Selective complexation of alkaline earth metal ions with nanotubular cyclopeptides: DFT theoretical study

Fereshte Shahangia, Alireza Najafi Chermahini*a, Hossein Farrokhpoura and Abbas Teimourib
aDepartment of Chemistry, Isfahan University of Technology, Isfahan 84156-83111, Iran. E-mail: anajafi@cc.iut.ac.ir; Fax: +983113913250; Tel: +983113913251
bChemistry Department, Payame Noor University, 19395-4697 Tehran, Iran

Received 7th August 2014 , Accepted 1st December 2014

First published on 1st December 2014


Abstract

The interaction of alkaline earth metal cations including Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ with cyclic peptides containing 3 or 4 (S) alanine molecules (CyAla3 and CyAla4) was investigated by density functional theory (DFT-CAM-B3LYP and DFT-B3LYP). A mixed basis set including 6-31+G(d) for C, H, O, Be2+, Mg2+, Ca2+ and LANL2DZ for Sr2+ and Ba2+ were used for calculations. The optimized structures, binding energies, and various thermodynamic parameters of free ligands and related metal cation complexes were determined. The order of strength of interaction energies was found as Be2+ > Mg2+ > Ca2+ > Sr2+ > Ba2+. Vibrational frequency calculations showed that the selected cyclic peptides and their complexes with the alkaline earth metal cations were at local minima of their potential energy surfaces. In addition, it was found that the larger cavity CyAla4 ligand, can hold the alkaline metal cations better than CyAla3 molecule when the same metal cation is in the structure of complex. Moreover, analyzing the geometry of [M/CyAla3]2+ and [M/CyAla4]2+ complexes indicated that the aggregation with metal cation, caused substantial changes in the geometrical parameters of ligands.


1. Introduction

Supramolecular chemistry appeared when the Nobel Prize was awarded to Charles J Pedersen, Donald J Cram and Jean-Marie Lehn in 1987. Lehn specified supramolecular chemistry as ‘chemistry beyond the molecule’, i.e. the chemistry of molecular aggregates assembled via non-covalent interactions.1 After two decades, supramolecular chemistry is an essential, knowledge based branch of science encompassing opinions of physical and biological processes. Host–guest chemistry is an example of supramolecular chemistry. It is the study of complexes that are composed of molecules or ions held together by intermolecular forces, such as electrostatic interactions, hydrogen bonding, and dispersion interactions, and solvophobic effects not by covalent bonds.2 The discovery of crown ethers is the milestone for starting the extensive evolution of host–guest chemistry in 1967.3 Shortly afterwards, various classes of macrocyclic ligands with structures of increasing complexity were synthesized, including cryptands,4 cavitands,5 carcerands,6 cyclodextrins (CDs),7,8 macrocyclic antibiotics,9,10 proteins11 and chiral micelles.12 In the host–guest chemistry an inclusion compound is a complex in which one chemical compound (“host”) forms an enclosed space in which molecules of a second “guest” compound are situated.13,14 There are various experimental15–17 and theoretical18–22 studies devoted to investigate different aspects of this phenomenon. The definition of inclusion compounds is very broad: for example in molecular encapsulation a guest molecule is actually trapped inside another molecule.23

In recent years, a new fascinating class of organic compounds has been reported in which amino acid unites make a macrocycle named cyclic peptide.24–28 Cyclic peptides have been defined in many natural environments and display a wide spectrum of biological activity.29 For example they have antibacterial,30 antiviral,31 antifungal,32 immunosuppressant,33 and antinociceptive properties.34 Their amphiphilic characteristics make them to be potential superior candidates of surfactants.35 Also, cyclic peptides can self-assemble into peptide nanotubes, as models of biological transmembrane channels.36,37 Such surfaces and their biological properties have attracted interest in the structures of cyclic peptides and their behaviors at the hydrophilic/hydrophobic interfaces. The structure and properties of cyclic peptides have been deeply studied and results reported in literature. Chen and co-workers have studied characteristics of cyclic peptides based on the density function theory (DFT-B3LYP) and examined the effect of the substituents and ring size on molecular structure of cyclic peptides.24 Poteau and Trinquier investigated the structures of all-cis cyclopolyglycines, cis cyclopolyalanines and cyclopolyphenylalanines based on theoretical approaches.28 Vijayaraj et al. reported structures and geometries of cyclic peptide nanotubes by molecular dynamic simulations.38,39 Mazurek and co workers studied structures and properties of cyclo glycine and compared these with its phosphor analogues.40 In addition, Jishi et al. investigated formation of dimers of cyclo[(Gly-D-Ala)4] and concluded that dimer formation is favored by hydrogen bonding.41 Hongge Zhao and co-workers used a cyclic decapeptide and the enantiomers of 1-phenyl-1-propanol as the host and guest molecules, respectively, to examine the separation ability of guest enantiomers by the cyclic peptide.42 Collision-induced dissociation (CID) of protonated peptides are the most frequently practiced MS/MS technology in proteomics.43–45 In collision-induced dissociation of a peptide, cleavage of an amide bond can result in namely b fragment ion with a five-membered oxazolone ring on the C-terminal side as first postulated by Harrison.46,47 These oxazolone structures can isomerize to macrocyclic peptides via a head to tail nucleophilic attack from the N-terminus.

The interaction of metal cations with cyclic peptides has been subjected of various studies especially for obtaining sequence information.48–58 For example Williams and Brodbelt used low energy collisionally activated dissociation (CAD) in a quadrupole ion trap were used to characterize the fragmentation of alkali, alkaline earth and transition metal complexes of five cyclic peptides.48 Moreover, Zhang et al. studied the interaction of disulfide-constrained cyclic tetrapeptides with Cu2+.57 In addition, Ruotolo and co-workers performed a conformational analysis of Gramicidin S, a cyclic antimicrobial peptide and found a β-sheet conformational preference.52

Recently, we have investigated the ability of cyclo alanines with different sizes for separating lactic acid enantiomers and metal alkali cations.59,60 Our previous theoretical calculations have pointed that CyAla3 and CyAla4 cyclic peptides are appropriate ligands for the separation of Li+ and Na+ from other alkali metal ions. Additionally, the binding energy of Li+ is greater than Na+ metal ion due to the smaller size of the Li+ ion. In continuum with our previous studies, our aim in this work is to employ the DFT approach along with a suitable basis set, to examine the influence of the alkaline earth metal ions nature on the metal binding selectivity by the cis CyAla3 and CyAla4 cyclic peptides. The second goal of this theoretical study is anticipating the efficiency of cyclic peptides for selective extracting of different metal ions. The results obtained in this work could be useful for predicting the applicability of an extractant for different metal ions, the material design of metal ion recognition and the other related fields. Also, investigation of interactions of cyclic peptides with guest molecules, as inclusion complexes, could help us to explain the features responsible for the remarkable potency of cyclic peptides.

2. Computational methods

DFT calculations were applied to optimize the structures of selected cyclic peptides in this work. Vibrational frequency calculations were also performed to verify that the optimized structures are in local minima on their potential energy surfaces. The original geometries of cyclic peptides were taken from the structures reported by Poteau and Trinquier.28 The optimized structures of cyclic peptides were used for studying the interaction of alkaline earth metal ions at the DFT/B3LYP and DFT/CAM-B3LYP level of theory. Briefly, the CAM-B3LYP method combines the features of hybrid functionals such as B3LYP61–63 with the long-range corrected functionals of Hirao et al.61 The exchange functional is considered as a mixture of exact, i.e., Hartree–Fock and DFT exchange, but, unlike B3LYP, the ratio of exact to DFT exchange varies in different regions of the molecule. The key improvement in this method is that the short range DFT exchange interaction is incorporated in the short-range DFT exchange functional but, the correct long-range interaction is described via HF exchange. In this work a mixed basis set including 6-31+G(d) for C, H, O, N, Be2+, Mg2+, Ca2+ and the effective core potential (ECP) of LANL2DZ for Sr2+ and Ba2+ have been used for the calculations. In addition, we decided to use B3LYP in the present study. The B3LYP functional has been widely used, and is generally considered satisfactory for alkali–ion complexes.64–66

All local energy minimum structures found by potential energy surface (PES) scan (relax) calculations were fully optimized at the B3LYP/6-31+G(d) level of theory.67 For scanning metal cations, we used the following coordinate system. The proper cyclic peptide was positioned around the z-axis where all oxygen or nitrogen atoms were in the xy plan. In addition, a dummy atom was put in the center of the macrocyle. Then, the earth alkaline metal cation was scanned along the z-axis. Initial positions were generated by movement of M2+ cations along the z-axis. Interaction energies were corrected by zero point energy (ZPE) and the basis set superposition error (BSSE)68 was taken into account by the counterpoise method. The natural bond orbital (NBO) analysis69,70 at the CAM-B3LYP/6-31+G(d) level of theory was performed to characterize the second-order interaction energy. All calculations were performed with the GAUSSIAN 09 computational chemistry package71 without any limitation. The atoms in molecule (AIM)72,74 at the CAM-B3LYP/6-31+G(d) level was used here to describe the binding characteristic between donor and acceptor.

3. Results and discussion

3.1 Geometrical parameters

In the present study, two cyclic peptides constructed from 3 or 4 l-alanine molecules, named CyAla3 and CyAla4, with amide groups in the cis conformation have been selected. The optimized structures for the cyclic peptides in their ground electronic states are shown in Fig. 1. Local minimum energy structures were confirmed by the absence of any imaginary frequency in the Hessian matrix. As seen in Fig. 1, the oxygen atoms in the optimized structures are pointing upward from the peptide rings. At the CAM-B3LYP level of theory, the calculated bond lengths of C[double bond, length as m-dash]O and C–N bonds in the amide group of CyAla3 are 1.226 and 1.369 Å, respectively. These values for the CyAla4 ring are 1.224 and 1.359 Å, respectively. It is notable that the calculated C[double bond, length as m-dash]O and C–N bond lengths are same in each free cyclic peptide. It is seen that the value of C–N bond length is sensitive to the ring size. In addition, with calculation at the B3LYP level, the calculated bond lengths of C[double bond, length as m-dash]O and C–N bonds in the amide group of CyAla3 are 1.232 and 1.375 Å, respectively. These values for the CyAla4 ring are 1.229 and 1.365 Å, respectively. The fully relaxed minimum energy structures of the metal ion-cyclic peptide complexes calculated at the CAM-B3LYP/6-31+G(d) and B3LYP/6-31+G(d) levels of the theory are given in Fig. 2, 3 and Fig. S2, S3, respectively.
image file: c4ra08302d-f1.tif
Fig. 1 The optimized structures of CyAla3 and CyAla4 obtained at the CAM-B3LYP/6-31+G(d) level of theory (H atoms are omitted for clarity). All CO and CN bonds are identical for each free cyclic peptide.

image file: c4ra08302d-f2.tif
Fig. 2 Optimized structures and important geometrical parameters of M/CyAla3 complexes calculated at the CAM-B3LYP level of theory, M = Be2+, Mg2+, Ca2+, Sr2+ and Ba2+.

image file: c4ra08302d-f3.tif
Fig. 3 Optimized structures and important geometrical parameters of M/CyAla4 complexes calculated at the CAM-B3LYP level of theory, M = Be2+, Mg2+, Ca2+, Sr2+ and Ba2+.

Moreover, it is evident from Fig. 2, 3, S2, and S3 that the metal cations are not located in the hollow of cyclic peptides and form stable complex with three or four O atoms of the cyclic peptide backbones, where these atoms point upward to the metal ion as seen in the case of the free CyAla3 or CyAla4 molecules. As one can see, comparison of the calculated structural parameters of metal complexes with free cyclic peptides indicates that the C[double bond, length as m-dash]O bond lengths increase and the C([double bond, length as m-dash]O)–N bond lengths decrease in M2+/CyAla3 (see Fig. 2). It must be noted that in the free CyAla3, all carbonyl bond lengths are identical. The analysis of metal–ligand distances may be valuable. All of the alkaline earth metal cations symmetrically interact with oxygen electron lone pairs so that the calculated M–O bond lengths are 1.606, 1.994, 2.346, 2.548, 2.746 Å for the Be2+, Mg2+, Ca2+, Sr2+ and Ba2+, respectively. It is obvious that the M–O bond length decreases with the decrease in the size of metal cation. The selected calculated important geometrical parameters of complexes of alkaline metal ions with CyAla3 molecule calculated at CAM-B3LYP/6-31+G(d) levels of the theory have been tabulated in Table 1. Comparing the geometries of the free cyclic peptide CyAla3 molecule with the corresponding cationic metal complexes indicates that the C–C(H3) bond lengths decrease in the range of 0.020–0.008 from the Be2+/CyAla3 at top of the alkaline earth metal group to Ba2+/CyAla3 at the end of the group but the C–C([double bond, length as m-dash]O) bond lengths decrease from 0.004 Å for Be2+/CyAla3 and 0.001 Å for Mg2+/CyAla3, but for Ca2+ to Ba2+ the C–C([double bond, length as m-dash]O) bond lengths increase about 0.001–0.002 Å. In addition, with the complex formation, the NH bond length increases in the range of 0.003–0.007 Å. For more investigation for the effect of complexation on the geometry of cyclic peptide, the dihedral angle between the carbonyl groups and N–H bonds was determined. As seen, aggregation causes non-negligible changes in the value of dihedral angles. For all φ (H–N–C–O) dihedral angles, Be2+/CyAla3 has maximum dihedral angle and from top to end of the group dihedral angle is increase. For example, the value of φ (H8–N7–C3–O5) dihedral angle changes from −4.4 to 24.0, 20.0, 18.6, 17.6, 16.5 degrees after complexation of CyAla3 with metal cations including Be2+, Mg2+, Ca2+, Sr2+, and Ba2+ metal cations, respectively.

Table 1 The selected geometrical parameters of M/CyAla3 complexes calculated at the CAM-B3LYP/6-31+G(d) level of theory
M/CyAla3 C–C([double bond, length as m-dash]O) C–C(H3) N–H φ 8-7-3-5 φ 14-12-10-13 φ 6-1-16-18 C[double bond, length as m-dash]O C([double bond, length as m-dash]O)–N
CyAla3 1.538 1.528 1.014 −4.4 −4.5 −4.4 1.226 1.369
Be2+ 1.534 1.516 1.021 24.0 23.9 23.9 1.277 1.350
Mg2+ 1.537 1.517 1.020 20.2 20.2 20.2 1.263 1.355
Ca2+ 1.539 1.518 1.018 18.6 18.7 18.6 1.252 1.361
Sr2+ 1.539 1.519 1.017 17.6 17.6 17.5 1.248 1.364
Ba2+ 1.540 1.520 1.017 16.5 16.5 16.6 1.246 1.365


Bond lengths are in Å, dihedral angles in degree. Because of symmetry of free molecule and corresponding complexes only one bond length is presented in table.

The important geometrical parameters of the “host” ligand constructed from four alanine molecule (CyAla4) and its “host–guest” complexes with Be2+, Mg2+, Ca2+, Sr2+, and Ba2+ ions calculated at the same level of theory than for CyAla3 are presented in Table 2. It is noteworthy that in the formation of Be2+ and Mg2+ complexes, only two alanine carbonyl oxygen atoms interact with the metal ions as seen in Fig. 3. The calculated distances between two carbonyl oxygen atoms nearby Be2+ and Mg2+ ion in the upward cavity are 1.577 and 1.980 Å for Be2+/CyAla4 and Mg2+/CyAla4 complexes, respectively. In addition, the distance between amide nitrogens and mentioned ions are 1.917 and 2.333 Å, respectively. Fig. 3 also shows that four oxygen atoms of CyAla4 are interacting with Ca2+, Sr2+, and Ba2+ metal ions. The average bond length for the Ca–O and Sr–O bond is 2.427 Å and 2.625 Å, respectively.

Table 2 The selected geometrical parameters of M/CyAla4 complexes calculated at the CAM-B3LYP level of theorya
M/CyAla4 C–C([double bond, length as m-dash]O) C–(CH3) N–H φ 12-2-1-3 φ 11-6-7-8 φ 14-13-16-17 φ 20-19-21-22 C[double bond, length as m-dash]O C([double bond, length as m-dash]O)–N
a Bond lengths in Å, dihedral angles in degree, a the average value, b the values in parenthesis are the distance between metal ions and faraway atoms.
CyAla4 1.537 a 1.529a 1.017a 6.6 4. 3 6.6 4.3 1.224 1.359
Be2+ (1.547)b, 1.523 1.530a 1.021a −15.1 119.5 −15.1 120.1 (1.191)b, 1.280 (1.495)b, 1.320
Mg2+ (1.549)b, 1.525 1.527a 1.024a −12.1 4.3 −12.1 4.3 (1.195)b, 1.264 (1.475)b, 1.330
Ca2+ 1.537 1.532 1.020 −14.5 −14.5 −14.5 −14.5 1.244 1.361
Sr2+ 1.538 1.532 1.020 −12.9 −12.9 −12.9 −12.9 1.242 1.361
Ba2+ 1.538 1.532 1.019 −11.5 −11.5 −11.5 −11.5 1.240 1.361


Comparison of the M/CyAla4 complexes with the corresponding free cyclic peptide molecule indicates that the C–C([double bond, length as m-dash]O) bond lengths in Be2+/CyAla4 and Mg2+/CyAla4 are different from the corresponding C–C([double bond, length as m-dash]O) bond length in free cyclic peptide (0.014–0.012 Å shift for C–C([double bond, length as m-dash]O) that is bound to metal ions) while for the larger metal ions the C–C([double bond, length as m-dash]O) bond lengths in the cationic metal complexes are almost unchanged.

Only in Be2+/CyAla4 and Mg2+/CyAla4 because of different geometry we use average values for some geometrical parameters such as C–(CH3) and N–H, but because of symmetry for other cationic metal complexes other geometrical parameters are similar. It is noted that because of the importance of C–C([double bond, length as m-dash]O), C[double bond, length as m-dash]O and C([double bond, length as m-dash]O)–N bond lengths in Be2+/CyAla4 and Mg2+/CyAla4 we have maintained the bond lengths values for these parameters.

Similar results have been obtained with B3LYP. The selected important calculated geometrical parameters of complexes of alkaline metal ions with CyAla3 molecule calculated at B3LYP/6-31+G(d) levels of the theory have been tabulated in Table S1 and structures are shown in Fig. S2.

All of the alkaline earth metals cations are approximately symmetrically interact with oxygen lone electron pairs so that the calculated M–O bond lengths are 1.612, 2.008, 2.363, 2.567, 2.768 Å for the Be2+, Mg2+, Ca2+, Sr2+ and Ba2+, respectively. The results of geometrical parameters of the metal complexes of [M/CyAla4]2+ that are similar to result of calculation with CAM-B3LYP method are presented in Fig. S3. In the formation of Be2+ and Mg2+ complexes, only two alanine carbonyl oxygen atoms interact with the metal ions. The calculated distances between two carbonyl oxygen atoms nearby Be2+ and Mg2+ ion in the upward cavity are 1.585 and 1.993 Å for the Be2+/CyAla4 and Mg2+/CyAla4 complexes, respectively. In addition, the distance between amide nitrogens and mentioned ions are 1.931, 1.937 and 2.360, 2.358 Å, respectively. Fig. S3 also shows that four oxygen atoms of CyAla4 are interacting with Ca2+, Sr2+, and Ba2+ metal ions. The bond length for the Ca–O, Sr–O and Ba–O bond is 2.451 Å, 2.653 Å and 2.846 Å respectively.

Comparison of M/CyAla4 complexes with the corresponding free cyclic peptide molecule indicates that the C–C([double bond, length as m-dash]O) bond lengths in Be2+/CyAla4 and Mg2+/CyAla4 are different from the corresponding C–C([double bond, length as m-dash]O) bond in free cyclic peptide (0.026–0.015 Å for C–C([double bond, length as m-dash]O) that is bond to metal ions) while for larger metal ions the C–C([double bond, length as m-dash]O) bond lengths in the cationic metal complexes are almost unchanged. The rest of important geometrical parameters listed in Table S1.

For more precise evaluation of the correct position of metal cations in the cavity of cyclic peptides, an exploration of the PES has been performed. For this purpose, the position of relevant metal cations was changed from the cavity center of cyclic peptide by 0.2 Å intervals. The graphical illustration of the energy changes occurring during the inclusion passing process of cations at different Z positions of cyclic peptides presented in Fig. 4. A closer look at this figure and comparison with geometrical parameters allows one remarking that inclusion process is thermodynamically favorable. Interestingly, a local minimum found for each cation about 6.4 Å above the cavity center. It is interesting that for the Be cation, a considerable local minimum related to method of calculation (CAM-B3LYP level) at 6.2 Å, found.


image file: c4ra08302d-f4.tif
Fig. 4 Potential energy surfaces of inclusion complexation of metal cations in the cavity of CyAla3 at different positions, calculated at B3LYP/6-31+G(d) and CAM-B3LYP levels of theory.

3.2. Binding energies and metal binding selectivity

Generally, the energy decreases with creation of a host–guest complex. The decreased energy is called the binding energy (BE), which is associated with the solidity of the equivalent host–guest complex and the extraction power of an extractant for a given metal ion. A steady complex all the time gives a negative value of ΔE. Therefore, the stability of complexes will increase with the negative value of ΔE, and the extraction power of an extractant for metal ions will be stronger. The BE of M/CyAla3 or M/CyAla4 complexes for the complexation reaction:
 
M2+ + CyAla3 → M/CyAla3 (1)
is defined by the following general equation:
 
BE = EM/CyAla2+ − (EM2+ + ECyAla) (2)
where EM/CyAla2+, EM2+, and ECyAla refer to the energy of the M2+/CyAla complex, M2+ ion and the cyclic peptide system, respectively. The calculated binding energies using B3LYP and CAM-B3LYP methods are listed in Table 3. The results clearly show the effect of the metal ion's nature on the selective binding capacity.
Table 3 The binding energies ΔE (kcal mol−1), zero point corrected binding energies ΔEZPE = (ΔE + ΔZPE), the value of basis set superposition error in energy (EBSSE), ΔEcorr = ΔEZPE + EBSSE, binding enthalpies, Gibbs free energies ΔG of binding (kcal mol−1) and formation equilibrium constants in gas phase for the complexes calculated at B3LYP and CAM-B3LYP levels of theorya
CAM-B3LYP ΔE ΔEZPE EBSSE ΔEcorr ΔH ΔG log[thin space (1/6-em)]K
a ΔE and ΔG in kcal mol−1.
M/CyAla3
Be2+ −327.09 −324.14 1.25 −322.89 −326.47 −311.86 228.64
Mg2+ −205.23 −203.37 1.34 −202.03 −205.07 −190.98 140.01
Ca2+ −139.85 −138.38 1.20 −137.18 −139.71 −126.25 92.56
Sr2+ −106.58 −105.39 1.47 −103.92 −106.44 −93.46 68.52
Ba2+ −90.17 −89.10 1.38 −87.72 −90.01 −77.40 56.74
[thin space (1/6-em)]
M/CyAla4
Be2+ −359.83 −356.61 1.78 −354.83 −358.64 −345.92 253.60
Mg2+ −207.52 −205.61 1.99 −203.62 −206.97 −195.07 143.01
Ca2+ −151.59 −149.89 1.40 −148.49 −151.18 −138.33 101.42
Sr2+ −116.42 −114.97 1.77 113.20 −115.98 −103.61 75.96
Ba2+ −99.49 −98.23 1.75 −96.48 −99.06 −87.17 63.91
[thin space (1/6-em)]
B3LYP
M/CyAla3
Be2+ −322.65 −319.82 1.19 −318.63 −321.51 −309.53 226.93
Mg2+ −200.63 −198.82 1.26 −197.56 −199.91 −188.40 138.12
Ca2+ −135.48 −134.01 1.17 −132.84 −134.72 −123.89 90.83
Sr2+ −102.51 −101.24 1.42 −99.82 −101.70 −91.28 66.92
Ba2+ −86.29 −85.19 1.36 −83.83 −85.47 −75.53 55.37
[thin space (1/6-em)]
M/CyAla4
Be2+ −354.26 −351.17 1.79 −349.38 −353.13 −340.49 249.62
Mg2+ −203.24 −201.51 1.88 −199.63 −202.80 −191.00 140.03
Ca2+ −146.32 −144.73 1.28 −143.45 −145.97 −133.18 97.64
Sr2+ −111.59 −110.18 1.69 −108.49 −111.11 −98.96 72.55
Ba2+ −95.03 −93.77 1.74 −92.03 −94.59 −82.74 60.66


The order of binding energies are Be2+ > Mg2+ > Ca2+ > Sr2+ > Ba2+ for both M/CyAla3 and M/CyAla4. The binding energies were also corrected for ZPE and BSSE corrections (ΔEZPE and ΔEcorr in Table 3). The binding enthalpy (ΔH) and binding free energy (ΔG) for the metal cyclic peptide complexation reactions were also calculated at the CAM-B3LYP and B3LYP levels at 298 K and the results have been listed in Table 3. It is obvious that the formation of metal ion complexes is exothermic as revealed from the values of ΔH given in Table 3. The binding enthalpy is increased in the order of Be2+ > Mg2+ > Ca2+ > Sr2+ > Ba2+ for both B3LYP and CAM-B3LYP methods. It is notable that with respect to the same cation, the larger cavity CyAla4 ligand, can hold the alkaline metal cations better than CyAla3 molecule. One can evaluate the regularity of B3LYP and CAM-B3LYP methods with looking to Fig. 5 which demonstrates a good correlation between the two levels of theory.


image file: c4ra08302d-f5.tif
Fig. 5 Correlation between B3LYP (red curve) and CAM-B3LYP (blue curve) binding energies for M/CyAla3 and M/CyAla4 complexes.

The calculated binding energies for the earth alkaline metal cations are much higher than with our previously reported data for the complexation of alkali metal cations and above mentioned cyclic peptides.44 The range of binding energies for the alkali metal cations has been found to be 47.11–15.75 and 31.19–15.24 kcal mol−1 for the M/CyAla3 and M/CyAla4, respectively. As it can be seen, alkaline earth metal cations form more tighten complexes with cyclic peptides.

3.3. Second-order interaction energies, energy gaps and charge transfers

To find the origin of the favorable interaction energies and clarify the reason for the different metal binding selectivity, the NBO analysis was carried out in this work. In NBO analysis, the stabilization energy (E(2)) is related to the strength of the coordination interaction. There is a direct relationship between the stability of complex and (E(2)) so that the more stability of complex is corresponds to the larger value for E(2). The stabilization energy E(2), associated with ij delocalization, can be estimated by the following equation:
image file: c4ra08302d-t1.tif
where qi is the donor orbital occupancy, εi and εj are orbital energies of donor and acceptor orbitals, respectively. F(i,j) are off-diagonal elements associated with NBO Fock matrix. The values of E(2), obtained by NBO analysis, for the considered complexes are summarized in Table 4. The value of stabilization energy depends on the strength of the charge-transfer interaction between a Lewis type NBOs (donor) and non-Lewis NBOs (acceptor). The stronger donor → acceptor interaction leads to the higher value for the relevant stabilization energy, and more charge will be transferred from the donor (cyclic peptide) to the acceptor (metal ion). Overall, the results of NBO analysis indicate that the origin of the interactions between the metal cations and the electron-donating oxygen or nitrogen atoms in the considered cyclic peptides are electrostatic. It is notable that for the M/CyAla3 complexes, with going from Be2+ as the smallest cation to Ba2+ as the largest one, the electron donation of oxygen atoms decreases from 17.24 to 1.38 kcal mol−1, respectively. However for the M/CyAla4 complexes when Mg2+ located in the cavity of ligand, the lone electron pairs of both N and O atoms contribute in stabilizing the cation. In addition, the values of E(2) for the M/CyAla3 complexes are lower than those for the M/CyAla4 when the same alkali metal cation contribute in the complexation.
Table 4 Selected stabilization interaction E(2) (kcal mol−1) for M/CyAla3 and M/CyAla4 complexes at the CAM-B3LYP level of theorya
M/CyAla3 M/CyAla4
Donor Acceptor   Donor Acceptor  
a LP, 1-center valence lone pair (LP1 and LP2 are the tow lone pairs of each oxygen and nitrogen atoms, respectively. One of the NBO is in the plane, the other is the corresponding NBO perpendicular to the plane): LP*, 1-center valance antibond lone pair: BD, 2-center bond. RY* corresponds to Rydberg NBOs.
Be2+ Be2+
BD(1)C3–O5 LP*(1)Be31 4.99 BD(1)O3–Be41 BD*(2)C1–N2 88.55
BD(2)C3–O5 LP*(1)Be31 10.71 BD(1)O17–Be41 BD*(2)N13–C16 88.41
BD(2)C3–O5 RY*(1)Be31 3.55 LP(1)O3 BD*(1)O3–Be41 8.09
BD(1)C10–O13 LP*(1)Be31 4.87 LP(1)O3 BD*(1)O17–Be41 6.44
BD(1)C16–O18 LP*(1)Be31 4.99 LP(2)O3 BD*(1)O3–Be41 6.81
BD(2)C16–O18 LP*(1)Be31 10.69 LP(1)N6 BD*(1)O17–Be41 6.83
BD(2)C16–O18 RY*(2)Be31 4.33 LP(1)O17 BD*(1)O3–Be41 6.45
LP(1)O5 LP*(1)Be31 17.24 LP(1)O17 BD*(1)O17–Be41 8.20
LP(2)O5 LP*(1)Be31 22.63 LP(2)O17 BD*(1)O17–Be41 6.61
LP(1)O13 LP*(1)Be31 17.12 LP(1)N19 BD*(1)O3–Be41 6.77
LP(2)O13 LP*(1)Be31 22.42 LP(1)N19 BD*(1)O17–Be41 5.28
LP(3)O13 LP*(1)Be31 14.90      
LP(1)O18 LP*(1)Be31 17.24      
LP(2)O18 LP*(1)Be31 22.63      
[thin space (1/6-em)]
Mg2+ Mg2+
LP(2)O5 LP*(1)Mg31 4.67 BD(2)C1–O3 LP*(1)Mg41 8.28
LP(1)O13 LP*(1)Mg31 10.23 LP(1)O3 LP*(1)Mg41 11.73
LP(2)O13 LP*(1)Mg31 4.64 LP(1)N6 LP*(1)Mg41 14.99
LP(1)O18 LP*(1)Mg31 10.22 LP(1)O17 LP*(1)Mg41 11.64
LP(2)O18 LP*(1)Mg31 4.66 LP(3)O17 LP*(1)Mg41 10.29
      LP(1)N19 LP*(1)Mg41 15.01
[thin space (1/6-em)]
Ca2+ Ca2+
LP(1)O5 LP*(1)Ca31 4.22 LP(1)O3 LP*(1)Ca41 4.94
LP(1)O13 LP*(1)Ca31 4.21 LP(1)O8 LP*(1)Ca41 4.93
LP(1)O18 LP*(1)Ca31 4.21 LP(1)O17 LP*(1)Ca41 4.94
      LP(1)O22 LP*(1)Ca41 4.94
[thin space (1/6-em)]
Sr2+ Sr2+
LP(1)O5 LP*(1)Sr31 2.90 LP(1)O3 LP*(3)Sr41 7.78
LP(2)O5 RY*(2)Sr31 1.02 LP(1)O8 LP*(1)Sr41 5.33
LP(1)O13 LP*(1)Sr31 2.92 LP(1)O8 LP*(2)Sr41 7.78
LP(1)O18 LP*(1)Sr31 2.92 LP(1)O17 LP*(1)Sr41 5.34
LP(2)O18 RY*(1)Sr31 1.05 LP(1)O17 LP*(3)Sr41 7.78
      LP(1)O22 LP*(1)Sr41 5.34
      LP(1)O22 LP*(2)Sr41 7.78
[thin space (1/6-em)]
Ba2+ Ba2+
LP(1)O5 LP*(1)Ba31 1.38 LP(1)O3 LP*(1)Ba41 1.81
LP(1)O13 LP*(1)Ba31 1.38 LP(1)O8 LP*(1)Ba41 1.81
LP(2)O13 RY*(2)Ba31 1.00 LP(1)O17 LP*(1)Ba41 1.81
LP(1)O18 LP*(1)Ba31 1.39 LP(1)O22 LP*(1)Ba41 1.81


In order to analyze the electrostatic interactions of the alkaline metal cations with the host molecules, the partial charges of the selected atoms in the complexes compared with the corresponding charges in the free ligand molecules, (see Table 5). It is well known that the complexation of metal ions and peptides can proceed through the electrostatic effects taking place between metal ions with main chain carbonyl groups or side chains groups. However, molecular modeling and experimental results suggested the preference of the interaction of backbone carbonyl groups of cyclic peptides with the metal ions inside the cavity.75–77 For the present complexes, the charge-transfer is defined as the charge difference between a free metal ion and its complexated form.

Table 5 Calculated NBO charges of the metals and selected atoms of M/CyAla3 and M/CyAla4 complexes at CAM-B3LYP level of theory
CyAla3   Be2+ Mg2+ Ca2+ Sr2+ Ba2+
N1 −0.671 −0.609 −0.619 −0.632 −0.636 −0.640
N7 −0.671 −0.609 −0.619 −0.632 −0.637 −0.639
N12 −0.671 −0.609 −0.619 −0.632 −0.637 −0.640
O5 −0.627 −0.837 −0.809 −0.787 −0.770 −0.759
O13 −0.627 −0.837 −0.809 −0.786 −0.770 −0.758
O18 −0.627 −0.837 −0.809 −0.787 −0.771 −0.759
M   1.744 1.856 1.942 1.918 1.931
[thin space (1/6-em)]
CyAla4
N2 −0.651 −0.579 −0.597 −0.646 −0.645 −0.649
N6 −0.654 −0.914 −0.871 −0.646 −0.645 −0.649
N13 −0.651 −0.579 −0.597 −0.646 −0.645 −0.649
N19 −0.654 −0.913 −0.871 −0.646 −0.645 −0.649
O3 −0.635 −0.831 −0.793 −0.752 −0.716 −0.736
O8 −0.635 −0.461 −0.453 −0.751 −0.716 −0.736
O17 −0.635 −0.831 −0.793 −0.752 −0.716 −0.736
O22 −0.635 −0.461 −0.453 −0.752 −0.716 −0.736
M   1.697 1.805 1.932 1.763 1.932


Table 5 also reveals that for the M/CyAla3, the negative charge on the nitrogen atoms changed from −0.671 esu in the free CyAla3 molecule to −0.609, −0.619, −0.632, −0.636, and −0.640 esu for the Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ metal cations, respectively. Moreover, the negative charge on the oxygen atoms increases from −0.627 esu in the free cyclic peptide molecule to −0.837, −0.809, −0.787, −0.770, and −0.759 esu for the Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ metal cations located in the cavity of cyclic peptide, respectively. The charge transfer values for the metal ions in the complex are 0.256, 0.144, 0.038, 0.082 and 0.069 esu, calculated for the Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ metal cations, respectively. For the Be2+ cation that interacts with contrary nitrogen atoms, N2 and N13 atoms take −0.651 but N6 and N19 atoms take −0.654 esu for the Be2+/CyAla4, respectively. In addition, for the Mg2+/CyAla4 complex the N2 and N13 atoms take −0.579 esu but, N6 and N19 atoms take −0.914 esu, respectively. As you can see the charge transfer values are in accordance with the radii of metal cations.

3.4. AIM topological parameters

The theory of atoms in molecules (AIM) was developed by Professor Richard F. W. Bader and his co workers, in 1990.72,73 AIM characterizes the chemical bonding of a system based on the topology of the quantum charge density. The bond critical point (BCP) is described in terms of topological parameters, such as the charge density and the corresponding Laplacian field. According to the topological analysis of electronic charge density in AIM theory, electronic charge density ρ(r) describes the strength of a bond (if ρ(r) value is big, the corresponding bond will be strong), and Laplacian of the electron charge density ∇2ρ(r) (The sum of eigenvalues (λ1, λ2 and λ3) of the Hessian matrix of electronic charge density is equal to the Laplacian) shows the characteristic of the bond. A negative value of Laplacian ∇2ρ(r) < 0 indicates the concentration of the electron density in the interatomic region and occurs for sharing interactions like covalent bonds whereas a positive value (∇2ρ(r) > 0) of Laplacian indicates the depletion of the electron density for the interaction of the closed-shell systems such as ionic bond, coordination bond, hydrogen bond, or van der Waals interaction. The values of ρ(r), ∇2ρ(r), the eigenvalues of the Hessian matrix and the ellipticity at the bond critical point of all studied complexes, calculated using the wave function obtained at the CAM-B3LYP/6-31+G(d) level of theory were listed in Table 6. As an example the molecular graphs of the Ca2+/CyAla3 and Ca2+/CyAla4 based on AIM theory are shown in Fig. 6, because the other M2+/CyAla3 and M2+/CyAla4 have similar shape we just show one graph. Based on Table 6, the calculated Laplacian values at corresponding BCP are positive and this means that interactions between cyclic peptide and alkaline earth metal cations were closed-shell interactions and there is no bond between them. On the other hand, the values of ∇2ρ(r) changed in the order of Be2+ > Mg2+ > Ca2+ > Sr2+ > Ba2+ for both M/CyAla3 and M/CyAla4 complexes. For all complexes except for Be/CyAla4 and Mg/CyAla4, the alkaline earth metal cations interact with O atoms of cyclic peptides and as it is obvious from Fig. 3 the shape of these complexes are symmetrical so the interactions between alkaline earth metal cations and all of the O atoms in one complex are nearly the same. The ρ(r) values show that when we go from top to end of the alkaline earth metal group, electronic charge density decreases so the interactions between cyclic peptide and alkaline earth metal cations weakens.
Table 6 The topological properties at BCP of complexesa
CyAla3 BCPs λ1 λ2 λ3 ρ(r) 2ρ(r)
a (ρ(r) in e au−3, ∇2ρ(r) in e au−5).
Be O5–M −0.1755 −0.1630 0.8959 0.0787 0.5574
O13–M −0.1754 −0.1629 0.8952 0.0786 0.5569
O18–M −0.1755 −0.1630 0.8961 0.0787 0.5575
Mg O5–M −0.0642 −0.0605 0.4511 0.0432 0.3264
O13–M −0.0642 −0.0605 0.4510 0.0431 0.3264
O18–M −0.0641 −0.0605 0.4509 0.0431 0.3262
Ca O5–M −0.0406 −0.0376 0.2676 0.0341 0.1894
O13–M −0.0405 −0.0375 0.2668 0.0341 0.1888
O18–M −0.0405 −0.0375 0.2669 0.0341 0.1889
Sr O5–M −0.0259 −0.0255 0.1914 0.0245 0.1400
O13–M −0.0261 −0.0256 0.1925 0.0246 0.1409
O18–M −0.0262 −0.0257 0.1932 0.0247 0.1414
Ba O5–M −0.0225 −0.0207 0.1652 0.0230 0.1219
O13–M −0.0224 −0.0206 0.1641 0.0229 0.1212
O18–M −0.0225 −0.0208 0.1653 0.0231 0.1221
[thin space (1/6-em)]
CyAla4
Be O3–M −0.2015 −0.1923 1.0215 0.0872 0.6277
O17–M −0.2016 −0.1924 1.0221 0.0873 0.6281
N6–M −0.0676 −0.0520 0.3339 0.0461 0.2143
N19–M −0.0657 −0.0501 0.3259 0.0455 0.2101
Mg O3–M −0.0688 −0.0653 0.4835 0.0463 0.3494
O17–M −0.0689 −0.0653 0.4838 0.0463 0.3496
N6–M −0.0292 −0.0244 0.1787 0.0261 0.1251
N19–M −0.0293 −0.0245 0.1793 0.0261 0.1255
Ca O3–M −0.0314 −0.0293 0.2076 0.0280 0.1469
O8–M −0.0313 −0.0293 0.2072 0.0280 0.1466
O17–M −0.0313 −0.0293 0.2075 0.0280 0.1468
O22–M −0.0314 −0.0293 0.2076 0.0280 0.1469
Sr O3–M −0.0206 −0.0204 0.1508 0.0205 0.1098
O8–M −0.0206 −0.0204 0.1507 0.0205 0.1097
O17–M −0.0206 −0.0204 0.1508 0.0205 0.1098
O22–M −0.0206 −0.0204 0.1507 0.0205 0.1098
Ba O3–M −0.0183 −0.0169 0.1358 0.0196 0.1005
O8–M −0.0183 −0.0169 0.1357 0.0195 0.1005
O17–M −0.0182 −0.0169 0.1356 0.0195 0.1004
O22–M −0.0183 −0.0169 0.1357 0.0195 0.1004



image file: c4ra08302d-f6.tif
Fig. 6 Molecular graphs of the Ca2+/CyAla3 and Ca2+/CyAla4 complexes at the CAM-B3LYP/6-31+G(d) level of theory.

3.5. The HOMO and LUMO energies and the values of the energy gap

In this part we examine the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of our complexes. HOMO and LUMO of molecules are quite essential to describe their reactivity. EHOMO depicts the molecular ability in donating electrons to appropriate acceptor molecules with low energy having empty molecular orbital. In contrast ELUMO indicates the ability of the molecule to accept electrons. The lower value of ELUMO, indicates that the molecule would accept electrons. Therefore, relating to the value of the energy gap, ΔE(ELUMOEHOMO), if this energy is high it means that the reactivity to a molecule is low on the contrary if the energy gap is low the reactivity to a molecule is high because the energy needed to promote one electron from the HOMO to the LOMO orbital will be low. The diagrams of frontier molecular orbital and the energies of Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ complexes with CyAla3 and CyAla4 such as ELUMO, EHOMO, and ΔE (in eV) estimated by the CAM-B3LYP/6-31+G(d) level are represented in Table 7 and Fig. 7. As one can see, for the CyAla3 and CyAla4 complexes, the HOMOs locate on heteroatoms of cyclic peptides. On the other hand the LUMOs show different patterns. For instance for the Be/CyAla3 or Be/CyAla4 complexes the LUMOs locate partly on the cation, for the other complexes the LUMOs extend on the whole molecule.
Table 7 Frontier molecular orbital diagrams and energies (eV) of Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ complexes with CyAla3 and CyAla4 estimated by the CAM-B3LYP/6-31+G(d) level
M/CyAla3 EHOMO ELUMO ΔE M/CyAla4 EHOMO ELUMO ΔE
Be −0.669 −0.428 0.241 Be −0.610 −0.503 0.107
Mg −0.666 −0.386 0.280 Mg −0.596 −0.483 0.113
Ca −0.661 −0.357 0.304 Ca −0.587 −0.386 0.201
Sr −0.605 −0.312 0.293 Sr −0.573 −0.360 0.213
Ba −0.607 −0.309 0.298 Ba −0.579 −0.307 0.272



image file: c4ra08302d-f7.tif
Fig. 7 Frontier molecular orbital diagrams of Be2+, Mg2+, Ca2+, Sr2+ and Ba2+ from top to bottom complexes with CyAla3 (left) and CyAla4 (right) calculated at the CAM-B3LYP/6-31+G(d) level.

As seen in Table 7 the values of EHOMO for M/CyAla3 complexes show the ranking Be2+ > Mg2+ > Ca2+ > Ba2+ > Sr2+ for this property. In addition, the values of ΔE show Ca2+ > Ba2+ > Sr2+ > Mg2+ > Be2+. For M/CyAla4 complexes the order is Be2+ > Mg2+ > Ca2+ > Ba2+ > Sr2+ for the EHOMO and the values of ΔE show Ba2+ > Sr2+ > Ca2+ > Mg2+ > Be2+. As can be seen from the HOMO and LUMO pictures in Fig. 7, the majority of HOMO and LUMO's are found on the donor atoms in the cyclic peptide.

4. Conclusion

The structure and interaction energies of nanotubular cyclic peptide complexes of M2+/CyAla3 and M2+/CyAla4, where M = Be2+, Mg2+, Ca2+, Sr2+, and Ba2+ have been studied using B3LYP and CAM-B3LYP method. The key findings are as follows:

(1) Analyzing the geometry of M/CyAla3 and M/CyAla4 complexes indicated that the aggregation caused substantial changes in geometrical parameters of ligands. In this manner, after insertion the metal ions in the cavity of cyclic peptides, the C[double bond, length as m-dash]O bond length increases in the range of 0.026–0.051 Å for the M/CyAla3 complexes while the C–N amide bond length decreased in the range of 0.004–0.019 Å. In addition, for the M/CyAla4 complexes, during the formation of Be2+ and Mg2+ complexes, only two alanine carbonyl oxygen atoms interact with the metal ions. Moreover, the calculated metal ligand bond lengths decrease with decreasing size of metal cation.

(2) Vibrational frequency calculations showed that these cyclic peptides and their complexes with the alkaline earth metal cations are all located at local minimum points of their potential energy surfaces. Therefore, they are all stable “host–guest” complexes.

(3) The order of binding energies calculated by B3LYP and CAM-B3LYP methods was found to be Be2+ > Mg2+ > Ca2+ > Sr2+ > Ba2+ for both M/CyAla3 and M/CyAla4, respectively. This trend indicates that these cyclic peptides might be used for separating agent of these cations.

(4) Based on the charges obtained by the NBO analysis, it can be concluded that the binding energies may be attributed to the strong polarization of the C[double bond, length as m-dash]O bonds of cyclic peptides by metal cations.

(5) Based on AIM calculations, Laplacian values at corresponding BCP are positive and this means that the interactions between cyclic peptide and alkaline earth metal cations are closed-shell interactions and there is no bond between them.

(6) The results of this study are comparable with our previous work. The results indicate that alkaline earth metal cations bind with much more strength than alkali metal cations. This could be due to the double positive charge of the former ions, compared to the single charge of alkali metal cations.

Acknowledgements

We would like to thank Isfahan University of Technology (IUT) for the financial support (Research Council Grant).

References

  1. J. M. Lehn, Pure Appl. Chem., 1978, 50, 871–892 CrossRef CAS.
  2. M. Vincenti, J. Mass Spectrom., 1995, 30, 925–939 CrossRef CAS.
  3. C. J. Pedersen, J. Am. Chem. Soc., 1967, 89, 2495–2496 CrossRef CAS.
  4. J. M. Lehn, J. P. Sauvage and B. Diedrich, J. Am. Chem. Soc., 1970, 92, 2916–2918 CrossRef CAS.
  5. R. Moran, S. Karbach and D. J. Cram, J. Am. Chem. Soc., 1982, 104, 5826–5828 CrossRef.
  6. D. J. Cram, S. Karbach, Y. H. Kim, L. Baczynskyj and G. W. Kalleymeyn, J. Am. Chem. Soc., 1985, 107, 2575–2576 CrossRef CAS.
  7. L. Seridi and A. Boufelfel, J. Mol. Liq., 2011, 158, 151–158 CrossRef CAS PubMed.
  8. S. Fanali, J. Chromatogr. A, 2000, 875, 89–122 CrossRef CAS.
  9. D. W. Armstrong and U. B. Nair, Electrophoresis, 1997, 18, 2331–2342 CrossRef CAS PubMed.
  10. T. J. Ward and T. M. Oswald, J. Chromatogr. A, 1997, 792, 309–325 CrossRef CAS.
  11. J. Haginaka, J. Chromatogr. A, 2000, 875, 235–254 CrossRef CAS.
  12. K. Otsuka and S. Terabe, J. Chromatogr. A, 2000, 875, 163–178 CrossRef CAS.
  13. E. Anslyn, Modern Physical Organic Chemistry, University Science Books, 2006 Search PubMed.
  14. B. De Sousa, A. M. L. Denadai, I. S. Lula, J. F. Lopes, H. F. Dos Santos, W. B. De Almeida and R. D. Sinisterra, Int. J. Pharm., 2008, 353, 160–169 CrossRef PubMed.
  15. J. K. Khedkar, W. Gobre, R. V. Pinjari and S. P. Gejji, J. Phys. Chem. A, 2010, 114, 7725–7732 CrossRef CAS PubMed.
  16. A. Maheshwari and D. Sharma, J. Inclusion Phenom. Macrocyclic Chem., 2010, 68, 453–459 CrossRef CAS.
  17. M. Jug, N. Mennini, F. Melani, F. Maestrelli and P. Mura, Chem. Phys. Lett., 2010, 500, 347–354 CrossRef CAS PubMed.
  18. X. H. Wen, Z. Y. Liu and T. Q. Zhu, Chem. Phys. Lett., 2005, 405, 114–117 CrossRef CAS PubMed.
  19. A. Zoppi, M. A. Quevedo, A. Delrivo and M. R. Longhi, J. Pharm. Sci., 2010, 99, 3166–3176 CAS.
  20. H. F. Dos Santos, H. A. Duarte, R. D. Sinisterra, S. V. De Melo Mattos, L. F. C. De Oliveira and W. B. De Almeida, Chem. Phys. Lett., 2000, 319, 569–575 CrossRef CAS.
  21. W. Snor, E. E. Liedl, P. Weiss Greiler, H. Virnstein and P. Wolschann, Int. J. Pharm., 2009, 381, 146–152 CrossRef CAS PubMed.
  22. D. J. Barbiric, E. A. Castro and R. H. de Rossi, J. Mol. Struct.: THEOCHEM, 2000, 532, 171–181 CrossRef CAS.
  23. H. A. Dabbagh, M. Zamani and H. Farrokhpour, Chem. Phys., 2012, 393, 86–95 CrossRef CAS PubMed.
  24. G. J. Chen, S. Su and R. Z. Liu, J. Phys. Chem. B, 2002, 106, 1570–1575 CrossRef CAS.
  25. M. Teranishi, H. Okamoto, K. Takeda, K. Nomura, A. Nakano, R. K. Kalia, P. Vashishta and F. Shimojo, J. Phys. Chem. B, 2009, 113, 1473–1484 CrossRef CAS PubMed.
  26. J. D. Hartgerink, J. R. Granja, R. A. Milligan and M. R. Ghadiri, J. Am. Chem. Soc., 1996, 118, 43–50 CrossRef CAS.
  27. H. W. Tan, W. W. Qu, G. J. Chen and R. Z. Liu, Chem. Phys. Lett., 2003, 369, 556–562 CrossRef CAS.
  28. R. Poteau and G. Trinquier, J. Am. Chem. Soc., 2005, 127, 13875–13889 CrossRef CAS PubMed.
  29. F. Yokoyama, N. Suzuki, M. Haruki, N. Nishi, S. Oishi, N. Fujii, A. Utani, H. K. Kleinman and M. Nomizu, Biochemical, 2004, 43, 13590–13597 CrossRef CAS PubMed.
  30. M. Bagheri, S. Keller and M. Dathe, Antimicrob. Agents Chemother., 2011, 55, 788–797 CrossRef CAS PubMed.
  31. M. Kracht, H. Rokos, M. Ozel, M. Kowall, G. Pauli and J. Vater, J. Antibiot., 1999, 52, 613–619 CrossRef CAS.
  32. S. R. Tendulkar, Y. K. Saikumari, V. Patel, S. Raghotama, T. K. Munshi, P. Balaram and B. B. Chattoo, J. Appl. Microbiol., 2007, 103, 2331–2339 CrossRef CAS PubMed.
  33. C. Weber, G. Wider, B. von Freyberg, R. Traber, W. Braun, H. Widmer and K. Wuthrich, Biochemical, 1991, 30, 6563–6574 CrossRef CAS.
  34. G. Trevisan, G. Maldaner, N. A. Velloso, S. Sant'Anna Gda, V. Ilha, C. Velho Gewehr Cde, M. A. Rubin, A. F. Morel and J. Ferreira, J. Nat. Prod., 2009, 72, 608–612 CrossRef CAS PubMed.
  35. H. Z. Gang, J. F. Liu and B. Z. Mu, J. Phys. Chem. B, 2010, 114, 2728–2737 CrossRef CAS PubMed.
  36. A. Banerjee and A. Yadav, Appl. Nanosci., 2012, 1–14 Search PubMed.
  37. J. Liu, J. Fan, M. Tang and W. Zhou, J. Phys. Chem. A, 2010, 114, 2376–2383 CrossRef CAS PubMed.
  38. R. Vijayaraj, S. Sundar Raman, R. Mahesh Kumar and V. Subramanian, J. Phys. Chem. B, 2010, 114, 16574–16583 CrossRef CAS PubMed.
  39. R. Vijayaraj, S. Van Damme and P. Bultinck, Phys. Chem. Chem. Phys., 2012, 14, 15135–15144 RSC.
  40. P. Zero, F. Plucinski and A. P. Mazurek, J. Mol. Struct.: THEOCHEM, 2009, 915, 182–189 CrossRef CAS PubMed.
  41. R. A. Jishi, R. M. Flores, M. Valderrama, L. Lou and J. Bragin, J. Phys. Chem. A, 1998, 102, 9858–9862 CrossRef CAS.
  42. H. Zhao, Y. Zhu, M. Tong, J. He, C. Liu and M. Tang, J. Mol. Model., 2012, 18, 851–858 CrossRef CAS PubMed.
  43. X. Chen, M. Tirado, J. D. Steill, J. Oomens and N. C. Polfer, J. Mass Spectrom., 2011, 46, 1011–1015 CrossRef CAS PubMed.
  44. P. Y. Iris Shek, J. Kai-Chi Lau, J. Zhao, J. Grzetic, U. H. Verkerk, J. Oomens, A. C. Hopkinson and K. W. Michael Siu, Int. J. Mass Spectrom., 2012, 316–318, 199–205 CrossRef PubMed.
  45. J. S. Klassen and P. l. Kebarle, J. Am. Chem. Soc., 1997, 119, 6552–6563 CrossRef CAS.
  46. T. Yalcin, C. Khouw, I. G. Csizmadia, M. R. Peterson and A. G. Harrison, J. Am. Soc. Mass Spectrom., 1995, 6, 1165 CrossRef CAS.
  47. T. Yalcin, I. G. Csizmadia, M. B. Peterson and A. G. Harrison, J. Am. Soc. Mass Spectrom., 1996, 7, 233 CrossRef CAS.
  48. S. M. Williams and J. S. Brodbelt, J. Am. Soc. Mass Spectrom., 2004, 15, 1039–1054 CrossRef CAS PubMed.
  49. K. Schwing, C. Reyheller, A. Schaly, S. Kubik and Ma. Gerhards, ChemPhysChem, 2011, 12, 1981–1988 CrossRef CAS PubMed.
  50. L. C. M. Ngoka and M. L. Gross, J. Mass Spectrom., 2000, 35, 265–276 CrossRef CAS.
  51. A. P. Mendham, T. J. Dines, M. J. Snowden, B. Z. Chowdhry and R. Withnall, J. Raman Spectrosc., 2009, 40, 1478–1497 CrossRef CAS.
  52. B. T. Ruotolo, C. C. Tate and D. H. Russell, J. Am. Soc. Mass Spectrom., 2004, 15, 870–878 CrossRef CAS PubMed.
  53. L. C. M. Ngoka and M. L. Gross, Int. J. Mass Spectrom., 2000, 194, 247–259 CrossRef CAS.
  54. C. M. N. Lambert, M. L. Gross and P. L. Toogood, Int. J. Mass Spectrom., 1999, 182/183, 289–298 CrossRef.
  55. S. Lin, S. Liehr, B. S. Cooperman and R. J. Cotter, J. Mass Spectrom., 2001, 36, 658–663 CrossRef CAS PubMed.
  56. N. S. Nagornova, M. Guglielmi, M. Doemer, I. Tavernelli, U. Rothlisberger, T. R. Rizzo and O. V. Boyarkin, Angew. Chem., Int. Ed., 2011, 50, 5383–5386 CrossRef CAS PubMed.
  57. L. Zhang, Z. Luo, L. Zhang, L. Jia and L. Wu, J. Biol. Inorg. Chem., 2013, 18, 277–286 CrossRef CAS PubMed.
  58. M. Ngu-Schwemlein, W. Gilbert, K. Askew and S. Schwemlein, Bioorg. Med. Chem. Lett., 2008, 16, 5778–5787 CrossRef CAS PubMed.
  59. F. Shahangi, A. N. Chermahini, H. A. Dabbagh, A. Teimouri and H. Farrokhpour, Comput. Theor. Chem., 2013, 1020, 163–169 CrossRef CAS PubMed.
  60. A. N. Chermahini, M. Rezapour and A. Teimouri, J. Inclusion Phenom. Macrocyclic Chem., 2014, 79, 205–214 CrossRef PubMed.
  61. M. Kamiya, T. Tsuneda and K. Hirao, J. Chem. Phys., 2002, 117, 6010–6015 CrossRef CAS PubMed.
  62. P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627 CrossRef CAS.
  63. T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS PubMed.
  64. N. C. Polfer, J. Oomens and R. C. Dunbar, ChemPhysChem, 2008, 9, 579–589 CrossRef CAS PubMed.
  65. C. Colas, S. Bouchonnet, F. Rogalewicz-Gilard, M. Popot and G. Ohanessian, J. Phys. Chem. A, 2006, 110, 7503–7508 CrossRef CAS PubMed.
  66. T. Marino, N. Russo and M. Toscano, Inorg. Chem., 2001, 40, 6439–6443 CrossRef CAS PubMed.
  67. A. D. Beck, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef PubMed.
  68. S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553–566 CrossRef CAS.
  69. A. E. Reed, R. B. Weinstock and F. Weinhold, J. Chem. Phys., 1985, 83, 735–746 CrossRef CAS PubMed.
  70. A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899–926 CrossRef CAS.
  71. M. J. Frisch, et al., Gaussian 09, Revision B.01, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.
  72. R. F. W. Bader, International Series of Monographs in Chemistry, Atoms in Molecules, A Quantum Theory, Oxford University Press, Oxford, 1990, vol. 22 Search PubMed.
  73. P. L. A. Popelier, Coord. Chem. Rev., 2000, 197, 169–189 CrossRef CAS.
  74. R. F. W. Bader, Chem. Rev., 1991, 91, 893–928 CrossRef CAS.
  75. S. De, A. Boda and S. M. Ali, J. Mol. Struct.: THEOCHEM, 2010, 941, 90–101 CrossRef CAS PubMed.
  76. S. Kubik and R. Goddard, Chem. Commun., 2000, 633–634 RSC.
  77. G. Praveena and P. Kolandaivel, J. Mol. Struct.: THEOCHEM, 2009, 900, 96–102 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra08302d

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.