-Musliu, R. Uzunov, K. Belichovska, M. Arapcheska, Mac.
1530 δ(HCH)(10 8 9) [19] S59, S69, S83 Vet. Rev. 38 (2015) 53–59.
Omega 9 1557 δ(HCH)(41 39 40) [22] S79, S83, S85 [12] P. la Mataa, A. Dominguez-Vidalb, J.M. Bosque, S.A. Ruiz, M.L. Cuadros,
1713 1713 ν(C=C)(28 26) [72] R.M.J. Ayora, Food Control 23 (2012) 449–455.
1721 1721 ν(C=O)(52 51) [78] [13] H. Yanga, J. Irudayaraj, M.M. Paradkarb, Food Chem. 93 (2005) 25–32.
2989 ν(CH)(20 22) [17] S5, S6, S11, S12 [14] T. De Beer, A. Burggraeve, M. Fonteyne, L. Saerensa, J.P. Remon, C. Vervaet, Int. J.
2998 ν(CH)(33 34) [40] S5, S7, S10, S13, Pharm. 30 (1-2) (2011) 32–47 417, doi: 10.1016/j.ijpharm.2010.12.012.
S16, S26, S27 [15] C.B. Silva, J.G. da Silva Filho, G.S. Pinheiro, A.M.R. Teixeira, P.T.C. Freire, Vibrat.
3122 ν(CH)(28 29) [49] S19 Spectrosc. 98 (2018) 128–133, https://doi.org/10.1016/j.vibspec.2018.08.001.
3641 3641 ν(OH)(53 54) [100] [16] A. Abkari, I. Chaabane, K. Guidara, Physica E 81 (2016) 136–144, https://doi.org/
1078 1078 ν(CC)(13 16) [31] S44, S45, S47 10.1016/j.physe.2016.03.010.
1093 1093 ν(CC)(1 4) [23] [17] X.F. Zhang, M.Q. Zou, X.H. Qi, F. Liu, C. Zhang, F. Yin, J. Raman Spectrosc. 42
(2011) 1784–1788.
1317 1317 δ(HCC)(25 24 26) [14]
[18] H. Sadeghi-Jorabchi, P.J. HendraR, H. Wilson, P.S. Belton, J. Am. Oil Chem. Soc.,
1556 δ(HCH)(11 10 12) [29] S56, S60, S64,
67, 483–486.
S66
[19] H. Yanga, J. Irudayaraj, M.M. Paradkarb, Food Chem. 93 (2005) 25–32.
1709 1709 ν(C=C) (24 22) + (29 31) S35 [20] G.F. Ghesti, J.L. de Macedo, I.S. Resck, J.A. Dias, S.C. Dias, Energy Fuels 21 (2007)
[39] 2475–2480.
Omega 6 1713 1713 ν(C=O)(50 49) [79] [21] P. Pavón, N. Sánchez, F. Laespada, M. Cordero, Anal. Bioanal. Chem. 394 (2009)
2957 2957 ν(CH)(26 27) + (26 28) [48] 1463–1470.
3007 3007 ν(CH) (45 47) + (45 48) [37] S8, S9, S12, S13, [22] F.P. Capote, J.R. Jiménez, M.D. de Castro, Anal. Bioanal. Chem. 388 (2007)
S30 1859–1865.
3141 3141 ν(CH)(24 25) [35] S7, S16, S17 [23] H. Chena, M. Angiuli, C. Ferrari, E. Tombari, G. Salvetti, E. Bramantib, Food Chem.
3637 3637 ν(OH) (51 52) [100] 125 (2011) 1423–1429.
948 948 ν(CC)(30 28) [16] [24] V. Baeten, Lipid Technol. 22 (2010) 36–38.
1055 1055 ν(CC)(30 28) [13] S37, S43 [25] A.D. Becke, J. Chem. Phys. 98 (1993) 5648.
[26] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 98 (1988) 785.
1202 1202 δ(HOC)(25 24 22) [24]
[27] X. Wu, S. Gao, J.S. Wang, H. Wang, Y.W. Huanga, Y. Zhaod, Analyst 137 (2012)
1441 1441 δ(HCH)(49 47 48) [26] S71, S77, S79
4226–4234.
Omega 3 1535 1535 δ(HCH)(46 44 45) [33] S74, S79
[28] F. Huang, Y. Li, H. Guo, J. Xu, Z. Chen, J. Zhang, Y. Wang, J. Raman Spectrosc. 47
1720 1720 ν(C=C) (28 26)[11]+(35 (2016) 860–864.
33)[38]+(42 40)[23] [29] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
1721 1721 ν(C=O) (23 22) [78] G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
2959 2959 ν(CH)(30 31) [98] X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada,
2998 ν(CH)(16 17) [36] S8, S11 M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
3133 3132 ν(CH) (35 36) [54] S21 O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, J.E. Peralta, F. Ogliaro,
3641 ν(OH) (24 25) [100] M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi,
J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi,
Here are presented the more general natures of vibrations, these being able to M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken,
C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin,
have specific variations according to the type movements of the atoms. Bending
R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski,
(δ) and stretching (ν). G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas,
J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, (2009).
Acknowledgements [30] B. Becke, Phys. Rev. B 38 (1998) 3098–3100.
[31] T.K. Kuruvilla, J.C. Prasana, S. Muthu, J. George, S.A. Mathew, Spectrochim. Acta
Part A: Mol. Biomol. Spectrosc. 188 (2018) 382–393.
The authors would like to give thanks to Amazon Struthio for the [32] M.H. Jamroz, Vibrational Energy Distribution Analysis, VEDA 4 Program, Warsaw,
sample assigned to this study, to the Grupo de Optica e Nanoscopia Poland, 2004.
[33] Y. Morino, K. Kuchitsu, J. Chem. Phys. 20 (1952) 1809–1810.
(GON) and Laboratorio de Caracterização e Microscopia de Materiais [34] W.J. Taylor, J. Chem. Phys. 22 (1954) 1780–1781.
(LCMMAT) of the Universidade Federal de Alagoas, to the research [35] T. Miyazawa, T. Shimanouchi, S. Mizushima, J. Chem. Phys. 29 (1958) 611–616.
group Estrutura da Matéria e Física Computacional/DEFIJI/UNIR and [36] G. Keresztury, G. Jalsovszky, J. Mol. Struct. 10 (1971) 304.
[37] P. Pulay, G. Fogarasi, F. Pang, J.E. Boggs, J. Am. Chem. Soc. 101 (1979)
FAPERO/CAPES (Grant number: 008/2018). 2550–2560.
[38] M.H. Jamróz, Spectrochim. Acta Part A: Mol. Biomol. Spectrosc. 114 (2013)
Appendix A. Supplementary data 220–230.
[39] N. Cebi, M.T. Yilmaz, O. Sagdic, H. Yuce, E. Yelboga, Food Chem. 225 (2017)
188–196 doi: 10.1016/j.foodchem.2017.01.013.
Supplementary data associated with this article can be found, in the [40] M.D. Guillén, N. Cabo, J. Sci. Food Agric. 80 (2000) 2028–2036 doi.org/10.1002/
online version, at https://doi.org/10.1016/j.vibspec.2019.102945. 1097-0010(200011)80:14 < 2028AID-JSFA713 > 3.0.CO;2-4.
[41] N.E. Maurer, B. Hatta-Sakoda, G. Pascual-Chagnan, L.E.R. Saona, Food Chem. 134
(2012) 1173–1180 doi: 10.1016/j.foodchem.2012.02.143.
References [42] K. Malek, E. Podstawka, J. Milecki, C. Schroeder, L.M. Proniewicz, Biophys. Chem.
142 (2009) 17–26 doi.org/10.1016/j.bpc.2009.02.007.
[1] R.D. O’Brien, AOCS Press, 2nd ed., 2000. [43] B. Innawong, P. Mallikarjunan, J. Irudayaraj, J.E. Marcy, Food Sci. Technol. 37
[2] S. Gavanji, B. Larki, A. Hosein, J. Novel Appl. Sci. 18 (2013) 650–654. (2004) 23–28 doi.org/10.1016/S0023-6438(03)00120-8.
[3] H. Ali, H. Nawaz, M. Saleem, F. Nurjis, M. Ahmed, J. Raman Spectrosc. 47 (2016) [44] A. Rohaman, Y.B. Che Man, Sci. World J. (2012) 2012, https://doi.org/10.1100/
706–711. 2012/250795.
[4] W. Dong, Y. Zhang, B. Zhang, X. Wang, J. Raman Spectrosc. 44 (2013) 1739–1745. [45] P. Lark, Infrared and Raman Spectroscopy: Principles and Spectral Interpretation,
[5] J.O. Horbańczuk, I. Malecki, R.G. Cooper, A. Jóźwik, J. Klewiec, J. Krzyźewski, Elsevier, 2011.

You might also like