Abstract
The quantum-chemical calculations based on density functional theory (DFT) have been performed on the diphenyltin(IV) derivative of glycyl-phenylalanine (H2L) at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory without any symmetry constraint. The harmonic vibrational frequencies were computed at the same level of theory to find the true potential energy surface minima. The various geometrical and thermochemical parameters for the studied complex are obtained in the gas phase. The atomic charges at all the atoms were calculated using the Mulliken population analysis, the Hirshfeld population analysis, and the natural population analysis. The charge distribution within the studied complex is explained on the basis of molecular electrostatic potential maps, frontier molecular orbital analysis, and conceptual DFT-based reactivity (global and local) descriptors, using the finite difference approximation method. The nature of O-Sn, N-Sn, N→Sn, and C-Sn bonds is discussed in terms of the conceptual DFT-based reactivity descriptors. The structural analysis of the studied complex has been conducted in terms of the selected bond lengths and bond angles. The structural and the atomic charge analyses suggest a distorted trigonal bipyramidal arrangement consisting of negatively charged centers around the positively charged central Sn atom.
Introduction
In recent decades, the chemistry of organotin(IV) compounds has drawn considerable research interest because of their structural diversity and wide range of synthetic and biological applications (Nath, 2008). These compounds, which are characterized by the presence of tetravalent Sn centers and at least one covalent Sn-C bond, often exhibit previously tetravalent Sn atoms expanding their coordination number and becoming hypervalent on inter- and/or intramolecular interactions with electron donor atoms. The appearance of such additional interactions results from the large size of Sn atom, the availability of low-lying empty 5d atomic orbitals, and the pronounced electron-acceptor ability of the Sn atoms (Pellerito and Nagy, 2002). However, apart from such unique structural features, organotin(IV) compounds hold significance owing to their possible use as potential biologically active nonplatinum chemotherapeutic metallopharmaceuticals having antitumor activity (Alama et al., 2009; Arjmand et al., 2014; Carraher and Roner, 2014). The speciation of organotin(IV) compounds in the biological systems has revealed that their antiproliferative activity depends on the availability of coordination positions at Sn center, number of Sn-C bonds, electronic and geometrical properties, mode of coordinating ligands, occurrence of relatively stable ligand-Sn bonds (e.g. Sn-N and Sn-S), and their slow hydrolytic decomposition (Pellerito and Nagy, 2002). As a result, in the last decades, several organotin(IV) derivatives of dipeptides have been modeled for metal-protein interactions and have also been shown to exhibit a wide range of biological activities (Katsoulakou et al., 2008; Nath et al., 2009; Girasolo et al., 2010). These studies have provided an impetus to understand the electronic properties of such derivatives so as to formulate a theoretical basis for the experimental observations.
In the contemporary research, the quantum-chemical methods based on density functional theory (DFT) have been used for understanding the electronic structure of molecules to correlate the calculated structural features with the experimental observations. The DFT-based methods have reasonably accounted for the electron correlation for organotin(IV) derivatives with heterodonor ligands; hence, several studies have successfully calculated the geometric structures for such systems by quantum-chemical calculations performed within a domain of DFT (Girichev et al., 2012; Latrous et al., 2013; Thomas et al., 2013; Matczak, 2015). Further, the study of conceptual DFT-based descriptors is indispensable to comprehend the electronic structure of an organotin(IV) complex, which has far-reaching consequences on its structure and reactivity. Moreover, with an aim to design new molecular entities possessing unique structural features and broad range of synthetic and biological applications, we have systematically initiated efforts on the theoretical investigation of the organotin(IV)-peptide system. Such studies hold significance as they will provide insight into the nature of interaction of peptide molecule with organotin(IV) moiety, and the reactivity behavior of the organotin(IV)-peptide complex will thus be formed. As a part of our systematic study on the structure and reactivity of organotin(IV)-peptide system in light of conceptual DFT-based quantum-chemical calculations (Pokharia, 2015), the present work highlights the theoretical investigation on previously synthesized diphenyltin(IV) derivative of glycylphenylalanine (Ph2SnL, where L is the dianion of glycylphenylalanine) (Figure 1A; Nath et al., 2004), so as to obtain a theoretical description for the nature of interaction of heterodonor atoms (ONN) in the dipeptide molecule with diphenyltin(IV) moiety within the complex and also its reactivity behavior. The structure and the nature of interaction of the studied complex are investigated in terms of the calculated atomic charges and frontier molecular orbital (FMO) analysis, and its reactivity is interpreted in light of conceptual DFT-based global and local reactivity descriptors.

Molecular geometry of diphenyltin(IV) derivative of glycylphenylalanine.
(A) Structure (along with the atom number) of the diphenyltin(IV) derivative of glycylphenylalanine (Ph2SnL). (B) Ground-state optimized geometry (in gas phase) of Ph2SnL calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory.
Results and discussion
Geometry optimization and structural analysis
The ground-state optimized geometry in the gas phase of Ph2SnL calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) computational level of theory is presented in Figure 1B (cf. Figure 1A for atom number notation). The frequency analysis indicates that the optimized geometric configuration is the local minima on the potential energy surface (PES) of Ph2SnL. The selected bond lengths and bond angles in Ph2SnL are presented in Table S1 (Supplementary Material).
As presented in Figure 1B, the calculated geometric environment around the central Sn atom in Ph2SnL is considerably distorted with terminal carboxylate oxygen O17 (Sn-O17=2.019 Å) and terminal amino nitrogen N1 (Sn-N1=2.430 Å) in the axial positions, and deprotonated peptide nitrogen N9 (Sn-N9=2.058 Å) and two phenyl carbons C30 and C41 (Sn-C30=2.126 Å and Sn-C41=2.124 Å) in the equatorial plane. The calculated equatorial angle C30-Sn-C41 in Ph2SnL is 118.6°, which is close to that in experimentally reported penta-coordinated diorganotin(IV) dipeptides [e.g. Ph2Sn(Gly-Gly), 117.5° (Huber et al., 1977); n-Bu2Sn(Gly-Val), 125.3° (Mundus-Glowacki et al., 1992)] but smaller to that experimentally reported in n-Bu2Sn(Trp-Gly), 151.6° (Nath et al., 2009), and Me2Sn(Tyr-Phe)·MeOH, 136.4° (Nath et al., 2008). The distortion in the molecule is evident from the calculated axial angle N1-Sn-O17 of 153.4°, which is close to the experimentally reported values for n-Bu2Sn(Trp-Gly), 147.9° (Nath et al., 2009), and Me2Sn(Tyr-Phe)·MeOH, 149.6° (Nath et al., 2008). Further, various bond angles involving the central Sn atom suggest a distorted trigonal bipyramidal arrangement. Also, there seems to be a change in bond length of various bonds involving N1, N9, and O17 in Ph2SnL relative to H2L, indicating the electron redistribution around the central Sn atom through these coordinating atoms. The calculated distorted trigonal bipyramidal arrangement is consistent with previously reported geometry of diorganotin(IV) derivatives of dipeptides (Huber et al., 1977; Mundus-Glowacki et al., 1992; Katsoulakou et al., 2008; Nath et al., 2008, 2009; Girasolo et al., 2010).
The various energetic and thermochemical parameters for Ph2SnL in the gas phase have been calculated at 1 atm and 298.15 K at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory, and the results are presented in Table S2. The total energy of Ph2SnL calculated after zero-point correction and the thermal correction to energy, enthalpy, and Gibbs free energy have been estimated; for instance, its calculated total energy after thermal correction to Gibbs free energy, which is the sum of electronic and thermal free energies (G=H-TS) (Foresman and Frisch, 1996), is -1227.9881 a.u. The contribution to internal thermal energy (Etot), entropy (Stot), and constant volume molar heat capacity (CVtot) has the value for Ph2SnL as 271.22 kcal/mol, 189.48 cal/mol-K, and 101.48 cal/mol-K, respectively.
Atomic charges and molecular electrostatic potential (MEP) map
The MEP is created in the space around a molecule by its nuclei and electrons and is related to the electron density. It holds significance because it provides substantial insight into the sites within the molecule where the electron distribution effect is dominant, for example, for the intermolecular association and molecular properties of small molecules, predicting molecular reactive behavior, and analyzing processes based on the ‘recognition’ of one molecule by another, such as between a drug and its cellular receptor, because it is through their potentials that the two species first ‘see’ each other (Politzer et al., 1985). The MEP surface of Ph2SnL calculated on the ground-state optimized geometry in gas phase at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory is presented in Figure 2A. The different values of the electrostatic potential at the surface are represented by different colors and potential increases in the order red<orange<yellow<green<blue. As evident from the MEP map (Figure 2A), the greater regions of the complex are of intermediary potential (green coded regions), but because the green color is toward the increasing range of the potential (toward the blue color), the complex possesses smaller electronegativity with electron-deficient regions having preference for an approach by nucleophiles.

Molecular electrostatic potential (MEP) map and atomic charges for diphenyltin(IV) derivative of glycylphenylalanine.
(A) Molecular electrostatic potential (MEP) map. (B) Plot of atomic charges at the selected atoms based on MPA, HPA, and NPA for the ground-state optimized geometry (in the gas phase) for Ph2SnL calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory.
The calculation of effective atomic charges is significant whenever quantum-chemical calculations are applied to molecular systems because they affect dipole moment, molecular polarizability, and electronic structure. In such systems, the atomic charges are attributed through population analysis. To study electrostatic arguments that would explain the probable structure and reactivity of Ph2SnL, an electron density distribution analysis has been performed on the basis of atomic charges determined by the Mulliken population analysis (MPA), the Hirshfeld population analysis (HPA), and the natural population analysis (NPA) charge schemes in the gas phase within Ph2SnL, at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory, and the results based on these schemes for the selected atoms in Ph2SnL are presented in Figure 2B and Table S3.
The population of the atomic orbitals suggest that in Ph2SnL, the natural electron configuration of the central Sn atom is [core]5s0.785p0.926p0.01, which differs significantly from Ph2Sn(IV)2+ cation configuration [core]5s2. The absolute value of the natural charge of the central Sn atom in Ph2SnL on the basis of NPA is 2.299. Further, the natural electron configuration of coordinating oxygen atom (O17) is [core]2s1.702p5.163d0.01, amino nitrogen (N1) is [core]2s1.432p4.523p0.01, and deprotonated peptide nitrogen (N9) is [core]2s1.352p4.523p0.01. The absolute values of the natural charge of N1, N9, and O17 on the basis of NPA are approximately -0.960, -0.884, and -0.872, respectively. Moreover, the natural electron configuration of two carbon atoms (covalently bonded to the central Sn atom), viz., C30 and C41, is [core]2s1.052p3.503p0.02. The absolute values of the natural charge of C30 and C41 on the basis of NPA are approximately -0.576 and -0.578, respectively. The existence of the high positive value of charge at the central Sn atom and such oppositely charged centers around it further confirms the ionic interaction in the Sn-O and Sn-N bonds, resulting in the formation of coordinate/dative bonds between the ONN system of deprotonated H2L and Ph2Sn(IV) moiety. The results, as presented in Figure 2B, indicate that the most negative atomic charges are attributed to oxygen, nitrogen, and organotin(IV) carbon atoms in the Ph2SnL derivative. Similarly, results are reported for tin systems with ligands having heterodonor atoms (Girichev et al., 2012; Latrous et al., 2013; Matczak, 2015). The population analysis based on MPA, HPA, and NPA also indicates an ionic interaction in these bonds (Table S3). The presence of such an ionic interaction in Ph2SnL is significant as it may lead to a slow hydrolytic decomposition of Sn-O/N bonds, suggesting that the complex can exhibit potential activity, such as antiproliferative activity owing to the existence of R2Sn(IV)2+ (R=Ph) moiety in the biological medium.
Conceptual DFT-based global reactivity descriptors
The molecular properties and the conceptual DFT-based global reactivity descriptors for the ground-state optimized geometries in the gas phase of H2L and Ph2SnL, calculated using finite difference approximation, are presented in Table 1. The results indicate that the dipole moment that accounts for the existence of charged separated regions within the system is greater for Ph2SnL (10.35 Debye) in comparison with H2L (3.31 Debye). The ionization potential (IP) and the electron affinity (EA) for Ph2SnL are lower than that for H2L. As a result, the band gap (ΔE) for Ph2SnL (7.89 eV) is lower than H2L (9.50 eV).
Calculated molecular properties and conceptual DFT-based global reactivity descriptors for the ground-state optimized geometries (in gas phase) of Gly-Phe (H2L) at B3LYP/6-31G(d,p) and Ph2SnL at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory.
Parameter/property | System | |
---|---|---|
Gly-Phe (H2L) | Ph2SnL | |
EN (a.u.)a | -762.824913 | -1228.331213 |
EN+1 (a.u.) | -762.773851 | -1228.313607 |
EN-1 (a.u.) | -762.526946 | -1228.059070 |
Dipole moment (Debye)b | 3.31 | 10.35 |
IP (eV)c | 8.11 | 7.41 |
EA (eV)d | -1.39 | -0.48 |
ΔE (eV)e | 9.50 | 7.89 |
EHOMO (eV)f | -8.11 | -7.41 |
ELUMO (eV)g | 1.39 | 0.48 |
Electronic chemical potential (μ) (eV)h | -3.36 | -3.46 |
Electronegativity (χ) (eV)i | 3.36 | 3.46 |
Global hardness (η) (eV)j | 9.50 | 7.89 |
Global softness (S) (/eV)k | 0.11 | 0.13 |
Electrophilicity index (ω) (eV)l | 0.59 | 0.76 |
aEN, EN+1, and EN-1 are the total energies of the system containing, respectively, N, N+1, and N-1 electrons.
bDipole moment for the system containing N number of electrons.
cIP is the ionization potential given by EN-1-EN.
dEA is the electron affinity given by EN-EN+1.
eΔE is the band gap given by IP-EA.
fEnergy of the highest occupied molecular orbital as -EHOMO=IP.
gEnergy of the lowest unoccupied molecular orbital as -ELUMO=EA.
hElectronic chemical potential of the system given by
iElectronegativity of the system given by -μ.
jGlobal hardness of the system given by ELUMO-EHOMO (Parr and Pearson, 1983; Geerlings et al., 2003).
kGlobal softness of the system given by 1/η (Yang and Parr, 1985; Geerlings et al., 2003).
lElectrophilicity index of the system given by μ2/2η (Parr et al., 1999; Geerlings et al., 2003; Chattaraj et al., 2006).
The FMO analysis is significant from the perspective of understanding the distribution of charge density within a system as it helps in estimating the energies and the type of frontier molecular orbitals because these orbitals are the sites of exchange of charge density, which leads to the interaction between molecules. The FMO analysis for Ph2SnL has been performed through the Koopman’s approximation within the molecular orbital theory (Geerlings et al., 2003), and the EHOMO and the ELUMO energies are presented in Table 1. The EHOMO and the ELUMO plots along with the band gap (ΔE) in the gas phase, as presented in Figure 3, indicate that, in Ph2SnL, the highest occupied molecular orbital (HOMO) is concentrated over the dipeptide, whereas the lowest unoccupied molecular orbital (LUMO) is concentrated around the central Sn atom and organic residues in organotin(IV) moiety, indicative of the interaction of H2L with Ph2Sn(IV) moiety through deprotonated oxygen (O17) and peptide nitrogen (N9) atoms, as reported previously for other diorganotin(IV)-dipeptide systems (Huber et al., 1977; Mundus-Glowacki et al., 1992; Nath et al., 2008, 2009; Girasolo et al., 2010). Further, the FMO analysis suggests that upon interaction with an electron donor/acceptor, in Ph2SnL, the region around the central Sn atom will behave as an acceptor of charge density, whereas the dipeptide region (L2-) will behave as a donor of charge density. Moreover, the studied complex interacts with macromolecular receptors through the central Sn atom upon the slow hydrolysis (owing to ionic interaction) of Sn-O/N bonds.

Highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) plots along with band gap (ΔE) in the gas phase for Ph2SnL calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory.
The global reactivity descriptors calculated based on FMO analysis are tabulated in Table 1. The electronic chemical potential (μ), which measures the sensitiveness of the system’s energy to a change in the number of electrons at fixed external potential
Conceptual DFT-based local reactivity descriptors
The conceptual DFT-based local reactivity descriptors, such as Fukui function index [f±(k)] (Parr and Yang, 1984; Geerlings et al., 2003; Roy and Saha, 2010), local softness [s±(k)] (Geerlings et al., 2003; Roy and Saha, 2010), local electrophilicity [ω±(k)] (Geerlings et al., 2003; Chattaraj et al., 2006; Roy and Saha, 2010), and hardness potential [Δ±h(k)] (Geerlings et al., 2003; Saha et al., 2013), considered to predict the regioselectivity of the atoms in Ph2SnL, have been calculated at all the atoms with three different population analysis schemes, viz., MPA, HPA, and NPA, and the results based on the NPA, HPA, and MPA schemes at the selected atoms of Ph2SnL are presented in Tables S4, S5, and S6, respectively. The variation of f+(k) and s+(k), f-(k) and s-(k), ω+(k) and ω-(k), and Δ+h(k) and Δ-h(k) at the selected atoms of Ph2SnL calculated on the basis of MPA, HPA, and NPA charge schemes is presented in Figure 4A–D, respectively.

Variation of (A) f+(k) and s+(k), (B) f-(k) and s-(k), (C) ω+(k) and ω-(k), and (D) Δ+h(k) and Δ-h(k) at the selected atoms of Ph2SnL based on MPA, HPA, and NPA charges, calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory (all values are given in eV).
Because there are some differences between the atomic charges obtained from MPA and NPA schemes (Reed et al., 1985), Fukui indices will be highly dependent on the population analysis method, and the results should therefore be analyzed with greater restraint. The Fukui function
As evident from the results (cf. Figure 4), in Ph2SnL, the trends based on the values of f+(k), s+(k), and ω+(k) with all the population scheme suggest that the central Sn atom is an electrophilic site and is more reactive toward a nucleophile and that such an attack increases the atomic population at this center. Similarly, the trends based on the values of f-(k), s-(k), and ω-(k) with all the population schemes suggest that O18 is the most reactive site toward an electrophilic attack. The results obtained are in accordance with the FMO analysis, which indicates that in Ph2SnL, the HOMO is concentrated over the dipeptide unit and the LUMO is concentrated around the central Sn atom encompassing the phenyl residues covalently bonded to the central Sn atom. These results outline the general behavior of the central Sn atom in the organotin(IV) complexes in accordance with the previously reported experimental observations, that R2Sn(IV)2+ is an active species in the biological medium (Pellerito and Nagy, 2002). The two variants of the hardness potential, viz., the electrophilic hardness potential [Δ+h(k)] and the nucleophilic hardness potential [Δ-h(k)], which measures the reactivity toward an approaching nucleophilic and electrophilic reagent, respectively, have also been calculated at all the atoms of Ph2SnL. The higher the value of Δ+h(k) at an atom, the higher the reactivity (i.e. electrophilicity) of that atom toward an approaching nucleophile (Saha et al., 2013). Likewise, the higher the positive value of Δ-h(k) for an atom, the higher the reactivity (i.e. nucleophilicity) of that atom toward an approaching electrophile. As evident from the results (cf. Table S4 and Figure 4D), in Ph2SnL, the trends based on the values of Δ+h(k) suggest that C30/C41 (carbon bonded covalently to the Sn atom) is the most reactive site toward an approaching nucleophilic reagent, and the trends based on the values of Δ-h(k) suggest that O18 is among the most reactive site toward an approaching electrophilic reagent. Further, in the coordination sphere around the central Sn atom, the higher value of Δ+h(k) at the central Sn atom, N1(amino), and C30/C41 indicates that these sites are more reactive toward a nucleophilic reagent. However, the higher value of Δ-h(k) at N9(peptidic) and O17/O18(carboxylate) indicates that these sites are more reactive toward an electrophilic reagent.
The variation of relative electrophilicity

Variation of (A) relative electrophilicity, (B) relative nucleophilicity, and (C) dual reactivity descriptor at the selected atoms of Ph2SnL based on MPA, HPA, and NPA charges, calculated at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory (all values in panel C are in eV).
Conclusion
The present study significantly demonstrates the efficacy of quantum-chemical methods in studying the electronic structure of diphenyltin(IV) derivative of glycylphenylalanine at the B3LYP/6-31G(d,p)/LANL2DZ(Sn) level of theory, which will have greater ramifications in getting the vital insight into the structural features of the complex as also to gain insight into the plausible mechanism for their reactivity, specifically upon interaction with macromolecular receptors. The calculated values for the coordinating bond length and various other geometrical parameters in Ph2SnL are closer to those reported for other organotin(IV) complexes with heterodonor atoms.
The conceptual DFT-based global and local reactivity descriptors have been effectively calculated on the basis of MPA, HPA, and NPA schemes. The calculated global reactivity descriptors indicate that complexation leads to softness in Ph2SnL relative to H2L. There seem to be some discrepancies in the calculation of conceptual DFT-based local reactivity descriptors on the basis of MPA and NPA schemes. However, irrespective of the population scheme, these descriptors can convincingly explain the charge distribution at all the atoms of Ph2SnL and noticeably identify various nucleophilic and electrophilic sites within the complex, which will further provide some mechanistic insight into the structure and reactivity of organotin(IV) complexes with heterodonor atoms, apart from its biological activity.
The theoretical calculations performed have identified the FMOs for Ph2SnL, wherein the HOMO is concentrated over the dipeptide unit and the LUMO is concentrated around the central Sn atom encompassing phenyl moiety bonded to the central Sn atom.
Experimental
Computational details
All the quantum-chemical calculations have been performed using the Gaussian 09 program package (Frisch et al., 2010). The molecular geometries of glycylphenylalanine (H2L) and Ph2SnL were fully optimized in the gas phase at the DFT level using the B3LYP functional, which is a combination of Becke’s three parameter (B3) gradient corrected hybrid exchange functional (Becke, 1993), with the dynamical correlation functional of Lee et al. (1988) (LYP). All the atoms except Sn were described by the 6-31G(d,p) basis set, which contains a reasonable number of basis set functions that are able to reproduce the experimental observations. The Sn atom in Ph2SnL was described by the LANL2DZ basis set in which Sn inner shells are described by effective core potential ECP46MWB (1s22s22p63s23p63d104s24p64d10) along with the basis set (3s3p)/[2s2p] (Hay and Wadt, 1985). The geometry optimization was conducted through an algorithm of iterative steps to locate true global minima on the PES with default parameters for convergence is met. The absence of an imaginary frequency in a harmonic frequency calculation conducted at the same level of theory indicates that the calculated geometry is a true global minimum on the PES. The atomic charges at all the atoms in the studied systems were calculated using MPA, HPA, and NPA at the same level of theory. The energies of FMOs and the conceptual DFT-based global and local reactivity descriptors based on MPA, HPA, and NPA have been calculated for the studied systems using finite difference approximation (Geerlings et al., 2003). The MEP maps and the visualization of all results have been performed using Gauss View 5.0 (Dennington et al., 2009).
Acknowledgments
The authors are thankful to Banaras Hindu University, Varanasi, for providing basic infrastructural and computational facilities.
References
Alama, A.; Tasso, B.; Novelli, F.; Sparatore, F. Organometallic compounds in oncology: implications of novel organotins as antitumor agents. Drug Discov. Today.2009, 14, 500–508.10.1016/j.drudis.2009.02.002Search in Google Scholar PubMed
Arjmand, F.; Parveen, S.; Tabassum, S.; Pettinari, C. Organo-tin antitumour compounds: their present status in drug development and future perspectives. Inorg. Chim. Acta. 2014, 423, 26–37.10.1016/j.ica.2014.07.066Search in Google Scholar
Becke, A. D. Density functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652.10.1063/1.464913Search in Google Scholar
Carraher, C. E.; Roner, M. R. Organotin polymers as anticancer and antiviral agents. J. Organomet. Chem.2014, 751, 67–82.10.1016/j.jorganchem.2013.05.033Search in Google Scholar
Chattaraj, P. K.; Sarkar, U.; Roy, D. R. Electrophilicity index. Chem. Rev.2006, 106, 2065–2091.10.1021/cr040109fSearch in Google Scholar PubMed
Dennington II, R. D.; Keith, T. A.; Millam, J. M. Gauss View 5.0. Gaussian, Inc.: Wallingford CT, 2009.Search in Google Scholar
Foresman, J. B.; Frisch, E. Exploring Chemistry with Electronic Structure Methods. Gaussian, Inc.: Pittsburgh, USA, 1996.Search in Google Scholar
Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian09, Revision B.01. Gaussian, Inc.: Wallingford, CT, 2010.Search in Google Scholar
Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual density functional theory. Chem. Rev.2003, 103, 1793–1873.10.1201/9781420065442.ch27Search in Google Scholar
Girasolo, M. A.; Rubino, S.; Portanova, P.; Calvaruso, G.; Ruisi, G.; Stocco, G. New organotin(IV) complexes with l-arginine, Nα-t-Boc-l-arginine and l-alanyl-l-arginine: synthesis, structural investigations and cytotoxic activity. J. Organomet. Chem.2010, 695, 609–618.10.1016/j.jorganchem.2009.11.002Search in Google Scholar
Girichev, G. V.; Giricheva, N. I.; Koifman, O. I.; Minenkov, Y. V.; Pogonin, A. E.; Semeikin, A. S.; Shlykov, S. A. Molecular structure and bonding in octamethyl-porphyrin tin(II), SnN4C28H28. Dalton Trans. 2012, 41, 7550–7558.10.1039/c2dt12499hSearch in Google Scholar PubMed
Hay, P. J.; Wadt, W. R. Ab initio effective core potentials for molecular calculations-potentials for K to Au including the outermost core orbitals. J. Chem. Phys.1985, 82, 299–310.10.1063/1.448975Search in Google Scholar
Huber, F.; Haupt, H. J.; Preut, H.; Barbieri, R.; LoGuidice, M. T. Preparation, crystal and molecular structure of diphenyltin-glycylglycinate (C6H5)2SnC4H6N2O3. Z. Anorg. Allg. Chem.1977, 432, 51–57.10.1002/zaac.19774320106Search in Google Scholar
Katsoulakou, E.; Tiliakos, M.; Papaefstathiou, G.; Terzis, A.; Raptopoulou, C.; Geromichalos, G.; Papazisis, K.; Papi, R.; Pantazaki, A.; Kyriakidis, D.; et al. Diorganotin(IV) complexes of dipeptides containing the α-aminoisobutyryl residue (Aib): preparation, structural characterization, antibacterial and antiproliferative activities of [(n-Bu)2Sn(H-1L)] (LH = H-Aib-L-Leu-OH, H-Aib-L-Ala-OH). J. Inorg. Biochem.2008, 102, 1397–1405.10.1016/j.jinorgbio.2008.01.001Search in Google Scholar
Latrous, L.; Tortajada, J.; Haldys, V.; Léon, E.; Correia, C.; Salpin, J.-Y. Gas-phase interactions of organotin compounds with glycine. J. Mass. Spectrom. 2013, 48, 795–806.10.1002/jms.3223Search in Google Scholar
Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B.1988, 37, 785–789.10.1103/PhysRevB.37.785Search in Google Scholar
Matczak, P. Theoretical investigation of the N→Sn coordination in (Me3SnCN)2. Struct. Chem.2015, 26, 301–318.10.1007/s11224-014-0485-4Search in Google Scholar
Morell, C.; Grand, A.; Toro-Labbé, A. New dual descriptor for chemical reactivity. J. Phys. Chem. A2005, 109, 205–212.10.1021/jp046577aSearch in Google Scholar
Mundus-Glowacki, B.; Huber, F.; Preut, H.; Ruisi, G.; Barbieri, R. Synthesis and spectroscopic characterization of dimethyl-, di-n-butyl, di-t-butyl- and diphenyl-tin(IV) derivatives of dipeptides: crystal and molecular structure of di-n-butyltin(IV)glycylvalinate. Appl. Organometal. Chem.1992, 6, 83–94.10.1002/aoc.590060111Search in Google Scholar
Nath, M. Toxicity and the cardiovascular activity of organotin compounds: a review. Appl. Organometal. Chem.2008, 22, 598–612.10.1002/aoc.1436Search in Google Scholar
Nath, M.; Pokharia, S.; Eng, G.; Song, X.; Kumar, A. Diorganotin(IV) derivatives of dipeptides containing at least one essential amino acid residue: synthesis, characteristic spectral data, cardiovascular, and anti-inflammatory activities. Synth. React. Inorg. Met.-Org. Chem.2004, 34, 1689–1708.10.1081/SIM-200030161Search in Google Scholar
Nath, M.; Singh, H.; Eng, G.; Song, X. New di- and triorganotin(IV) derivatives of tyrosinylphenylalanine as models for metal-protein interactions: synthesis and structural characterization. Crystal structure of Me2Sn(Tyr-Phe).MeOH. J. Organomet. Chem. 2008, 693, 2541–2550.10.1016/j.jorganchem.2008.04.032Search in Google Scholar
Nath, M.; Singh, H.; Kumar, P.; Kumar, A.; Song, X.; Eng, G. Organotin(IV) tryptophanylglycinates: potential nonsteroidal anti-inflammatory agents; crystal structure of dibutyltin(IV) tryptophanylglycinate. Appl. Organometal. Chem.2009, 23, 347–358.10.1002/aoc.1514Search in Google Scholar
Parr, R. G.; Pearson, R. G. Absolute hardness: companion parameter to absolute electronegativity. J. Am. Chem. Soc.1983, 105, 7512–7516.10.1021/ja00364a005Search in Google Scholar
Parr, R. G.; Yang, W. Density functional approach to the frontier-electron theory of chemical reactivity. J. Am. Chem. Soc. 1984, 106, 4049–4050.10.1021/ja00326a036Search in Google Scholar
Parr, R. G.; Szentpály, L. V.; Liu, S. Electrophilicity index. J. Am. Chem. Soc.1999, 121, 1922–1924.10.1021/ja983494xSearch in Google Scholar
Pellerito, L.; Nagy, L. Organotin(IV)n+ complexes formed with biologically active ligands: equilibrium and structural studies, and some biological aspects. Coord. Chem. Rev.2002, 224, 111–150.10.1016/S0010-8545(01)00399-XSearch in Google Scholar
Pokharia, S. Theoretical insights on organotin(IV)-protein interaction: density functional theory (DFT) studies on di-n-butyltin(IV) derivative of glycylvaline. Asian J. Res. Chem.2015, 8, 7–12.10.5958/0974-4150.2015.00002.4Search in Google Scholar
Politzer, P.; Laurence, P. R.; Jayasuriya, K. Molecular electrostatic potentials: an effective tool for the elucidation of biochemical phenomena. Env. Health. Perspec. 1985, 61, 191–202.10.1289/ehp.8561191Search in Google Scholar PubMed PubMed Central
Reed, A. E.; Wienstock, R. B.; Weinhold, F. Natural population analysis. J. Chem. Phys.1985, 83, 735–746.10.1063/1.449486Search in Google Scholar
Roy, R. K.; Krishnamurti, S.; Geerlings, P.; Pal, S. Local softness and hardness based reactivity descriptors for predicting intra- and intermolecular reactivity sequences: carbonyl compounds. J. Phys. Chem. A1998, 102, 3746–3755.10.1021/jp973450vSearch in Google Scholar
Roy, R. K.; Saha, S. Studies of regioselectivity of large molecular systems using DFT based reactivity descriptors. Annu. Rep. Prog. Chem., Sect. C2010, 106, 118–162.10.1039/b811052mSearch in Google Scholar
Saha, S.; Bhattacharjee, R.; Roy, R. K. Hardness potential derivatives and their relation to Fukui indices. J. Comput. Chem.2013, 34, 662–672.10.1002/jcc.23177Search in Google Scholar PubMed
Thomas, R.; Nelson, J. P.; Pardasani, R. T.; Pardasani, P.; Mukherjee, T. Novel tin complexes containing an oximato ligand: synthesis, characterization, and computational investigation. Helv. Chim. Acta2013, 96, 1740–1749.10.1002/hlca.201200259Search in Google Scholar
Yang, W.; Parr, R. G. Hardness, softness, and the fukui function in the electronic theory of metals and catalysis. Proc. Natl. Acad. Sci. U.S.A.1985, 82, 6723–6726.10.1073/pnas.82.20.6723Search in Google Scholar PubMed PubMed Central
Supplemental Material:
The online version of this article (DOI: 10.1515/mgmc-2016-0009) offers supplementary material, available to authorized users.
©2016 Walter de Gruyter GmbH, Berlin/Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Frontmatter
- Research Articles
- Synthesis, spectroscopic, and theoretical studies of tin(II) complexes with biologically active Schiff bases derived from amino acids
- A density functional theory insight into the structure and reactivity of diphenyltin(IV) derivative of glycylphenylalanine
- Synthesis, crystal structure, and antibacterial activity of tricyclohexyltin salicylates
- Fungal strain Aspergillus flavus F3 as a potential candidate for the removal of lead (II) and chromium (VI) from contaminated soil
- Two SrII coordination compounds based on tetrazole-carboxylate ligands
- Short Communications
- Crystal structure of the triphenyltin(IV) chloride dimethyl N-cyanodithioiminocarbonate adduct
- Triorganotin carboxylates – synthesis and crystal structure of 2-methyl-1H-imidazol-3-ium catena-O,O′-oxalatotriphenylstannate
- Organotin(IV) scorpionates – X-ray structure and crystal packing of TpSn(Cl)2(n-Bu) [Tp=hydrotris(pyrazol-1-yl)borate]
Articles in the same Issue
- Frontmatter
- Research Articles
- Synthesis, spectroscopic, and theoretical studies of tin(II) complexes with biologically active Schiff bases derived from amino acids
- A density functional theory insight into the structure and reactivity of diphenyltin(IV) derivative of glycylphenylalanine
- Synthesis, crystal structure, and antibacterial activity of tricyclohexyltin salicylates
- Fungal strain Aspergillus flavus F3 as a potential candidate for the removal of lead (II) and chromium (VI) from contaminated soil
- Two SrII coordination compounds based on tetrazole-carboxylate ligands
- Short Communications
- Crystal structure of the triphenyltin(IV) chloride dimethyl N-cyanodithioiminocarbonate adduct
- Triorganotin carboxylates – synthesis and crystal structure of 2-methyl-1H-imidazol-3-ium catena-O,O′-oxalatotriphenylstannate
- Organotin(IV) scorpionates – X-ray structure and crystal packing of TpSn(Cl)2(n-Bu) [Tp=hydrotris(pyrazol-1-yl)borate]