This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
1. Introduction
Tuftsin is a naturally occurring linear tetrapeptide with amino acid sequence of threonine- (Thr-) lysine- (Lys-) proline- (Pro-) arginine (Arg-), and it was first identified by Nishioka et al. [1]. Tuftsin corresponds to residues 289-292 of leukokinin, a cytophilic γ-globulin, and it has been considered to be responsible for the activity in the stimulation of phagocytosis (mechanism of removing pathogens) [2], and any modification in the sequence results in the loss of activity. Tuftsin has been found to exhibit several biological activities connected with immune system function [3], and its deficiency is involved in severe infections in skin, lymph nodes, and lungs [4]. In addition, tuftsin shows antitumor activity [5], and it can be used as an activator of macrophages in cancer treatment [6], and it is supposed that tuftsin deficiency leads to the risk of bacterial infection in HIV-positive patients [7]. Apart from the above natal significance, because of the broad spectrum of the other biological activities such as mobility of granulocytes, humoral antibody formation, and tumoricidal activity of phagocytic cells as well as bacterial killing properties [8], tuftsin has been considered as the subject of several theoretical and experimental studies with a view to analyze its structure-activity relationships.
Siemion et al. [9] performed 13C–NMR measurements for the tuftsin and its analogs in D2O to understand the role of proline residue in the creation of the tetrapeptides. The same research group [10] performed circular dichroism measurements with tuftsin and its few sequential isomers and showed that the tuftsin’s lowest tendency to form turn. Tzehoval et al. [11] tested tuftsin’s capability to augment phagocytosis and its function as an immunogenic stimulator. Perkowska et al. [8] investigated the antibacterial properties of tuftsin and its analogs on twenty bacterial strains and found that tuftsin possesses the strongest antibacterial activity of all the analogs tested. Fridkin and Gottlieb [3] analyzed the structure-function relationships of tuftsin and found a direct correlation between tuftsin levels in the human blood serum of normal as well as various pathological origin and susceptibility to bacterial infections. Najjar [4] discussed the clinical and physiological aspects of tuftsin deficiency syndromes and stated that the therapy for these syndromes has been limited to gamma globulin injection along with appropriate chemotherapy. Fitzwater et al. [12] studied the conformational energy space of tuftsin and derived the characteristics of the molecular structure of tuftsin from the group of computed minimum energy conformations. Blumenstein and coworkers [13] studied the solution conformation of tuftsin using the NMR spectroscopy and did not find any preferred conformation of tuftsin.
Nishioka et al. [14] synthesized tuftsin using the Merrifield solid-phase method and noted that the synthetic tuftsin has the same physical, chemical, and biological properties as that of its natural compound. Ursi and his coworkers [15] have studied tuftsin in water and dimethyl sulfoxide solutions using 2D NMR spectroscopy. Sekacis et al. [16] found the tendency of β-formation of tuftsin in dimethyl sulfoxide solution. O’Connor and others [17] employed high temperature quenched molecular dynamics to search the conformational space of tuftsin in dimethyl sulfoxide and water. Siemion et al. [18] studied the 13C–NMR and circular dichroic spectra of tuftsin in connection with their hypothesis that β-turn is its biologically active conformation and concluded that Lys-Pro residues are important for the tuftsin conformation.
Nikiforovich [19] proposed a quasi-cyclic conformation of tuftsin with
2. Computational Details
The density functional theory (DFT) with Becke’s three parameter exact exchange functional (B3) [36] combined with gradient-corrected Lee-Yang-Parr correlation functional (LYP) [37] has been employed to optimize the tuftsin and its retro form using 6-31G
3. Results and Discussion
3.1. Structure and Energy
For convenience, the mono hydrated tuftsin and retro tuftsin complexes at four amino acid residues Thr, Lys, Pro, and Arg are represented as TTW, TLW, TPW, and TAW and RTTW, RTLW, RTPW, and RTAW, respectively. The solvated tuftsin complexes with four water molecules are represented as Tuftsin…4W and Retro tuftsin…4W and are depicted in Figures 1 and 2, respectively, along with the mono hydrated complexes where the atom numbering convention is detailed and the values mentioned are in Å. According to Siemion and Konopinska [42], the tuftsin was considered to be sensitive to structural changes, and it is substantiated by the energy of isolated tuftsin which is 8.16 and 4.39 kcal/mol lower than its retro form at MP2 and B3LYP levels of theory, respectively. The optimized structural parameters (selected values) of the bare and hydrated tuftsin are given in Table 1, and the structural parameters for retro tuftsin are listed in Table 2. The conformational changes are revealed through representative dihedral angles of tuftsin, and its hydrated complexes calculated at B3LYP/6-31G
[figure(s) omitted; refer to PDF]
Table 1
Selected geometrical parameters (bond lengths (in Å), bond angles (in degree)) of tuftsin and its hydrated complexes calculated at B3LYP/6-31G
| Parameter | Tuftsin | TTW | TLW | TPW | TAW | Tuftsin…4W |
| C56-N55 | 1.352 | 1.347 | 1.346 | 1.352 | 1.352 | 1.341 |
| N55-C40 | 1.460 | 1.462 | 1.454 | 1.460 | 1.459 | 1.458 |
| C40-C38 | 1.543 | 1.542 | 1.542 | 1.543 | 1.542 | 1.541 |
| N3-C2 | 1.468 | 1.471 | 1.470 | 1.466 | 1.468 | 1.470 |
| C2-C9 | 1.542 | 1.542 | 1.544 | 1.542 | 1.542 | 1.544 |
| C9-N11 | 1.367 | 1.369 | 1.364 | 1.355 | 1.367 | 1.355 |
| N11-C13 | 1.453 | 1.454 | 1.455 | 1.458 | 1.459 | 1.459 |
| C13-C14 | 1.549 | 1.547 | 1.550 | 1.551 | 1.540 | 1.550 |
| C66-C63-C58 | 113.63 | 113.70 | 112.57 | 113.68 | 113.99 | 113.36 |
| C63-C58-C56 | 111.46 | 111.22 | 109.67 | 111.47 | 112.37 | 110.34 |
| C58-C56-O57 | 123.15 | 122.13 | 121.66 | 123.19 | 123.52 | 122.75 |
| C58-C56-N55 | 113.77 | 114.17 | 116.46 | 113.77 | 113.61 | 114.50 |
| C56-N55-C40 | 121.55 | 122.11 | 119.96 | 121.42 | 121.11 | 122.31 |
| N55-C40-C38 | 109.54 | 108.15 | 109.88 | 109.51 | 109.38 | 109.19 |
| C40-C38-N3 | 117.79 | 118.17 | 118.40 | 117.67 | 117.84 | 118.26 |
| C38-N3-C2 | 119.22 | 118.86 | 119.91 | 119.36 | 119.44 | 119.46 |
| N3-C2-C9 | 114.92 | 114.66 | 116.14 | 115.03 | 115.28 | 115.35 |
| C2-C9-N11 | 116.94 | 116.98 | 118.01 | 117.11 | 117.23 | 118.34 |
| C9-N11-C13 | 121.04 | 120.91 | 120.21 | 123.86 | 120.98 | 122.57 |
| N11-C13-C14 | 110.53 | 110.42 | 110.31 | 109.88 | 110.81 | 109.57 |
| C13-C14-C16 | 116.11 | 115.82 | 115.95 | 116.04 | 116.29 | 115.77 |
| C14-C16-C20 | 113.73 | 114.13 | 113.01 | 113.84 | 116.25 | 114.00 |
| C16-C20-N28 | 110.49 | 110.22 | 111.01 | 110.38 | 109.38 | 110.19 |
| C20-N28-C29 | 118.36 | 118.70 | 118.09 | 118.53 | 118.10 | 118.89 |
Table 2
Selected geometrical parameters (bond lengths (in Å), bond angles (in degree)) of retro tuftsin and its hydrated complexes calculated at B3LYP/6-31G
| Parameter | Retro tuftsin | RTTW | RTLW | RTPW | RTAW | Retro tuftsin…4W |
| C29-N28 | 1.281 | 1.281 | 1.280 | 1.281 | 1.286 | 1.282 |
| N28-C20 | 1.452 | 1.447 | 1.452 | 1.452 | 1.451 | 1.447 |
| C20-C16 | 1.535 | 1.541 | 1.534 | 1.536 | 1.535 | 1.540 |
| C14-C13 | 1.550 | 1.553 | 1.554 | 1.548 | 1.551 | 1.549 |
| N11-C9 | 1.368 | 1.368 | 1.351 | 1.362 | 1.369 | 1.355 |
| C9-C2 | 1.543 | 1.543 | 1.537 | 1.548 | 1.547 | 1.543 |
| C1-C2 | 1.542 | 1.542 | 1.547 | 1.539 | 1.540 | 1.542 |
| C42-C43 | 1.540 | 1.541 | 1.540 | 1.540 | 1.545 | 1.543 |
| C43-C44 | 1.541 | 1.541 | 1.541 | 1.542 | 1.538 | 1.537 |
| C29-N28-C20 | 118.73 | 121.45 | 118.87 | 118.70 | 117.45 | 121.06 |
| N28-C20-C16 | 110.11 | 109.90 | 110.22 | 110.11 | 110.58 | 109.79 |
| C20-C16-C14 | 114.02 | 114.80 | 113.89 | 113.91 | 113.95 | 115.25 |
| C16-C14-C13 | 116.13 | 116.28 | 115.99 | 115.47 | 116.96 | 115.84 |
| C13-C15-O18 | 112.31 | 113.25 | 112.29 | 112.21 | 112.65 | 113.20 |
| C14-C15-O17 | 125.05 | 123.07 | 124.98 | 125.89 | 124.96 | 124.06 |
| C14-C13-N11 | 110.91 | 110.98 | 110.93 | 109.98 | 110.46 | 109.36 |
| C13-N11-C9 | 122.14 | 122.15 | 123.07 | 120.69 | 122.50 | 120.71 |
| N11-C9-C2 | 115.78 | 115.78 | 116.50 | 117.74 | 115.77 | 118.20 |
| C9-C2-N3 | 113.48 | 113.42 | 113.35 | 115.87 | 113.06 | 115.28 |
| C2-N3-C38 | 119.92 | 119.96 | 119.68 | 121.38 | 119.71 | 121.29 |
| N3-C38-C40 | 116.10 | 116.06 | 116.41 | 116.83 | 115.92 | 116.50 |
| C38-C40-C41 | 112.75 | 112.93 | 112.45 | 112.89 | 113.73 | 112.81 |
| C41-C42-C43 | 114.46 | 114.30 | 114.50 | 114.54 | 113.32 | 114.42 |
| C43-C44-N45 | 116.05 | 116.27 | 116.08 | 116.03 | 114.98 | 115.00 |
Table 3
The representative dihedral angles of tuftsin and its hydrated complexes calculated at B3LYP/6-31G
| Dihedral angles | Tuftsin | TTW | TLW | TPW | TAW | Tuftsin…4W |
| ΨThr(N60C58C56N55) | 39.27 | 37.65 | 43.58 | 39.23 | 43.44 | 30.05 |
| ωThr(C58C56N55C40) | -177.74 | -171.97 | 178.61 | -178.18 | -177.10 | -176.61 |
| χ1Thr(N60C58C56O57) | -142.12 | -142.98 | -136.96 | -142.26 | -137.53 | -150.55 |
| ΦLys(C56N55C40C38) | 70.29 | 78.69 | 57.21 | 71.72 | 66.23 | 62.70 |
| ΨLys(N55C40C38N3) | -118.63 | -131.17 | -127.66 | -117.79 | -123.34 | -135.44 |
| ωLys(C40C38N3C2) | -179.12 | -179.31 | -176.56 | -178.65 | -177.64 | -176.77 |
| χ1Lys(N55C40C41C42) | -58.99 | -60.50 | -70.22 | -59.48 | -60.71 | -63.06 |
| χ2Lys(C40C41C42C43) | -176.34 | -176.43 | 106.41 | -177.92 | -178.41 | 179.91 |
| ΦPro(C38N3C2C9) | -74.36 | -76.82 | -74.24 | -77.08 | -73.03 | -82.89 |
| ΨPro(N3C2C9N11) | -13.77 | -9.81 | -8.62 | -10.16 | -10.65 | 4.22 |
| ωPro(C2C9N11C13) | 171.98 | 168.33 | 176.45 | 172.78 | 173.58 | -178.46 |
| χ1Pro(N3C2C1C5) | 31.71 | 29.32 | 34.13 | 32.11 | 32.21 | 32.57 |
| χ2Pro(C2C1C5C4) | -37.60 | -38.16 | -38.05 | -37.80 | -37.76 | -38.22 |
| ΦArg(C9N11C13C15) | -99.29 | -90.96 | -92.07 | -103.79 | -96.85 | -95.68 |
| χ1Arg(N11C13C14C16) | 173.69 | 175.54 | 170.17 | 174.30 | -173.89 | 179.15 |
| χ2Arg(C13C14C16C20) | -94.02 | -93.54 | -95.38 | -94.14 | -75.62 | -92.33 |
[figure(s) omitted; refer to PDF]
Compared to tuftsin, all the C-N bonds of Arg are observed to be shortened in retro tuftsin except C29-N30 and N28-C20. Only marginal deviations are noted for Arg C-C and C=O bonds in the tuftsin complexes. All the dihedral angles show minor variations, and a notable difference in the dihedral angle associated with Cα-C (χ1Arg) is spotted in TAW compared to its parent molecule. The Cα-N of Lys is found to be stretched while the C-N bond associated with the side chain is shortened in retro tuftsin compared to tuftsin. No much difference is detected for Cα-C and C=O bonds. Larger variation is noted in the dihedral angle associated with the Cβ-Cγ of the Lys side chain (χ2Lys) of TLW where the water molecule makes one O-H…O H-bond with C=O of Lys (1.877 Å), another O-H…O H-bond with carboxylic C=O of Arg (1.982 Å) and one N-H…O H-bond (1.943 Å) with Thr residues which resembles to the results pertaining to H-bonds obtained in our previous studies [45, 46]. Being the end residue, there is no much variation in Thr conformations between tuftsin and its retro form.
While considering Pro residue, being at the third position of the peptide chain of tuftsin, it plays an important role in the stabilization of the biologically active conformation of tuftsin [42, 47]. Based on Siemion’s approach, the
Earlier studies suggest that one probable tuftsin conformation involves a β-turn with H-bond located between Thr and Arg [50] as is observed in the present study. Three types of intramolecular H-bonds between Thr and Arg(O57(Thr)..H12-N11(Arg), H61(Thr)..O17-C15(Arg), and O64(Thr)..H33-N30(Arg)) are observed in tuftsin and its mono hydrated complexes within the distances ranging from 2.07 to 2.49 Å, 2.17 to 2.41 Å, and 2.14 to 2.22 Å, respectively, which is shown in Figure 4. This yet again validates the β-turn conformation of tuftsin in its present form which is its biologically active conformation [18] of the peptide chain as previously reported by Konopinska et al. [50] and Sobczyk et al. [51] and supported by Kahn and Devens [52]. But only O57 (Thr)..H12-N11(Arg) H-bond is observed at 2.35 Å in Tuftsin…4W complex, and the β-turn structure is observed to be slightly disturbed with
[figure(s) omitted; refer to PDF]
The Arg 1 ← Lys 3 (C15-O17…H53) type intramolecular H-bond exists in isolated and hydrated retro tuftsin complexes (Figure 5) with distances varying from 2.196 to 2.396 Å except RTLW and Retro tuftsin…4W complexes. In RTLW, the above intramolecular H-bond is replaced by two intermolecular H-bonds O76-H77(W)…O17(Arg) and O76-H78(W)…O39 (Lys) at 1.917 and 1.947 Å, respectively, as a result of addition of water molecule which acts as a channel between Arg and Lys residues. This causes an increase in the Arg 1 ← Lys 3 intramolecular distance in RTLW up to 4.978 Å. In Retro tuftsin…4W complex, this distance is found to be 5.066 Å. In this complex also, the Arg 1 ← Lys 3 H-bond is replaced by two intermolecular H-bonds O76-H77(W)…O17 (Arg) and O76-H78(W)…O39 (Lys) at 1.958 and 1.927 Å, respectively.
[figure(s) omitted; refer to PDF]
Another intramolecular H-bond Lys 3 ← Thr 4 (C55-O56…H75) is also traced in the retro tuftsin complexes (Figure 5). The most effective intramolecular interaction takes place on O56…H75 in RTLW (2.244 Å). In RTTW, the Lys 3 ← Thr 4 H-bond is increased equal to 3.221 Å (from 2.318 Å in isolated retro tuftsin) due to the interaction of water molecule at O56 of Thr which acts as protonic acceptor from water molecule and increases the intramolecular distance. It is appealing to note that in the Retro tuftsin…4W complex, due to the insertion of water molecule in between Lys and Thr, the Lys 3 ← Thr 4 H-bond is replaced by a N54-H75 (Lys)…O85(W) and O85-H86(W)…O56 H-bonds at 2.076 and 1.82 Å, respectively. It is worth mentioning that an intraresidual N-H…N H-bond is traced at Thr in isolated and hydrated retro tuftsin complexes with distances ranging from 2.079 to 2.149 Å.
A weak H-bond between Lys 2 CO (C38-O39) and Arg 4 NH (N11-H12) is noted (3.007 Å) as is observed in the previous study by Kothekar et al. [23]. Except ΦPro (present -74.36° and previous -79.0°), all the other dihedral angles differ from the values reported by Ursi and coworkers [15] which illustrates that the present tuftsin conformation conflicts with the previously reported structure. Structural proposals of tuftsin made by different authors differ considerably even for studies performed in the same environment [15], but globally, the tuftsin conformation is referred to as “hairpins with split ends” as observed in the present study (Figure 1) and supported by Fitzwater et al. [12] in addition to Scheraga and coworkers [12] as one of the low energy conformations of tuftsin. The Cα (Thr)-Cα (Arg) distance of tuftsin in this present conformation is 5.09 Å (should be ≤7.0 Å for the structures to be in their trans conformation [15] corroborate its trans configuration). This interatomic distance between the alpha carbons of Thr and Arg in the hydrated tuftsin complexes is observed to vary between 5.11 and 5.67 Å and shows that the hydration does not alter the trans structure of tuftsin tetrapeptide.
It is to be noted that TLW in which the water molecule bridges Thr, Lys, and Arg of tuftsin is the lowest energy complex (interaction energy is -13.99 and -13.05 kcal/mol at B3LYP and MP2 levels of theory, respectively) among the complexes formed from the binding of a water molecule at oxygen sites of four amino acid residues of tuftsin. In this structure, Lys C=O and Arg carboxylic C=O are acting as proton acceptor, and Thr N-H is acting as a proton donor with H-bond distances of 1.877, 1.982, and 1.943 Å, respectively. It is also interesting to examine the relative stability of monohydrated tuftsin and retro tuftsin complexes as a function of interaction energy calculated at B3LYP/6-31G
[figure(s) omitted; refer to PDF]
With -6.65 kcal/mol interaction energy, TTW stands as the third member in the stability order. The water molecule in TTW is involved in bidentate mode forming two H-bonds, one with Thr oxygen and the other with Cδ-H hydrogen of Pro residues. The TPW is found to be the least stable mono solvated tuftsin complex with relative energy difference of 10.79 kcal/mol calculated at B3LYP level of theory while the TTW is found to be the least stable complex at MP2 level of theory with a relative energy of 12.93 kcal/mol. In TPW, the water molecule is attached between Pro C=O oxygen and Arg Cα-H hydrogen with H-bond distances of 1.885 and 2.308 Å, respectively. In Tuftsin…4W complex, the four binding water molecules retain similar interactions with the four amino acid residues as in their individual mono hydrated complexes with marginally longer H-bonds. Same trend has been observed in the formation of C=O oxygen and Cα-H hydrogen bonds previously [46]. The interaction energy difference between the Tuftsin…4W and the most stable mono hydrated (TLW) complexes is found to be 23.97 and 22.91 kcal/mol at B3LYP and MP2 levels of theory, respectively. This difference is obviously due to the added number of H-bonds between the tuftsin tetra peptide and water molecules. Ten strong H-bonds (two N-H…O, two C-H…O, and six O-H…O types) are traced in this four water complex with H-bond lengths varying from 1.858 Å to 2.316 Å.
In monohydrated retro tuftsin complexes, the most effective interaction takes place on the Arg carboxylic oxygen (O18-H19) in RTAW, the most stable monohydrate, resulting in the strengthening of O-H…O bonds (1.783 Å) and higher interaction energy (-14.24 kcal/mol). In RTAW, the binding water is acting as protonic acceptor and donor simultaneously and forming a closed interaction at Arg–COOH with H-bond lengths 1.783 and 1.881 Å, respectively. In the Retro tuftsin…4W complex, still stronger interactions persist at the same carboxylic oxygen of Arg (1.765 Å) along with additional eleven H-bonds (1.82-2.376 Å) render almost three times higher interaction energy (-41.85 kcal/mol) for the above complex than the most stable mono hydrated retro tuftsin complex. At all applied levels, the stability order for hydrated retro tuftsin complexes based on relative and interaction energies is found to be the same, and it is predicted to be RTAW>RTTW>RTLW>RTPW. An energy difference of -1.51 kcal/mol is noted between the most and the next stable hydrated retro tuftsin complexes exactly as that of their tuftsin counterparts (-1.5 kcal/mol).
It is interesting to note down that irrespective of the position in the peptide chain, monohydrated tuftsin complexes at Pro residue stand last at the stability order with minimum interaction energy among the monohydrates (Tables 4 and 5). The energy differences between the monohydrated tuftsin, retro tuftsin complexes at Pro, and the corresponding most stable monohydrated complexes are found to be 8.22 and 7.59 kcal/mol and 7.97 and 6.71 kcal/mol at B3LYP and MP2 levels of theory, respectively. With -12.73 and -12.3 kcal/mol interaction energy at B3LYP and MP2 levels of theory, respectively, the RTTW has been the second nominee in the stability order of retro tuftsin complexes. In RTTW, the binding water acts as intraresidual viaduct within Arg through one O-H…O (2.041 Å) and one N-H…O (1.988 Å) interactions, and it makes interresidual link between the two end amino acid residues Arg and Thr via another O-H…O interaction at 1.894 Å.
Table 4
Total energy
| Complex | ||||||||
| MP2 | B3LYP | MP2 | B3LYP | MP2 | B3LYP | MP2 | B3LYP | |
| Tuftsin | 1708.448 | 1713.730 | — | — | — | — | 8.225 | 7.304 |
| TTW | 1784.6594 | 1790.154 | 12.93 | 10.67 | 6.15 | 6.65 | 6.321 | 5.650 |
| TLW | 1784.680 | 1790.171 | 0.00 | 0.00 | 13.05 | 13.99 | 7.952 | 6.794 |
| TPW | 1784.6595 | 1790.1538 | 12.86 | 10.79 | 5.08 | 5.77 | 9.792 | 8.744 |
| TAW | 1784.666 | 1790.161 | 8.79 | 6.28 | 11.73 | 12.49 | 9.482 | 8.611 |
| Tuftsin…4W | 2013.315 | 2019.444 | — | — | 35.96 | 37.96 | 9.483 | 8.421 |
Table 5
Total energy
| Complex | ||||||||
| MP2 | B3LYP | MP2 | B3LYP | MP2 | B3LYP | MP2 | B3LYP | |
| Retro tuftsin | 1708.435 | 1713.723 | — | — | — | — | 7.694 | 6.864 |
| RTAW | 1784.661 | 1790.158 | 0.00 | 0.00 | 12.99 | 14.24 | 8.170 | 7.327 |
| RTPW | 1784.648 | 1790.148 | 8.16 | 6.28 | 6.28 | 6.65 | 6.374 | 5.720 |
| RTLW | 1784.655 | 1790.156 | 3.77 | 1.25 | 12.17 | 12.49 | 4.708 | 3.771 |
| RTTW | 1784.659 | 1790.157 | 1.26 | 0.63 | 12.30 | 12.73 | 8.963 | 8.016 |
| Retro tuftsin…4W | 2013.310 | 2019.445 | — | — | 39.53 | 41.85 | 6.060 | 5.400 |
In RTAW, the water molecule acts as a biprotonic donor, one to Arg carboxylic oxygen (1.917 Å) and other to Lys C=O oxygen atom (1.947 Å). It accepts one proton from the Arg N-H group with H-bond distance of 2.08 Å. The TLW is energetically favorable than RTLW. The energy difference between these two complexes is estimated to be 0.88 and 1.5 kcal/mol at MP2 and B3LYP levels of theory, respectively. But it is interesting to note that all the other three monohydrated retro tuftsin complexes (RTAW, RTTW, and RTPW) are found to be stable than their tuftsin partners with energy difference of 1.75, 6.08, and 0.88 kcal/mol at B3LYP and 1.26, 6.15, and 1.2 kcal/mol at MP2 levels of theory, respectively. In line with the above complexes, the Retro tuftsin…4W is also energetically favorable than Tuftsin…4W complex with almost 4 kcal/mol at both the tested levels of theory. This appears to be a result of the variation in the number and strength of H-bonds existing in the Retro tuftsin…4W complex. The relative energies of tuftsin and retro tuftsin complexes vary between 6.28-10.79 and 0.63-6.28 kcal/mol at B3LYP and 8.79-12.93 and 1.26-8.16 kcal/mol at MP2 levels of theory, respectively. The relative and interaction energy values are found to agree, and the strong bonds are associated with maximum interaction energy and hence higher stability in the hydrated tuftsin complexes.
The geometrical parameters of the intermolecular H-bonds are tabulated in Tables 6 and 7, respectively, for the hydrated tuftsin and retro tuftsin complexes. It is to be noted that the (O-H)water bond lengths have been elongated from its corresponding monomer in all the hydrated tuftsin and retro tuftsin complexes up to the maximum of 0.015 Å in tuftsin (TTW) and 0.016 Å in retro tuftsin (RTAW) complexes. Maximum elongation (0.021 Å) is observed for the carboxylic O-H (O18-H19) of Arg residue in Retro tuftsin…4W complex where the water molecule acts as a protonic donor as well as acceptor and links the ends of the Arg carboxylic group. The bonds present in the hydrated tuftsin and retro tuftsin complexes are stretched as a result of hydration at the sites related with the atoms involving H-bonds as that of our previous studies [44–46]. The analysis shows that the proton donor N-H involved in these hydration interactions is stretched both in tuftsin and retro tuftsin complexes with values ranging from 0.005 to 0.009 Å and 0.001 to 0.009 Å, respectively. The variation of
Table 6
Hydrogen bond parameters (bond length R (in Å), bond angles θ (in degree)), electron density
| Complex | Hydrogen bond | ε | |||||||||
| Monomer | Complex | ||||||||||
| TTW | O76-H78…O57 | 0.969 | 0.980 | 0.015 | 1.849 | 2.824 | 172.89 | 0.032 | 0.113 | 0.036 | 7.98 |
| C4-H7…O76 | 1.093 | 1.094 | 0.001 | 2.314 | 3.401 | 172.63 | 0.016 | 0.043 | 0.017 | 4.79 | |
| TLW | N60-H61…O76 | 1.022 | 1.031 | 0.009 | 1.943 | 2.965 | 171.22 | 0.031 | 0.087 | 0.033 | 15.56 |
| O76-H78…O39 | 0.969 | 0.979 | 0.010 | 1.877 | 2.844 | 168.72 | 0.030 | 0.106 | 0.024 | 6.81 | |
| O76-H77…O17 | 0.969 | 0.972 | 0.003 | 1.982 | 2.870 | 150.68 | 0.022 | 0.088 | 0.071 | 6.12 | |
| TPW | O76-H78…O10 | 0.969 | 0.979 | 0.010 | 1.885 | 2.837 | 163.43 | 0.033 | 0.108 | 0.018 | 7.26 |
| C13-H23…O76 | 1.091 | 1.092 | 0.001 | 2.308 | 3.394 | 172.16 | 0.017 | 0.045 | 0.020 | 4.14 | |
| TAW | O76-H77…O17 | 0.969 | 0.974 | 0.005 | 1.976 | 2.937 | 168.63 | 0.025 | 0.096 | 0.012 | 4.51 |
| O76-H78…O64 | 0.969 | 0.976 | 0.007 | 1.926 | 2.863 | 160.03 | 0.030 | 0.105 | 0.045 | 10.16 | |
| N30-H33…O76 | 1.017 | 1.023 | 0.006 | 1.949 | 2.951 | 165.92 | 0.032 | 0.097 | 0.021 | 15.33 | |
| Tuftsin…4W | N30-H33…O82 | 1.017 | 1.022 | 0.005 | 1.974 | 2.956 | 160.44 | 0.029 | 0.099 | 0.040 | 13.38 |
| O82-H84…O64 | 0.969 | 0.979 | 0.010 | 1.915 | 2.864 | 162.58 | 0.032 | 0.105 | 0.044 | 10.71 | |
| O82-H83…O17 | 0.969 | 0.972 | 0.003 | 2.063 | 3.000 | 161.27 | 0.020 | 0.087 | 0.010 | 3.14 | |
| O79-H81…O17 | 0.969 | 0.971 | 0.002 | 2.008 | 2.969 | 169.54 | 0.022 | 0.094 | 0.017 | 5.83 | |
| N60-H61…O79 | 1.022 | 1.030 | 0.008 | 1.941 | 2.926 | 159.21 | 0.033 | 0.101 | 0.036 | 14.50 | |
| O79-H80…O39 | 0.969 | 0.977 | 0.008 | 1.858 | 2.807 | 162.96 | 0.034 | 0.110 | 0.008 | 8.50 | |
| O76-H78…O57 | 0.969 | 0.979 | 0.010 | 1.859 | 2.830 | 170.85 | 0.035 | 0.106 | 0.017 | 6.47 | |
| C4-H7…O76 | 1.093 | 1.094 | 0.001 | 2.254 | 3.332 | 167.78 | 0.017 | 0.068 | 0.010 | 5.70 | |
| O85-H87…O10 | 0.969 | 0.977 | 0.008 | 1.950 | 2.849 | 160.13 | 0.030 | 0.106 | 0.029 | 6.77 | |
| C13-H23…O85 | 1.091 | 1.093 | 0.002 | 2.316 | 3.395 | 169.25 | 0.015 | 0.058 | 0.129 | 4.03 | |
Table 7
Hydrogen bond parameters (bond length R (in Å), bond angles θ (in degree)), electron density
| Complex | Hydrogen bond | ε | |||||||||
| Monomer | Complex | ||||||||||
| RTAW | O18-H19…O76 | 0.977 | 0.996 | 0.019 | 1.783 | 2.704 | 152.25 | 0.045 | 0.104 | 0.025 | 21.36 |
| O76-H78…O17 | 0.969 | 0.985 | 0.016 | 1.881 | 2.719 | 141.16 | 0.035 | 0.111 | 0.078 | 9.52 | |
| RTPW | O76-H77…O10 | 0.969 | 0.979 | 0.010 | 1.869 | 2.814 | 161.37 | 0.033 | 0.108 | 0.008 | 7.26 |
| C14-H24…O76 | 1.095 | 1.095 | 0.000 | 2.559 | 3.609 | 160.27 | 0.007 | 0.036 | 0.054 | 4.14 | |
| RTLW | O76-H77…O17 | 0.969 | 0.978 | 0.009 | 1.917 | 2.857 | 160.22 | 0.031 | 0.103 | 0.017 | 8.03 |
| O76-H78…O39 | 0.969 | 0.977 | 0.008 | 1.947 | 2.828 | 148.74 | 0.028 | 0.103 | 0.057 | 3.22 | |
| N11-H12…O76 | 1.014 | 1.019 | 0.005 | 2.080 | 2.979 | 145.76 | 0.024 | 0.088 | 0.045 | 8.72 | |
| RTTW | O76-H77…O56 | 0.969 | 0.974 | 0.005 | 1.894 | 2.822 | 158.14 | 0.029 | 0.110 | 0.021 | 9.83 |
| N30-H33…O76 | 1.012 | 1.021 | 0.009 | 1.988 | 3.002 | 172.16 | 0.027 | 0.096 | 0.036 | 12.26 | |
| O76-H78…O17 | 0.969 | 0.972 | 0.003 | 2.014 | 2.984 | 176.08 | 0.022 | 0.091 | 0.019 | 4.43 | |
| Retro tuftsin…4W | O79-H81…O10 | 0.969 | 0.976 | 0.007 | 1.962 | 2.905 | 161.84 | 0.025 | 0.080 | 0.039 | 5.40 |
| O79-H80…O18 | 0.969 | 0.970 | 0.001 | 2.330 | 3.118 | 137.96 | 0.012 | 0.042 | 0.054 | 1.89 | |
| C13-H23…O79 | 1.092 | 1.093 | 0.001 | 2.322 | 3.315 | 150.16 | 0.015 | 0.044 | 0.057 | 3.59 | |
| O18-H19…O82 | 0.977 | 0.998 | 0.021 | 1.765 | 2.694 | 153.31 | 0.042 | 0.123 | 0.020 | 22.84 | |
| O82-H83…O17 | 0.969 | 0.983 | 0.014 | 1.911 | 2.727 | 138.59 | 0.030 | 0.094 | 0.049 | 7.25 | |
| N30-H33…O82 | 1.012 | 1.014 | 0.002 | 2.376 | 3.313 | 153.37 | 0.001 | 0.038 | 0.028 | 3.17 | |
| N11-H12…O76 | 1.014 | 1.017 | 0.003 | 2.289 | 3.050 | 130.67 | 0.015 | 0.049 | 0.127 | 3.94 | |
| O76-H77…O17 | 0.969 | 0.976 | 0.007 | 1.958 | 2.905 | 162.77 | 0.026 | 0.080 | 0.019 | 5.50 | |
| O76-H78…O39 | 0.969 | 0.977 | 0.008 | 1.927 | 2.842 | 154.79 | 0.027 | 0.088 | 0.034 | 3.55 | |
| N54-H53…O76 | 1.021 | 1.020 | 0.001 | 2.369 | 3.223 | 140.68 | 0.011 | 0.039 | 0.128 | 1.33 | |
| N54-H75…O85 | 1.022 | 1.024 | 0.002 | 2.076 | 3.100 | 178.38 | 0.023 | 0.062 | 0.026 | 10.27 | |
| O85-H86…O56 | 0.969 | 0.984 | 0.015 | 1.820 | 2.783 | 165.18 | 0.035 | -1.713 | 0.018 | 6.05 | |
The dipole moment of the tuftsin is found to be 7.3 Debye at B3LYP and 8.23 Debye at MP2 levels of theory, respectively. This value is found to lie between 5.65 and 8.74 Debye at B3LYP and 6.32 and 9.79 Debye at MP2 levels of theory, respectively. It is noteworthy that the least stable complex TPW is observed to possess the higher value of dipole moment calculated at MP2 and B3LYP levels of theory (9.8 and 8.7 Debye, respectively). The dipole moments of the retro tuftsin complexes are lower than tuftsin, and it lies in between 3.77 and 8.02 Debye at B3LYP and 4.71 and 9.0 Debye at MP2 levels of theory, respectively.
Pictorial representation of HOMO and LUMO of isolated tuftsin, TLW, and the most stable tuftsin monohydrate and Tuftsin…4W complexes is presented in Figure 7. Figure 8 shows the HOMO and LUMO for isolated retro tuftsin along with its most stable monohydrate RTAW and Retro tuftsin…4W complexes. The analysis of these frontier molecular orbitals of tuftsin shows that the HOMO is concentrated on the distal end of the Arg residue which is capped by the guanidinium group where the LUMO is across the carboxylic acid group of Arg. Similar characteristics prevail for the HOMO and LUMO of TLW and Tuftsin…4W complexes. As observed in Figure 8, it is interesting to note that the HOMO of retro tuftsin, RTAW, and Retro tuftsin…4W complexes are across the guanidinium group of Arg similar to their tuftsin counterparts irrespective of the position of Arg in the peptide chain of the tuftsin and its retro form. The LUMO of retro tuftsin is found to be concentrated on Lys where the LUMO of RTAW and Retro tuftsin…4W are across the carboxylic group of Arg. The electrostatic potential map [53] is a very useful three-dimensional diagram of molecules to visualize their size and shape. The electrostatic potential map for tuftsin, retro tuftsin, and their four water complexes is shown in Figures 9 and 10, respectively. The C=O oxygen atom of Thr residue in tuftsin, Lys residue in retro tuftsin, and Retro tuftsin…4W complexes are colored red corresponding to the strong negative potential charge (-0.006 a.u.) that confirmed the presence of electrostatic interaction. In Tuftsin…4W complex, the nitrogen atom which belongs to the guanidinium group of Arg shows the larger electron density in its region. The amino group of Lys side chain in tuftsin and one end of Thr side chain in retro tuftsin is white denoting the concentration of strong positive potential charge (0.006 a.u.). Similar style is observed for Tuftsin…4W and Retro tuftsin…4W complexes. From Figures 9 and 10, it can be noticed that the tuftsin and retro tuftsin tetra peptides have interacted electro statically with the neighboring water molecules mainly through the C=O and N-H groups. The blue color symbolizes the concentration of modest positive charges. The electro static interactions decrease from red to white through green, violet, and pink.
[figure(s) omitted; refer to PDF]
3.1.1. AIM Analysis
The nature of bonding between the atoms can be characterized through AIM theory by the values of electron density at the bond critical point (BCP). The low values of charge density (
For retro tuftsin complexes, the H-bond distance, bond angle, and
The positive values of
The correlation between the H-bond distance, electron density, and Laplacian of electron density has been established for Tuftsin…4W and Retro tuftsin…4W complexes (Figures 11 and 12, respectively). This indicates that the bond length and electron density are inverse to each other with correlation coefficients 0.937, 0.955, 0.989, and 0.975for the electron density and its Laplacian with H-bond lengths, respectively. The maximum ellipticity value (0.129) is noted for the Cα-H (Arg)…O (W) interaction in Tuftsin…4W complex which indicates the higher chance of structural deformation under external perturbations. In hydrated retro tuftsin complexes, the N54-H53 (Lys)…O (W) interaction in Retro tuftsin…4W complex has maximum ellipticity value (0.128).
[figure(s) omitted; refer to PDF]
3.1.2. NBO Analysis
The stabilization energies
In hydrated retro tuftsin complexes, the interactions where the oxygen of water offers lone pairs to the O-H antibond orbital of retro tuftsin are found to be stronger with
4. Conclusion
The tuftsin is sensitive to structural changes, and the results indicate that water-tuftsin H-bonds, in addition to intramolecular H-bonds, stabilize the β-turn structure with H-bonds between Thr and Arg residues. Difference in the stability of the hydrated complexes is confined to the amino acid residues at which the water molecule is attached to tuftsin. When the water molecules are attached to the oxygen sites of amino acid residues, strong H-bonds are observed to be formed between tuftsin, retro tuftsin complexes, and water molecules. The relative energy values of the hydrated tuftsin complexes match with the interaction energies where the strong H-bonds are said to possess the maximum interaction energy and hence higher stability. The H-bonds present in the hydrated tuftsin complexes are noticed to be stronger than those of retro tuftsin counterparts. It is interesting to notice that irrespective of the position in the peptide chain, monohydrated tuftsin complexes at Pro residue stand last at the stability order with minimum interaction energy among the monohydrates.
The H-bonds in the hydrated tuftsin are observed to be stronger than those present in the retro tuftsin complexes. The analysis of the frontier molecular orbitals of tuftsin shows that the HOMO is concentrated on the distal end of the Arg residue which is capped by the guanidinium group where the LUMO is across the carboxylic acid group of Arg. The AIM calculation indicates that O-H…O, N-H…O, and C-H…O interactions present in the hydrated tuftsin and retro tuftsin complexes possess low
[1] K. Nishioka, A. Constantopoulos, P. S. Satoh, V. A. Najjar, "The characteristics, isolation and synthesis of the phagocytosis stimulating peptide tuftsin," Biochemical and Biophysical Research Communications, vol. 47 no. 1, pp. 172-179, DOI: 10.1016/S0006-291X(72)80025-1, 1972.
[2] M. Fridkin, Y. Stabinsky, V. Zakuth, Z. Spirer, "Tuftsin and some analogs: synthesis and interaction with human polymorphonuclear leukocytes," Biochimica et Biophysica Acta, vol. 496 no. 1, pp. 203-211, DOI: 10.1016/0304-4165(77)90129-5, 1977.
[3] M. Fridkin, P. Gottlieb, T.-L.-P.-A. Tuftsin, "Tuftsin, Thr-Lys-Pro-Arg. Anatomy of an immunologically active peptide," Molecular and Cellular Biochemistry, vol. 41, pp. 73-97, 1981.
[4] V. A. Najjar, "The clinical and physiological aspects of tuftsin deficiency syndromes exhibiting defective phagocytes," Klinische Wochenschrift, vol. 37, pp. 751-756, DOI: 10.1007/BF01478032, 1979.
[5] K. Nishioka, "Anti-tumour effect of the physiological tetrapeptide, tuftsin," British Journal of Cancer, vol. 39 no. 3, pp. 342-345, DOI: 10.1038/bjc.1979.59, 1979.
[6] I. Florentin, M. Bruley-Rosset, N. Kiger, J. L. Imbach, F. Winternitz, G. Mathe, "In vivo immunostimulation by tuftsin," Cancer Immunology, Immunotherapy, vol. 5, pp. 211-216, DOI: 10.1007/BF00199631, 1978.
[7] G. R. Corazza, G. Zoli, L. Ginaldi, C. Cancellieri, V. Profeta, G. Gasbarrini, D. Quaglino, "Tuftsin deficiency in AIDS," Lancet, vol. 337 no. 8732, pp. 12-13, DOI: 10.1016/0140-6736(91)93331-3, 1991.
[8] D. B. Perkowska, F. Muzalewski, D. Konopinska, "Antibacterial properties of tuftsin and its analogs," Antimicrobial Agents and Chemotherapy, vol. 25 no. 1, pp. 134-136, DOI: 10.1128/AAC.25.1.134, 1984.
[9] I. Z. Siemion, M. Lisowski, K. Sobczyk, "Conformational investigations in the tuftsin group," Annals of the New York Academy of Sciences, vol. 419 no. 1 Antineoplasti, pp. 56-63, DOI: 10.1111/j.1749-6632.1983.tb37091.x, 1983.
[10] I. Z. Siemion, P. Stefanowicz, A. Kluczyk, J. J. Slon, "Circular dichroism of the peptides of tuftsin-kentsin group," Polish Journal of Chemistry, vol. 67, pp. 2171-2180, 1993.
[11] E. Tzehoval, S. Segal, Y. Stabinsky, M. Fridkin, Z. Spirer, M. Feldman, "Tuftsin (an Ig-associated tetrapeptide) triggers the immunogenic function of macrophages: implications for activation of programmed cells," Proceedings of the National Academy of Sciences, vol. 75 no. 7, pp. 3400-3404, DOI: 10.1073/pnas.75.7.3400, 1978.
[12] S. Fitzwater, Z. I. Hodes, H. A. Scheraga, "Conformational energy study of tuftsin," Macromolecules, vol. 11 no. 4, pp. 805-811, DOI: 10.1021/ma60064a036, 1978.
[13] M. Blumenstein, P. P. Layne, V. A. Najjar, "Nuclear magnetic resonance studies on the structure of the tetrapeptide tuftsin L-threonyl-L-lysyl-L-prolyl-L-arginine, and its pentapeptide analog L-threonyl-L-lysyl-L-prolyl-L-prolyl-L-arginine," Biochemistry, vol. 18 no. 23, pp. 5247-5253, DOI: 10.1021/bi00590a032, 1979.
[14] K. Nishioka, P. S. Satoh, A. Constantopoulos, V. A. Najjar, "The chemical synthesis of the phagocytosis-stimulating tetrapeptide tuftsin (Thr-Lys-Pro-Arg) and its biological properties," Biochimica et Biophysica Acta (BBA)-Protein Structure, vol. 310 no. 1, pp. 230-237, DOI: 10.1016/0005-2795(73)90025-1, 1973.
[15] A. D. Ursi, M. Pegna, P. Amodeo, H. Molinari, A. Verdini, L. Zetta, P. A. Temussi, "Solution conformation of tuftsin," Biochemistry, vol. 31 no. 40, pp. 9581-9586, DOI: 10.1021/bi00155a010, 1992.
[16] I. P. Sekacis, E. E. Liepins, N. I. Veretennikova, G. I. Chipens, "1H and 13C-NMR studies of tuftsin conformation," Bioorganicheskaia Khimiia, vol. 5, pp. 1617-1622, 1979.
[17] S. D. O’Connor, P. E. Smith, F. Al-Obeidi, B. M. Pettitt, "Quenched molecular dynamics simulations of tuftsin and proposed cyclic analogues," Journal of Medicinal Chemistry, vol. 35 no. 15, pp. 2870-2881, DOI: 10.1021/jm00093a021, 1992.
[18] I. Z. Siemion, M. Lisowski, D. Konopinska, E. Nawrocka, "13 C nuclear magnetic resonance and circular dichroism studies of the tuftsin conformation in water," European Journal of Biochemistry, vol. 112 no. 2, pp. 339-343, DOI: 10.1111/j.1432-1033.1980.tb07210.x, 1980.
[19] G. V. Nikiforovich, "Calcualtion of stable conformations of tuftsin," Bioorganicheskaia Khimiia, vol. 4, pp. 1427-1430, 1978.
[20] G. I. Chipens, N. I. Veretennikova, G. I. Nikiforovich, Z. A. Atare, "Elongated and Cyclic Analogues of Tuftsin and Rigin," Peptides 1980, Copenhagen: Scriptor, 1981.
[21] M. Z. Siddiqui, A. K. Sharma, S. Kumar, "Solution conformation of tuftsin," International Journal of Biological Macromolecules, vol. 19 no. 2, pp. 99-102, DOI: 10.1016/0141-8130(96)01108-7, 1996.
[22] I. Z. Siemion, A. Kluczyk, "Tuftsin: on the 30-year anniversary of Victor Najjar’s discovery," Peptides, vol. 20 no. 5, pp. 645-674, DOI: 10.1016/S0196-9781(99)00019-4, 1999.
[23] V. Kothekar, G. Ashish, D. Gupta, R. Kishore, "Theoretical study of conformational flexibility of tuftsin in vacuum and in aqueous environment," Indian Journal of Biochemistry & Biophysics, vol. 36 no. 1, pp. 14-28, 1999.
[24] A. S. Sobczyk, I. Z. Siemion, E. Nawrocka, "The influence of configuration changes in tuftsin peptide chain on its folding properties," Acta Biochimica Polonica, vol. 27, pp. 353-363, 1980.
[25] C. V. Valdeavella, H. O. Blatt, B. M. Pettitt, "Simulations of conformers of tuftsin and a cyclic tuftsin analog," International Journal of Peptide and Protein Research, vol. 46 no. 5, pp. 372-380, 1995.
[26] M. Fridkin, V. A. Najjar, "Tuftsin: its chemistry, biology, and clinical potentia," Critical Reviews in Biochemistry and Molecular Biology, vol. 24 no. 1,DOI: 10.3109/10409238909082550, 1989.
[27] K. Yasumura, K. Okamoto, S. Shimamura, "Synthesis of peptides related to tuftsin," Yakugaku zasshi: Journal of the Pharmaceutical Society of Japan, vol. 97 no. 3, pp. 324-329, DOI: 10.1248/yakushi1947.97.3_324, 1977.
[28] D. Voet, J. G. Voet, C. W. Prat, Fundamentals of Biochemistry, 1998.
[29] D. Liu, T. Wyttenbach, P. E. Barran, M. T. Bowers, "Sequential hydration of small protonated peptides," Journal of the American Chemical Society, vol. 125 no. 28, pp. 8458-8464, DOI: 10.1021/ja034638x, 2003.
[30] T. Wyttenbach, M. T. Bowers, "Hydration of biomolecules," Chemical Physics Letters, vol. 480 no. 1-3,DOI: 10.1016/j.cplett.2009.08.042, 2009.
[31] N. Samir Kumar, A. Ashish, S. Veena, "Molecular dynamics simulation of tuftsin and its analogs in a receptor like environment," The Pharma Innovation Journal, vol. 3 no. 11, pp. 55-67, 2015.
[32] A. E. Kovalenko, E. A. Pashkina, L. Y. Kanazhevskaya, A. N. Masliy, V. A. Kozlov, "Chemical and biological properties of a supramolecular complex of tuftsin and cucurbit[7]uril," International Immunopharmacology, vol. 47, pp. 199-205, DOI: 10.1016/j.intimp.2017.03.032, 2017.
[33] T. Kolomin, M. Shadrina, P. Slominsky, S. Limborska, N. Myasoedov, "A new generation of drugs: synthetic peptides based on natural regulatory peptides," Neuroscience and Medicine, vol. 4 no. 4, pp. 223-252, DOI: 10.4236/nm.2013.44035, 2013.
[34] M. Alam Khan, "Targeted drug delivery using tuftsin-bearing liposomes: implications in the treatment of infectious diseases and tumors," Current Drug Targets, vol. 22 no. 7, pp. 770-778, DOI: 10.2174/1389450121999201125200756, 2021.
[35] J. Huang, J. Wang, Z. Wang, M. Chu, Y. Wang, "Tuftsin: a natural molecule against SARS-CoV-2 infection," Frontiers in Molecular Biosciences, vol. 9,DOI: 10.1101/2022.01.10.475746, 2022.
[36] A. D. Becke, "Density – functional exchange – energy approximation with correct asymptotic behaviour," Physical Review A, vol. 38, pp. 3098-3100, DOI: 10.1103/PhysRevA.38.3098, 1998.
[37] C. Lee, W. Yang, R. G. Parr, "Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density," Physical Review B, vol. 37 no. 2, pp. 785-789, DOI: 10.1103/PhysRevB.37.785, 1988.
[38] C. Mollar, M. S. Plesset, "Note on an approximation treatment for many-electron systems," Physics Review, vol. 46 no. 7, pp. 618-622, DOI: 10.1103/PhysRev.46.618, 1934.
[39] S. F. Boys, F. Bernardi, "The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors," Molecular Physics, vol. 19 no. 4, pp. 553-566, DOI: 10.1080/00268977000101561, 1970.
[40] R. F. W. Bader, Atoms in Molecules: A Quantum Theory, 1990.
[41] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, V. G. Zakrzewski, J. A. Montgomery, R. E. Stratmann, J. C. Burant, S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, M. C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui, K. Morokuma, N. Rega, P. Salvador, J. J. Dannenberg, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. Cioslowski, J. V. Ortiz, A. G. Baboul, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, J. L. Andres, C. Gonzalez, M. Head-Gordon, E. S. Replogle, J. A. Pople, Gaussian03, Revision B.05, 2003.
[42] I. Z. Siemion, D. Konopinska, "Tuftsin analogs and their biological activity," Molecular and Cellular Biochemistry, vol. 41, pp. 99-112, 1981.
[43] B. Yogeswari, R. Kanakaraju, A. Abiram, P. Kolandaivel, "Molecular dynamics and quantum chemical studies on incremental solvation of glycine," Computational and Theoretical Chemistry, vol. 967 no. 1, pp. 81-92, DOI: 10.1016/j.comptc.2011.03.045, 2011.
[44] B. Yogeswari, R. Kanakaraju, S. Boopathi, P. Kolandaivel, "Microsolvation and hydrogen bond interactions in glycine dipeptide: molecular dynamics and density functional theory studies," Journal of Molecular Graphics & Modelling, vol. 35, pp. 11-20, DOI: 10.1016/j.jmgm.2012.02.002, 2012.
[45] B. Yogeswari, R. Kanakaraju, S. Boopathi, P. Kolandaivel, "Molecular dynamics and quantum chemical studies of solvent effects on cyclo glycylglycine and glycylalanine dipeptides," Molecular Simulations, vol. 39 no. 8, pp. 670-687, DOI: 10.1080/08927022.2012.758852, 2013.
[46] B. Yogeswari, R. Kanakaraju, S. Boopathi, P. Kolandaivel, "Combined theoretical studies on solvation and hydrogen bond interactions in glycine tripeptide," Molecular Simulations, vol. 40 no. 12, pp. 942-958, DOI: 10.1080/08927022.2013.828837, 2014.
[47] D. Konopinska, I. Z. Siemion, S. Szymaniec, S. Slopek, "Tuftsin and its analogs. Part IV. Synthesis of new tuftsin analogs based on simulation of genetic mutations," Polish Journal of Chemistry, vol. 52, pp. 573-580, 1978.
[48] J. Symersky, K. Blaha, J. Jecny, "Structure of cyclo tri-D-azetidine-2-carboxylic acid," Acta Crystallographica Section C: Crystal Structure Communications, vol. 44, pp. 148-150, 1988.
[49] R. Chandrasekaran, A. V. Lakshminarayan, U. V. Pandya, G. N. Ramachandran, "Conformation of the LL and LD hairpin bends with internal hydrogen bonds in proteins and peptides," Biochimica et Biophysica Acta, vol. 303 no. 1, pp. 14-27, DOI: 10.1016/0005-2795(73)90143-8, 1973.
[50] D. Konopinska, E. Nawrocka, I. Z. Siemion, S. Szymaniec, S. Slopek, Peptides, Proceedings of 14th European peptide symposium, 1976.
[51] A. S. Sobczyk, I. Z. Siemion, D. Konopinska, "Infrared spectroscopic investigations of tuftsin and its analogs," European Journal of Biochemistry, vol. 96 no. 1, pp. 131-139, DOI: 10.1111/j.1432-1033.1979.tb13022.x, 1979.
[52] M. Kahn, B. Devens, "The design and synthesis of a nonpeptide mimic of an immunosuppressing peptide," Tetrahedron Letters, vol. 27 no. 40, pp. 4841-4844, DOI: 10.1016/S0040-4039(00)85077-7, 1986.
[53] B. D. Boyd, "On the rhodanines and their presence in biologically active ligands," Journal of Molecular Structure, vol. 401 no. 3, pp. 227-234, DOI: 10.1016/S0166-1280(97)00024-9, 1997.
[54] U. Koch, P. L. A. Popelier, "Characterization of C-H-O hydrogen bonds on the basis of the charge density," The Journal of Physical Chemistry, vol. 99 no. 24, pp. 9747-9754, DOI: 10.1021/j100024a016, 1995.
[55] P. L. A. Popelier, R. F. W. Bader, "The existence of an intramolecular C-H-O hydrogen bond in creatine and carbamoyl sacrosine," Chemical Physics Letters, vol. 189 no. 6, pp. 542-548, DOI: 10.1016/0009-2614(92)85247-8, 1992.
[56] H. Raissi, M. Yoosefian, F. Mollania, F. Farzad, A. R. Nowroozi, D. Loghmaninejad, "Ab initio and DFT computational studies on molecular conformations and strength of the intramolecular hydrogen bond in different conformers of 3-amino-2-iminomethyl acryl aldehyde," Computational and Theoretical Chemistry, vol. 966 no. 1-3, pp. 299-305, DOI: 10.1016/j.comptc.2011.03.026, 2011.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Copyright © 2022 B. Yogeswari et al. This is an open access article distributed under the Creative Commons Attribution License (the “License”), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License. https://creativecommons.org/licenses/by/4.0/
Abstract
Density functional B3LYP method has been used to study the molecular properties of the tuftsin tetrapeptide (threonine-lysine-proline-arginine) and its retro form (arginine-proline-lysine-threonine). The influence of single water molecule on the conformations and relative stabilities of solvated tuftsin complexes has been studied by placing the water molecule at the individual amino acid residues of both the tuftsin complexes. The contribution of four water molecules to the system energetics of tuftsin complexes has also been analyzed. The conformational changes occurred in the solvated tuftsin complexes have been explored through their dihedral angles. The tuftsin is found to be sensitive to structural changes, and our results indicate that water-tuftsin hydrogen bonds (H-bonds), in addition to intramolecular H-bonds, stabilize the β-turn structure with H-bonds between threonine and arginine residues of tuftsin. Difference in the stability of the hydrated complexes is confined to the amino acid residues at which the water molecule is attached to tuftsin. The interaction energy calculations have been used to investigate the strength of the intermolecular H-bond interactions. The AIM theory and NBO analysis were employed to survey the H-bonding patterns in hydrated tuftsin complexes. The maximum ellipticity value (0.129) is noted for the Cα-H (Arg)…O (W) interaction in Tuftsin…4W complex which indicates the higher chance of structural deformation under external perturbations. The interactions between oxygen lone pairs in water and C-H antibond orbitals of tuftsin and retro tuftsin complexes exist with
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details
; Tamilselvan, K S 2 ; Thanikaikarasan, S 3 ; Lal, N Dayanand 4 ; Anandaram, Harishchander 5 ; Madhusudhanan, J 6 ; Karthik, R 7
; Batu, Areda 8
1 Department of Physics, Sri Eshwar College of Engineering, Coimbatore, 641 202 Tamil Nadu, India
2 Department of Electronics and Communication Engineering, KPR Institute of Engineering and Technology, Coimbatore, India
3 Department of Physical Sciences, Saveetha School of Engineering, Saveetha Institute of Medical And Technical Sciences (SIMATS), Thandalam, 602105 Tamil Nadu, India
4 Department of Computer Science and Engineering, GITAM School of Technology Bengaluru, India
5 Centre for Excellence in Computational Engineering and Networking, Amrita Vishwa Vidyapeetham, Coimbatore, Tamil Nadu, India
6 Department of Biotechnology, Anand Institute of Higher Technology, OMR, Kazhipattur, 603103, Chennai, Tamil Nadu, India
7 School of Electronics and Communication Engineering, REVA University, Bangalore, India
8 Center of Excellence for Artificial Intelligence & Robotics, Department of Chemical Engineering, College of Biological and Chemical Engineering, Addis Ababa Science and Technology University, Addis Ababa, Ethiopia





