Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Vibrational Spectroscopy 98 (2018) 128–133

Contents lists available at ScienceDirect

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

Vibrational and structural properties of L-Alanyl-L-Phenylalanine dipeptide T


by Raman spectroscopy, infrared and DFT calculations

C.B. Silvaa, , J.G. da Silva Filhoa, G.S. Pinheirob, A.M.R. Teixeirac, P.T.C. Freirea
a
Departamento de Física, Universidade Federal do Ceará, C.P. 6030, Campus do Pici, 60.455-760, Fortaleza, CE, Brazil
b
Departamento de Física, Universidade Federal do Piaui, campus Ministro Petrônio Portella, 64.049-550, Teresina, PI, Brazil
c
Departamento de Física, Universidade Regional do Cariri, 63.010-970, Juazeiro do Norte, CE, Brazil

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

Keywords: A study of vibrational and structural properties of L-Alanyl-L-Phenylalanine hydrophobic dipeptide (molecular
Dipeptide formula C12H16N2O3) is reported. The L-alanyl-L-phenylalanine is formed by two hydrophobic amino acids, L-
Raman spectroscopy alanine and L-phenylalanine, through a peptide bond. Individually these compounds have many important
FT-IR properties, including increasing immunity and providing energy to muscle tissue, brain and central nervous
DFT
system (L-alanine). We have performed measurements of Raman and Fourier transform infrared spectroscopy,
between 3500-50 cm−1 and 4000-130 cm−1, respectively, at ambient conditions. To support the experimental
results calculations using the density functional theory (DFT) with B3LYP functional, 6–31 G ++ (d, p) basis
sets and the polarizable continuum model of solvatation in an isolated molecule in the zwitterionic form were
done. The assignment of the normal modes was described by the potential energy distribution analysis. In this
way, it was possible to analyze the vibrational and structural properties of L-alanyl-L-phenylalanine, allowing
the assignment of each vibrational normal mode of the molecular structure, associating them with those results
obtained experimentally.

1. Introduction A series of works, using the most varied solvents, shows these di-
peptides tend to incorporate molecules of organic solvents into their
Short peptides have arisen the interest of the scientific community structure, which then act as acceptors for one of the three H atoms of
around the world. Although the peptides have less complex structures the N-terminal amino group NH3+ [14–18]. The change of solvent can
than the proteins, they exhibit important biological activities. result in the formation of different structures, for example, nanotubes,
Hydrogels formed by self-assembly of peptides have been used in 3D nanofilms, nanospheres, nanofibers, and materials with large channels
cell culture as constituents of compounds with regenerative properties filled with solvent [19–21]. These structures have been used in different
[1,2], which act in the regeneration of several tissues with low re- areas, such as in the construction of biosensors [21], nanocarriers for
generative capacity, such as bone tissue, dental tissue, skeletal tissue drug and gene delivery [22] and natural encapsulates [23,24]. In ad-
and cartilage [3–6]. Some peptides have antioxidant properties and act dition, the use of dipeptides has been tested in the fight against diseases
as pH regulator in muscle cells [7], avoiding the oxidation of organic such as zyka virus [25], cancer [26], diabetic osteoporosis [27], HIV
molecules and metals, helping in glycemic control [8] and the treat- [28], malaria [29], anti-tumor [30,31], and antibacterial and antifungal
ment of type 2 diabetes [9], besides being used in the diet of slimming agents [32].
[10]. In previously published works, crystalline structures with L-Ala-L-
The L-alanyl-L-phenylalanine (L-Ala-L-Phe) is a dipeptide formed by Phe dipeptide was obtained with 2-propanol solvent [33], ammonium
two hydrophobic amino acid residues. The hydrophobic dipeptides [34], hydrochloride dehydrate [35], as well as structures interacting
constitute a very diversified group of crystalline structure and have with copper [26] and gold [36], although the crystalline structure of
been a source of stable microporous materials [11,12], which are ori- pure L-Ala-L-Phe dipeptide is still unknown. Similar structures con-
ginated from molecular self-assembling dictated by the formation of taining the phenylalanine showed to be an effective hydrogelator, when
hydrogen bonds and by the aggregation of hydrophobic entities in the used with the specific solvent. In this manner the L-Ala-L-Phe is a good
side chains [13]. candidate to gel-forming [37–40].


Corresponding author.
E-mail address: cristiano.balbino@fisica.ufc.br (C.B. Silva).

https://doi.org/10.1016/j.vibspec.2018.08.001
Received 26 February 2018; Received in revised form 25 July 2018; Accepted 4 August 2018
Available online 09 August 2018
0924-2031/ © 2018 Elsevier B.V. All rights reserved.
C.B. Silva et al. Vibrational Spectroscopy 98 (2018) 128–133

Structural analysis and vibrational properties of dipeptides using the


density functional theory (DFT) [41,42] have been reported in the lit-
erature, being important to understanding the behavior of aggregation
of amino acid in complex structures as peptides and proteins [43–49].
The present work reports a detailed study about the structural and vi-
brational properties of L-Ala-L-Phe dipeptide (molecular formula
C12H16N2O3) including data about important torsion angles of the
molecular structure and the assignment of vibrational modes.

2. Experimental

L-Alanyl-L-Phenylalanine (L-Ala-L-Phe) dipeptide, which consists of


a white polycrystalline powder, was used as provided by Sigma-Aldrich.
The Raman spectra of L-Ala-L-Phe were performed using a triple spec-
trometer (Jobin-Yvon T64000) equipped with an N2-cooled charge-
coupled device detector (CCD). The 532 nm line of a semiconductor
laser was used as the exciting source, with an output power of 70 mW.
The spectral resolution was ∼ 2 cm−1. An Olympus microscope lens Fig. 1. Representation of molecular structure of L-Ala-L-Phe in zwitterionic
with focal distance 20.5 mm (N.A. = 0.35) was used to focus the laser form.
beam on the sample at room temperature. The spectral region analyzed
covered the region from 3500 to 40 cm−1.
modes of the Ala-Phe molecule.
The Fourier Transform infrared (FT-IR) measurement was obtained
through a Bruker Vertex 70v spectrometer. The infrared spectrum was
recorded from 4000 to 130 cm−1, using the Attenuated Total 4. Results and discussions
Reflectance (ATR) technique with vacuum by absorbance of radiation
to eliminate CO2 residues and humidity with a spectral resolution of The molecular structure of L-Ala-L-Phe (illustrated in Fig. 1) has 33
2 cm−1 and accumulating 128 scans per spectrum. atoms, molecular weight of 236.27 g/mol and molecular area of ap-
proximately 95.5 Å2. The data of positions of atomic coordinates and
3. Calculations the geometric parameters (bond lengths, bond angles and dihedral
angles) of the optimized molecular structure of L-Ala-L-Phe dipeptide
Calculation and normal mode assignment of IR and Raman spectra are listed in Tables S1 and S2 of Supplementary Material, respectively.
is widely utilized to understand organic materials like amino acids,
peptides and proteins [50–59]. The vibrational modes of L-Ala-L-Phe 4.1. Potential energy surface (PES)
dipeptide were calculated using density functional theory (DFT). For
this purpose, we used the Gaussian 09 package [60], with the B3LYP An important torsion angle in dipeptides deserves attention, due to
functional [61] and 6-31++G(d,p) basis set. Solvatation was also in- it considerably influences the conformation of peptides and proteins.
corporated for assignment by using the polarizable continuum model This torsion angle is formed by the alpha carbon (αC) and the beta
(PCM). Our calculations were established for an isolated molecule in carbon (βC) of the first amino acid residue with the alpha carbon and
the zwitterionic form. In general, zwitterionic states are predominantly the beta carbon of the second amino acid, interchanged by amide plane
observed in the crystalline medium for a variety of amino acids and (OCNH), θ = βC1- αC1•••αC2- βC2, which defines the relative positions of
dipeptides [62]. the two sides chain in the structure of dipeptide [19].
The initial molecular structure of L-Ala-L-Phe dipeptide used was This dihedral angle is the simplest way to describe the conforma-
that reported by Görbitz [33], and then optimized for a minimum of tional configuration of the molecule, characterizing the influence of the
energy. The optimized molecular structure was subjected to investigate inter- and intra-molecular interactions in the crystalline medium. In
the conformational structure by calculation of energy surface potential order to investigate the energetic stability of different conformational
(scan) of torsion dipeptide angle and, finally, to calculate the frequency configurations of L-Ala-L-Phe, the potential energy surface (PES) cal-
of vibrational modes for this molecule (3N-6). culation was carried out, varying the dihedral angle θ from -180° to
Using VEDA program [63] for molecular visualization and potential +180° degrees in steps of 10°. The relative energy with respect to the
energy distribution (PED) calculations, the vibrational assignments of global minimum (ΔE) was plotted against the dihedral angle, as shown
the normal modes were obtained. Here we consider only values with in Fig. 2. In this figure, we can see a global minimum at -110° and local
contributions of at least 10%. For a better correspondence between the minima around -150°, +20° and +150°. At room temperature the
theoretical and experimental spectra, we made use of the scale factors thermal energy (kT) is approximately 2.45 kJ/mol, which corresponds
suggested by the reference [64] of 0.977 for vibrations modes under to the dashed red line in the plot of Fig. 2. We can observe two regions
1800 cm−1, and 0.955 for vibrational modes above 1800 cm−1. above the thermal energy at room temperature, the first region between
The calculations were made in the conformer of the lowest energy. -85° to -115° and the second region between -177° to -130° separated by
Although the calculations had been performed in a molecule, the vi- a barrier of 1.96 kJ/mol.
brations of both single molecule and molecules in the crystal are si- In the polypeptide chain the definition of protein folding depends on
milar, exceptions occur with vibrations related to part of the molecule the torsion angles prior to alpha carbon, called phi angle (ϕ), and on the
involved in hydrogen bond, for example, and vibrations related to the torsion angle after the alpha carbon, called the psi angle (ψ) [65]. The
lattice modes that obviously do not appear in the calculation of the analysis of rotation of these angles led to the identification of permitted
molecule. It is important mentioning that the possible coincidence of regions where there is no clash between atoms, and non-permitted re-
the molecular conformation in the solid state with the molecular con- gions, where there is clash between the atoms. In Fig. 2 we also show
formation in the gas phase or in solution has supporters and opponents; the correspondence between each value of ϕ and ψ with dihedral angle
this question remains unresolved. As can be seen in Table 2, there is a θ. In this case, while the θ angle vary from −180° to + 180°, the ϕ
good agreement between the calculated wavenumbers and the experi- angle vary between −40° and 25° and the ψ angle vary between 80°
mental wavenumbers of the Raman and infrared bands to the internal and 110°, respectively.

129
C.B. Silva et al. Vibrational Spectroscopy 98 (2018) 128–133

Table 2
Calculated and scaled wavenumbers and experimental Raman and IR wave-
numbers (in units by centimeter) and classification of the vibrational modes of
L-Ala-L-Phe dipeptide with PED.
ωcalc ωscale ωRam ωIR Assignment of the molecular vibrations with PED
[%]

3638 3474 ν(N2H)[100]


3546 3386 3381 νas(N1H3+)[100]
3479 3322 3255 3256 νs(N1H3+)[98]
3205 3061 3062 3063 νs(CH-ring)[93]
3202 3058 3058 3058 νs(N1H3+)[96]
3193 3049 νas(CH-ring)[93]
3186 3043 3047 νas(CH-ring)[99]
3175 3032 3035 νas(CH-ring)[90]
3170 3027 3026 3025 νas(CH-ring)[95]
3151 3009 νs(C2H3)[99]
3145 3003 3003 3006 νas(C2H3)[96]
3133 2992 2981 νas(C5H2)[98]
3109 2969 2974 2971 ν(C1H)[95]
Fig. 2. Variation of energy and dihedral angle (ψ, ϕ) with the change in di-
3084 2945 2946 2955 νs(C5H2)[95]
hedral angle θ of L-Ala-L-Phe dipeptide. 3068 2929 2933 2829 ν(C4H)[94]
3066 2928 2920 2922 νs(C2H3)[100]
1709 1670 1673 1678 ν(C = O)[63] + ν(N2C3)[19]
Depending on the molecular interactions in the crystal structure, we
1670 1632 1631 δ(N1H3+)[85]
can observe different conformations at ambient temperature. This wide 1647 1609 1603 ν(C10=C11-ring)[30]
angular range can leave the side chains practically on opposite sides. 1637 1599 δ(N1H3+)[91]
This allows the formation of crystalline microporous structures in hy- 1625 1588 ν(C = C-ring)[41]
drophobic dipeptides, with hydrophilic internal superficies and hy- 1624 1587 1583 1565 νas(CO2−)[78]
1582 1546 1555 δ(C3N1H)[55]+ ν(N2C3)[21]
drophobic external superficies [13,66]. The majority of the known di- 1528 1493 1497 1497 δ(CCH-ring)[67]
peptides has the torsion dipeptide angle θ (in absolute values) higher 1505 1470 δ(C2H3)[89]
than 135° and other considerable portion between 90° and 135°, as 1492 1458 1457 1453 δ(C2H3)[77]
shown by Gorbitz [19]. The two ranges showed in PES of torsion di- 1490 1456 sc(C5H2)[53]
1481 1447 1444 1443 sc(C5H2)[27] + δ(R)(CCH-ring)[18]
peptide angle in L-Ala-L-Phe dipeptide are in accordance with the re-
1470 1436 wagg(N1H3+)[63] + wagg(C2H3)[13]
ported analysis. 1441 1409 1398 1389 wagg(C2H3)[81]
Another important effect on dipeptide structures is the solvent used 1398 1366 1369 γ(HC1C3N2)[62] + δ(C1C2H)[14]
in the synthesis process or the crystal growth. Many of the structures 1388 1356 1357 1356 δ (N2C4H)[24] + γ(HC4C12O3)[23] +
are established by a particular solvent; sometimes the change of it leads νs(CO2−)[11]
1364 1333 1349 νs(CO2−)[51]
obtaining different forms of the same dipeptide: nanotubes, hydrogels, 1359 1328 δ(CCH-ring)[37] + δ(C5N2H)[10]
nanolayers, vesicles, thin films, among others. For this reason, we 1353 1322 ν(CC-ring)[50] + δ(CCH-ring)[20]
submitted the molecular structure of L-Ala-L-Phe to optimize changing 1346 1315 1317 1318 δ(C2C1H)[35]
the solvent in solvatation. We chose some solvents available in Gaussian 1332 1301 1290 1289 wagg(C5H2)[33]
1297 1267 1250 1264 ν(N2C3)[26] + δ(C3N2H)[20] + δ(C1C2H)[18]
code that generally are used in the growth of some crystal structure of
1261 1232 1248 1231 γ(HN2C1C2)[45] + δ(HC4N2)[26]
dipeptides: 1-propanol, 2-propanol, methanol, ethanol, 2,2,2-tri- 1223 1195 1202 ν(C5C6)[33] + δ(HC7N8)[10]
fluoroethanol and water. 1199 1171 1182 1181 δ(HCC-ring)[39] + τ(C2H2)[12]
Table 1 shows the dielectric constant of each solvent and the cor- 1196 1168 1169 1167 δ(HCC-ring)[24] + τ(C2H2)[10]
respondent torsion dipeptide angle of the most stable conformation. 1176 1149 1155 δ(HCC-ring)[78]
1145 1119 1120 1124 r(C2H3)[17] + γ(HN1C1C2)[10] +
The two conformers of lowest energy were obtained using the methanol
γout(C2N1C3C1)[11]
and water as solvents. We can still see that torsion dipeptide angle 1105 1080 1081 1088 δ(HCC-ring)[32] + ν(CC)[25]
presents a small variation with respect to the change of dielectric 1087 1062 1047 1050 ν(C4C5)[22] + ν(N2C4)[20]
constant of solvent, increasing slightly as shown in Fig. 3. 1048 1024 1029 1028 ν(C8C9)[47] + δ(C9C10C11)[13] + δ(HC7C8)
[11]
1022 998 1012 r(C2H2)[39]
1013 990 r(N1H3+)[21] + γ(HC1C2N1)[18] + r(C2H2)[15]
4.2. Vibrational analysis
1012 989 989 δ(CCC)[62] + ν[C9C10)[12]
1004 981 γout(HCCC-ring)[67] + γout (C8C9C10C11)[16]
For the study of the vibrational properties (IR and Raman) the op- 1001 978 980 978 ν(C1C2)[23] + r(N1H3+)[17] + r(C2H2)[12]
timized structure was submitted to the frequency calculation, and the 987 964 γout(HCCC-ring)[78]
theoretical Raman spectrum was converted from Raman activity to 959 937 948 942 ν(C12C4)[19]+ γout(C12C5N2C4)[13]
932 911 918 917 γout(HCCC-ring)[65]
907 886 880 876 ν(N1C1)[20] + ν(C1C3)[16]
Table 1 868 848 854 853 ν(N2C4)[17] + ν(C4C5)[10]
Solvent effect in dihedral angle θ of L-Ala-L-Phe peptide. 859 839 844 844 γout(HCCC-ring)[96]
854 834 827 827 ν(N1C1)[26] + ν(C1C2)[15] + δ(O1C16N2)[13]
Solvent Dielectric constant (ε) Dihedral angle ΔE (kj/mol)
825 806 819 811 δ(C8C9C10)[24] + ν(C5C6)[16] + δ(CO2−)[13]
C2-C1—C4-C5 (°)
775 757 761 761 δ(CO2−)[23] + γout(O1C1N2C3)[15] +
γout(O2C4O3C12)[10]
2-propanol 19.264 −133.594 6.934
760 743 743 742 γout(HCCC-ring)[48] + γout(C6C7C8C9)[10]
1-propanol 20.524 −137.392 9.998
746 729 721 718 γout(O1C1N2C3)[30]+ δ(CO2−)[12]
2,2,2-trifluoroethanol 26.726 −138.191 10.859
+γout(O2C4O3C12)[11]
Ethanol 24.852 −138.199 7.701
710 694 698 696 γout(CCCC)[53] + γout(HCCC-ring)[41]
Methanol 32.613 −138.365 0
651 636 622 621 δ(CO2−)[14] + γout(O2C4O3C12)[14]
Acetonitrile 35.688 −138.389 4.365
632 617 δ(C9C10C11-ring)[75]
Water 78.355 −147.456 0.287
(continued on next page)

130
C.B. Silva et al. Vibrational Spectroscopy 98 (2018) 128–133

Table 2 (continued)

ωcalc ωscale ωRam ωIR Assignment of the molecular vibrations with PED
[%]

615 601 ν(C1C3)[20] + δ(O1C3N2)[12] + γout(O1C1N2C3)


[12]
581 568 566 564 δ(C6C7C8)[15] + δ(C4C12N2)[15] + δ(O3C4C12)
[11]
552 539 547 548 δ(O3C4C12)[27] + δ(C1C3N2)[10]
539 527 γ(HN2C3C1)[78]
489 478 493 471 γout(C2N1C3C1)[24] + γ(C7C8C9C10)[15]
444 434 δ(C1C2N1)[26] + δ(N2C4C5)[15] +
γout(N1C1C2C3)[12]
436 426 422 δ(N2C4C5)[28] + δ(O3C12C4)[18]
416 406 407 γout(CCCC-ring)[53]
387 378 398 395 γ(N1H3+)[37] + δ(O1C3N2)[17]
367 359 353 350 δ(C1C2N1)[11] + δ(N1C1C3)[10] +
γ(N1H3+)[10]
348 340 340 γ(CO2−)[13] + δ(C5C6C7)[11]
250 244 241 270 γ(C2H3)[33]
237 232 229 228 δ(C4C5C6)[14] + γ(C10C9C8C7)[12]
225 220 210 γ(C2H3)[18] + γout(C2N1C3C1)[16] +
γout(C12C9C8C7)[11]
202 197 γ(N1H3+)[66]
170 166 168 δ(C12C4N2)[20] + δ(C1C3N2)[14] + δ(N2C4C5)
[11]
138 135 133 δ(C3N2C4)[44] + δ(C1C3N2)[10]
102 100 γ(C3N2C4C5)[56]
83 81 85 δ(C4C5C6)[34] + γout(C5C7C11C6)[33]
66 64 γ(N1C1C3N2)[47]
61 60 γ(C1C3N2C4)[42] + γ(N1C1C3N2)[18] +
γ(N2C4C5C6)[11]
34 33 γ(C4C5C6C7)[54] + γ(C1C3N2C5)[16] +
γ(O3C12C4C5)[11]
33 32 γ(C4C5C6C7)[46] + γ(C1C3N2C5)[19]
22 21 γ(N2C4C5C6)[65] + γ(C1C3N2C5)[10]

Nomenclature: γ=torsion, γout = torsion out of plane, wagg = wagging,


δ=deformation, ν=stretching, νs=symmetric stretching, νas=anti-symmetric
stretching, r = rocking, τ=twisting.

Fig. 4. Experimental and calculated (scaled) infrared spectra above and ex-
perimental and calculated (scaled) Raman spectra below of L-Ala-L-Phe di-
peptide.

where calculated bands are not present in the experimental spectrum.


This is due the small difference in the frequency of the modes between
the calculated and the experimental values; however, these few cases do
not detract the quality of the assignment as a whole.
Table 2 presents a detailed description of the assignments of all
vibrational modes of the molecule of L-Ala-L-Phe dipeptide. In addition,
the table shows the calculated (ωcalc) and scaled (ωscale) wavenumbers,
as well as the IR and Raman experimental bands associated with the
relative assignment. Using the scaled wavenumber as a reference, we
can assign molecular vibrations for each normal modes observed in the
experimental measurements of Raman and IR spectroscopy. To facil-
itate the discussion, we separated the spectral ranges in four subsec-
tions.

Fig. 3. Evolution of torsion dipeptide angle of L-Ala-L-Phe with different sol-


4.2.1. Spectral region from 3500 to 2800 cm−1
vents.
The vibrations in this region are characteristic of stretching vibra-
tions of the functional groups NH, NH3+, CH, CH2 and CH3. Potential
Raman intensity using the procedure adopted in a previous work [67]. energy distribution presented a high contribution of these portions of
The Fig. 4 illustrates the comparison between theoretical and experi- the molecule in the normal modes (between 90–100%). A band ob-
mental Raman and infrared spectra. The Raman effect is due the in- served in the IR spectrum at 3386 cm−1 is attributed to an anti-sym-
elastic scattering of light, a process that involves the molecular polar- metric stretching of NH3+ group. Symmetric stretching of NH3+ was
izability, while infrared spectroscopy is based on effects of absorption situated at 3322 and 3058 cm−1 and was observed in both spectra (IR
of radiation through the electric dipole. Eventually, because the phy- and Raman). At 3061 cm−1, we have the symmetric stretching vibra-
sical effects are distinct, some modes in the Raman spectrum are not tion of CH belonging to the ring of the structure, while the anti-sym-
present in the IR spectrum, and some modes in the IR spectrum are not metric stretching can be observed between 3049 and 3027 cm−1. Anti-
observed in the Raman spectra. Additionally, when comparing the symmetric stretching of the CH3 group are observed at 3009 and
theoretical and experimental spectra one notes there are some few cases 3003 cm−1, and the symmetric stretching of the same group is assigned

131
C.B. Silva et al. Vibrational Spectroscopy 98 (2018) 128–133

at 3009 to 2928 cm−1. Symmetric stretching of the CH2 group is at- occurrence of phase transitions either under pressure [53] or under
tributed to the band at 2992 cm−1, while the CH2 anti-symmetric temperature [54] variations as previously reported. Admittedly, the
stretching corresponds to the band at 2945 cm−1. The bands at 2969 understanding of this spectral region helps us to realize the stability of
and 2929 cm-1 are due to the CH stretching vibration. Additionally, the the structure under diverse thermodynamics parameters.
CH/OH stretching region in the IR experimental spectrum contains a Finally, between 3200 and 1800 cm−1, the Raman and FT-IR
couple of features (in particular the one at 3200 cm−1) which can be spectra showed the absence of bands. This is characteristic of the ma-
understood as due the presence of water. jority of amino acids since the units that form the molecular structure
do not have vibrations with energy in this region. So, it is expected that
4.2.2. Spectral region from 1800 to 1200 cm−1 similar behavior occurs in dipeptides. This absence was reproduced by
In this region, the mixed modes begin to emerge due to vibrations of DFT calculations, showing a good agreement with experimental results.
different parts/groups of the molecule. At 1670 cm−1 we observed a Furthermore, assignments for other dipeptides discussed in previous
band attributed to C]O stretching plus N2C3 stretching, this band is works indicate a good agreement with the present work [43,44,69,70].
denominated amide I and appears as the most intense in the IR spec- As a conclusion of this section, we have furnished a complete descrip-
trum [68], where the latter corresponds to the peptide bond. Anti- tion of vibrational modes of the L-Ala-L-Phe dipeptide, giving in-
symmetric stretching of CO2- group occurs at 1587 cm−1. At 1546 cm formation about the normal modes of the material and linking theore-
−1
appears the band corresponding to the amide II, being a combina- tical calculations with experimental spectra.
tion of CN and NH vibrations [68].
Symmetric stretching is observed at 1357 and 1349 cm−1. 5. Conclusions
Deformation of NH3+ group is observed between 1632 and 1599 cm−1.
The C]C stretching vibrations in the ring of the structure is observed In this work, the structural and vibrational properties of the mole-
from 1609 to 1588 cm−1, and CCH deformations, between 1493 and cular structure of L-Ala-L-Phe dipeptide in the zwitterionic form were
1322 cm−1. The band situated at 1546 cm-1 is related to the C3N1H investigated using the density functional theory and vibrational spec-
deformation mixed with N2C3 stretching; these atoms correspond to troscopy. Analysis of important torsion angle in the molecular structure
the amide plane, being observed a very intense band in the IR spectrum, of dipeptides (θ=βC1αC1•••αC2- βC2) showed two regions susceptive to
whereas in the Raman spectrum it appears with low intensity. low energy conformation of L-Ala-L-Phe at room temperature. The first
Deformations of CH3 group occur between 1470 and 1447 cm−1. The region, found between 85° < |θ| < 115° and the second region, found
CH3 wagging vibrations are observed from 1436 to 1409 cm−1, where between 177 < |θ| < 130°, are separated by a barrier of 1.96 kJ/mol.
we can still observe an NH3+ wagging vibration. Wagging vibration of In these angular ranges, the side chains can carry practically on oppo-
CH2 occurs at 1301 cm−1 and the scissoring vibrational modes of the site sides, which may facilitate the formation of structures with hy-
CH2 group are observed at 1456 and 1447 cm−1. In-plane torsions of drophilic and hydrophobic properties as occurs with some dipeptides.
HCCN and HCCO and in-plane deformations of CCH and CNH are Conformers of lower energy were obtained for different solvents using
commons in this spectral region. We also observed NC and CC the CPMC method. The calculations showed that there is a small var-
stretching vibrations between 1267 and 1195 cm−1. iation of energy and the torsion dipeptide angle increases subtly with
the dielectric constant of the solvent. For L-Ala-L-Phe, conformations
4.2.3. Spectral region from 1200 to 600 cm−1 with lower energy were obtained using methanol and water as solvent.
In this spectral region predominates deformation and torsional vi- Regarding the vibrational properties of the material, the calculated
brations, and NC and CC stretchings are also observed. Deformations of wavenumbers reproduced the experimental results with good agree-
HCC belonging to the ring occur from 1171 to 1080 cm−1. The CH3 ment, allowing us to assign all normal modes of vibrations of L-Ala-L-
rocking vibration gives rise to the band at 1119 cm−1, while the CH2 Phe with the respective potential energy distribution.
rocking is observed at 998 and 990 cm−1, where we also observed the
NH3+ rocking vibration. Out-of-plane torsions of HCCC belonging to Acknowledgments
the ring are observed from 981 to 694 cm−1, where also occurs out-of-
plane torsion of CCCC, CCNC, OCNC, OCOC groups from the skeleton of The authors would like to acknowledge the financial support from
the molecular structure. Deformation of CO2- group contributes to the the Brazilian agencies CAPES and CNPq. We also thank the CENAPAD-
bands between 806 and 622 cm−1. Out-of-plane deformation of HCCC SP for the use of the GAUSSIAN 09 software package and the compu-
of the ring presents contributions in the bands from 981 to 698 cm−1. tational facilities through the project of reference proj373. PTCF ac-
Twisting vibration modes of the CH2 group contributes in the bands at knowledges support from FUNCAP-CNPq PRONEX PR2-011-00006.01-
1171 and 1168 cm−1. Rocking vibration modes (CH2, CH3 and NH3+) 00/15. J.G.S.F acknowledges the Brazilian agency CAPES for the
are present in the bands between 1119 and 990 cm−1. The CC postdoctoral program fellowship (process 1746352).
stretching has contribution in bands of the 1062 - 601 cm−1 region, and
the NC stretching presents contributions in bands located at 1047, 886, Appendix A. Supplementary data
848 and 827 cm−1.
Supplementary material related to this article can be found, in the
4.2.4. Spectral region from 600 to 150 cm−1 online version, at doi:https://doi.org/10.1016/j.vibspec.2018.08.001.
In this region predominates deformation and torsional vibrations of
the molecular skeletal of L-Ala-L-Phe dipeptide. The NH3+ torsional References
vibration is observed between 378 and 359 cm−1, the CO2- torsion
occurs at 340 cm−1 and the CH3 torsion at 220 cm−1. The low wave- [1] R.V. Ulijn, A.M. Smith, Chem. Soc. Rev. 37 (2008) 664.
number region (lower than 150 cm−1) is associated with lattice modes, [2] S. Zhang, F. Gelain, X. Zhao, Semin. Cancer Biol. 15 (2005) 413.
[3] T.D. Sargeant, S.M. Oppenheimer, D.C. Dunand, S.I. Stupp, J. Tissue Eng. Regen.
so the calculation in a single molecule is not precise for this region. In Med. 2 (2008) 455.
Table 2 we furnish a tentative assignment of the modes with wave- [4] R.N. Shah, N.A. Shah, M.M. Del Rosario Lim, C. Hsieh, G. Nuber, S.I. Stupp, Proc.
number lower than 150 cm−1 as obtained from the DFT calculation for Natl. Acad. Sci. 107 (2010) 3293.
[5] F.F. Demarco, M.C. Conde, B.N. Cavalcanti, L. Casagrande, V.T. Sakai, J.E. Nor,
one molecule. This indicates that although most of the bands in this Braz. Dent. J. 22 (2011) 3.
region are associated with lattice modes, possibly they are coupled to [6] F. Gelain, D. Bottai, A. Vescovi, S. Zhang, Designer self-assembling, PLoS One 1
some internal modes. However, the most important point is the possi- (2006) 413.
[7] H. Fu, Y. Katsumura, M. Lin, Y. Muroya, K. Hata, K. Fujii, A. Yokoya, Y. Hatano,
bility to use the bands appearing in this region to define or not the

132
C.B. Silva et al. Vibrational Spectroscopy 98 (2018) 128–133

Radiat. Phys. Chem. 78 (2009) 1192. [44] S. Celik, A.E. Ozel, S. Kecel, S. Akyuz, Vib. Spectrosc. 61 (2012) 54.
[8] Y. Dinçer, Z. Alademir, H. Ilkova, T. Akçay, Clin. Biochem. 35 (2002) 297. [45] P. Leyton, C. Paipa, A. Berrios, A. Zárate, M.V. Castillo, S.A. Brandán, J. Mol. Struct.
[9] Y. Dinçer, T. Akçay, Z. Alademir, H. Ilkova, Mutat. Res. 505 (2002) 75. 1031 (2013) 110.
[10] A.D. Mooradian, M. Smith, M. Tokuda, Clin. Nutr. ESPEN 18 (2017) 1. [46] M.A. Ziganshin, I.G. Efimova, V.V. Gorbatchuk, S.A. Ziganshina, A.P. Chuklanov,
[11] V.N. Yadav, A. Comotti, P. Sozzani, S. Bracco, T. Bonge-Hansen, M. Hennum, A.A. Bukharaev, D.V. Soldatov, J. Pept. Sci. 18 (2012) 209.
C.H. Görbitz, Angew. Chemie - Int. Ed. 54 (2015) 15684. [47] L. Zheng, Y. Zhao, H. Dong, G. Su, M. Zhao, J. Funct. Foods 21 (2016) 485.
[12] C.H. Görbitz, M. Nilsen, K. Szeto, L.W. Tangen, Chem. Commun. (2005) 4288. [48] S. Mondal, D.S. Chowdhuri, S. Ghosh, A. Misra, S. Dalai, J. Mol. Struct.
[13] C.H. Görbitz, Chem. - A Eur. J. 13 (2007) 1022. THEOCHEM. 810 (2007) 81.
[14] C.H. Görbitz, E. Gundersen, Acta Crystallogr. Sect. C Cryst. Struct. Commun. 52 [49] S. Ghosh, S. Mondal, A. Misra, S. Dalai, J. Mol. Struct. THEOCHEM. 805 (2007) 133.
(1996) 1764. [50] P. Lagant, G. Vergoten, G. Fleury, M.H.L. Lefebvre, Eur. J. Biochem. 139 (1984)
[15] T. Andrade-Filho, T.C. Martins, F.F. Ferreira, W.A. Alves, A.R. Rocha, Theor. Chem. 137.
Acc. 135 (2016) 1. [51] D.L. Rousseau, R.P. Bauman, S.P.S. Porto, J. Raman Spectrosc. 10 (1981) 253.
[16] C.H. Görbitz, Acta Crystallogr. Sect. E Crystallogr. Commun. 72 (2016) 635. [52] T.C. Cheam, J. Mol. Struct. 295 (1993) 259.
[17] C.H. Görbitz, K. Rissanen, A. Valkonen, Å. Husabø, Acta Crystallogr. Sect. C Cryst. [53] P. Vennila, M. Govindaraju, G. Venkatesh, C. Kamal, J. Mol. Struct. 1111 (2016)
Struct. Commun. 65 (2009) 267. 151.
[18] D.M. M, S.P. S, M. a Castanho, N.C. Santos, J. Pept. Sci. 14 (2008) 1084. [54] N. Sundaraganesan, S. Ilakiamani, H. Saleem, P.M. Wojciechowski, D. Michalska,
[19] C.H. Görbitz, Chem. - A Eur. J. 7 (2001) 5153. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 61 (2005) 2995.
[20] A.C. Mendes, E.T. Baran, R.L. Reis, H.S. Azevedo, Rev. Nanomed. Nanobiotechnol. 5 [55] V.M. Naik, S. Krimm, Biophys. J. 49 (1986) 1147.
(2013) 582. [56] A. Arbor, 45 (1989) 271.
[21] Y. Gong, X. Chen, Y. Lu, W. Yang, Biosens. Bioelectron. 66 (2015) 392. [57] L.D. Bannon, A.R. Gargaro, L. Hecht, Spectrochim. Acta 47A (1991) 1001.
[22] R.F. Silva, D.R. Araujo, E.R. Silva, R.A. Ando, W.A. Alves, L - Langmuir 29 (2013) [58] P.K. Bose, P.L. Polavarapu, L.D. Barron, L. Hecht, J. Phys. Chem. 94 (1990) 1734.
10205. [59] Y. Imai, K. Aida, J. Inorg, Nucl. Chem. 41 (1979) 963.
[23] Y. Okahata, K. Ljiro, J. Chem. Soc. Perkin Trans. I 11 (1988). [60] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
[24] T.D. Do, M.T. Bowers, Anal. Chem. 87 (2015) 4245. G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
[25] Y. Li, Z. Zhang, W.W. Phoo, Y.R. Loh, W. Wang, S. Liu, M.W. Chen, A.W. Hung, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada,
T.H. Keller, D. Luo, C.B. Kang, Structure. 25 (2017) 1242. M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
[26] S. Iglesias, N. Alvarez, M.H. Torre, E. Kremer, J. Ellena, R.R. Ribeiro, R.P. Barroso, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, J.E. Peralta, F. Ogliaro,
A.J. Costa-filho, M.G. Kramer, G. Facchin, J. Inorg. Biochem. 139 (2014) 117. M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi,
[27] L. Baerts, L. Glorie, W. Maho, A. Eelen, A. Verhulst, P. D’Haese, A. Covaci, I. De J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi,
Meester, Pharmacol. Res. 100 (2015) 336. M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken,
[28] T. Ju, D. Hu, S.H. Xiang, J. Guo, Bioorg. Chem. 68 (2016) 105. C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin,
[29] L. Pérez-Picaso, H.F. Olivo, R. Argotte-Ramos, M. Rodríguez-Gutiérrez, M.Y. Rios, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski,
Bioorg. Med. Chem. Lett. 22 (2012) 7048. G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas,
[30] S. Celik, A.E. Ozel, S. Akyuz, Vib. Spectrosc. 83 (2016) 57. J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian09, Gaussian 09, (2009).
[31] B. Xu, W.Q. Yan, X. Xu, G.R. Wu, C.Z. Zhang, Y.T. Han, F.H. Chu, R. Zhao, [61] M.P. Andersson, P. Uvdal, J. Phys. Chem. A 109 (2005) 2937.
P.L. Wang, H.M. Lei, Eur. J. Med. Chem. 130 (2017) 26. [62] G. Wu, Amino acids: Biochem. Nutr. (2010).
[32] M. Bajaj, S.K. Pandey, T. Nain, S.K. Brar, P. Singh, S. Singh, N. Wangoo, [63] M.H. Jamróz, Vibrational Energy Distribution Analysis VEDA 4, Warsaw (2010).
R.K. Sharma, Colloids Surf. B Biointerfaces 158 (2017) 397. [64] C.W. Bauschlicher, S.R. Langhoff, Spectrochim. Acta A. Mol. Biomol. Spectrosc. 53
[33] C.H. Görbitz, Acta Crystallogr. Sect. C C55 (1999) 2171. (1997) 1225.
[34] C.H. Görbitz, Acta Crystallogr. Sect. E Struct. Reports Online 60 (2004) 2065. [65] J. Heller, D.D. Laws, M. Tomaselli, D.S. King, D.E. Wemmer, A. Pines, R.H. Havlin,
[35] G.F. Pauli, P. Junior, Science 38 (1995) 2–7. E. Oldfield, J. Am. Chem. Soc. 119 (1997) 7827.
[36] B.B. Koleva, T. Kolev, S.Y. Zareva, M. Spiteller, J. Mol. Struct. 831 (2007) 165. [66] C.H. Görbitz, New J. Chem. 27 (2003) 1789.
[37] A. Mahler, M. Reches, M. Rechter, S. Cohen, E. Gazit, Adv. Mater. 18 (2006) 1365. [67] A.N.L. Marques, J. Mendes Filho, P.T.C. Freire, H.S. Santos, M.R.J.R. Albuquerque,
[38] D.J. Adams, L.M. Mullen, M. Berta, L. Chen, W.J. Frith, Soft Matter 6 (2010) 1971. P.N. Bandeira, R.V. Leite, R. Braz-Filho, G.O.M. Gusmão, C.E.S. Nogueira,
[39] V. Jayawarna, M. Ali, T.A. Jowitt, A.F. Miller, A. Saiani, J.E. Gough, R.V. Ulijn, Adv. A.M.R. Teixeira, J. Mol. Struct. 1130 (2017) 231.
Mater. 18 (2006) 611. [68] R. Schweitzer-Stenner, Vib. Spectrosc. 42 (2006) 98.
[40] L. Chen, S. Revel, K. Morris, D.J. Adams, Chem. Commun. 46 (2010) 4267. [69] J.G. da Silva Filho, J.M. Filho, F.E.A. Melo, J.A. Lima, P.T.C. Freire, Vib. Spectrosc.
[41] W. Kohn, A.D. Becke, R.G. Parr, J. Phys. Chem. 100 (1996) 12974. 92 (2017) 173.
[42] M.W. Wong, Chem. Phys. Lett. 256 (1996) 391. [70] J.G. Silva, L.M. Arruda, G.S. Pinheiro, C.L. Lima, F.E.A. Melo, A.P. Ayala, J.M. Filho,
[43] S. Kecel, A.E. Ozel, S. Akyuz, S. Celik, G. Agaeva, J. Mol. Struct. 993 (2011) 349. P.T.C. Freire, Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 148 (2015) 244.

133

You might also like