Open access

Using Molecular Modelling to Study Interactions Between Molecules with Biological Activity

Written By

María J. R. Yunta

Submitted: 03 October 2011 Published: 28 November 2012

DOI: 10.5772/54007

From the Edited Volume

Bioinformatics

Edited by Horacio Pérez-Sánchez

Chapter metrics overview

4,601 Chapter Downloads

View Full Metrics

1. Introduction

To better understand the basis of the activity of any molecule with biological activity, it is important to know how this molecule interacts with its site of action, more specifically its conformational properties in solution and orientation for the interaction. Molecular recognition in biological systems relies on specific attractive and/or repulsive interactions between two partner molecules. This study seeks to identify such interactions between ligands and their host molecules, typically proteins, given their three-dimensional (3D) structures. Therefore, it is important to know about interaction geometries and approximate affinity contributions of attractive interactions. At the same time, it is necessary to be aware of the fact that molecular interactions behave in a highly non-additive fashion. The same interaction may account for different amounts of free energy in different contexts and any change in molecular structure might have multiple effects, so it is only reliable to compare similar structures. In fact, the multiple interactions present in a single two-molecule complex are a compromise between attractive and repulsive interactions. On the other hand, a molecular complex is not characterised by a single structure, as can be seen in crystal structures, but by an ensemble of structures. Furthermore, changes in the degree of freedom of both partners during an interaction have a large impact on binding free energy [Bissantz et al., 2010].

The availability of high-quality molecular graphics tools in the public domain is changing the way macromolecular structure is perceived by researchers, while computer modelling has emerged as a powerful tool for experimental and theoretical investigations. Visualisation of experimental data in a 3D, atomic-scale model can not only help to explain unexpected results but often raises new questions, thereby affecting future research. Models of sufficient quality can be set in motion in molecular dynamic (MD) simulations to move beyond a static picture and provide insight into the dynamics of important biological processes.

Computational methods have become increasingly important in a number of areas such as comparative or homology modelling, functional site location, characterisation of ligand-binding sites in proteins, docking of small molecules into protein binding sites, protein-protein docking, and molecular dynamic simulations [see for example Choe & Chang, 2002]. Current results yield information that is sometimes beyond experimental possibilities and can be used to guide and improve a vast array of experiments.

To apply computational methods in drug design, it is always necessary to remember that to be effective, a designed drug must discriminate successfully between the macromolecular target and alternative structures present in the organism. The last few years have witnessed the emergence of different computational tools aimed at understanding and modelling this process at the molecular level. Although still rudimentary, these methods are shaping a coherent approach to help in the design of molecules with high affinity and specificity, both in lead discovery and in lead optimisation. Moreover, current information on the 3D structure of proteins and their functions provide a possibility to understand the relevant molecular interactions between a ligand and a target macromolecule. As a consequence, a comprehensive study of drug structure–activity relationships can help identify a 3D pharmacophore model as an aid for rational drug design, as a pharmacophore model can be defined as ‘an ensemble of steric and electronic features that is necessary to ensure the optimal supramolecular interactions with a specific biological target and to trigger (or block) its biological response’, and a pharmacophore model can be established either in a ligand-based manner, by superposing a set of active molecules and extracting common chemical features that are essential for their bioactivity, or in a structure-based manner, by probing possible interaction points between the macromolecular target and ligands.

Molecular recognition (MR) is a general term designating non-covalent interactions between two or more compounds belonging to host-guest, enzyme-inhibitor and/or drug-receptor complexes. A rigorous approach to an MR study should involve the adoption of a computational method independent from the chemical intuition of the researcher. Drug design purposes prompt another challenging feature of such an ideal computational method, the ability to make sufficiently accurate thermodynamic predictions about the recognition process.

Advertisement

2. Molecular modelling methods and their usefulness

Molecular recognition is a central phenomenon in biology, for example, with enzymes and their substrates, receptors and their signal inducing ligands, antibodies and antigens, among others. Given two molecules with 3D conformations in atomic detail, it is important to know if the molecules bind to each other and, if it is so, what does the formed complex look like (“docking”) and how strong is the binding affinity (that can be related to the “scoring”functions).

Molecules are not rigid. The motional energy at room temperature is large enough to let all atoms in a molecule move permanently. That means that the absolute positions of atoms in a molecule, and of a molecule as a whole, are by no means fixed, and that the relative location of substituents on a single bond may vary with time. Therefore, any compound containing one or several single bonds exists at every moment in many different conformers, but generally only low energy conformers are found to a large extent [Kund, 1997, as cited in Tóth et al., 2005].

The biological activity of a drug molecule is supposed to depend on one single unique conformation amongst all the low energy conformations, the search for this so-called bioactive conformation for compound sets being one of the major tasks in Medicinal Chemistry. Searching for all low energy conformations is possible with molecular modelling studies, since molecular modelling is concerned with the description of the atomic and molecular interactions that govern microscopic and macroscopic behaviours of physical systems. These molecular interactions are classified as: (a) bonded (stretching, bending and torsion), (b) non-bonded (electrostatic (including interactions with metals), van der Waals and π-stacking), and (c) derived, as they result from the previous ones (hydrogen bonds and hydrophobic effect).

Protein-ligand or, in general, molecule-molecule binding free energy differences can a priori be computed from first principles using free energy perturbation techniques and a full atomic detailed model with explicit solvent molecules using molecular dynamics simulations. However, these are computationally demanding. More affordable approaches use end-point molecular dynamic simulations and compute free energies accounting for solvent effects with continuum methods, such as MM-PBSA (molecular mechanics Poisson-Boltzman surface area) or MM-GBSA (generalized Born surface area) [Kollman et al., 2000; Wang et al., 2005]. One of the first approaches was comparative molecular field analysis (CoMFA) [Cramer et al., 1988], which enabled interpretation and understanding of enzyme active sites when the crystal structure was absent. However, this type of analysis was not possible until in vitro drug-drug interaction studies were widely used (through the 1990s).

2.1. Molecular mechanics, molecular dynamics and docking

Molecular mechanics (MM) is often the only feasible means with which to model very large and non-symmetrical chemical systems such as proteins and polymers. Molecular mechanics is a purely empirical method that neglects explicit treatment of electrons, relying instead on the laws of classical physics to predict the chemical properties of molecules. As a result, MM calculations cannot deal with problems such as bond breakage or formation, where electronic or quantum effects dominate. Furthermore, MM models are wholly system-dependent. MM energy predictions tend to be meaningless as absolute quantities, as the zero or reference value depends on the number and types of atoms and their connectivity, and so they are generally useful only for comparative studies. A force field is an empirical approximation for expressing structure-energy relationships in molecules and is usually a compromise between speed and accuracy.

Molecular mechanics have been shown to produce more realistic geometry values for the majority of organic molecules, owing to the fact that they are highly parameterised. Parameterisation of structures should be performed with care and non-“standard” molecules will need to have new parameters. This is usually done by analogy for bonded terms and assigning charges by a procedure consistent with the used force field.

There are many levels of theory in which computational models of 3D structures can be constructed. The overall aim of modelling methods is often to try to relate biological activity to structure. An important step towards this goal is to be able to compute the potential energy of the molecule as a function of the position of the constituent atoms. Once a method for evaluating the molecular potential energy is available, it is natural to search for an optimum molecular geometry by minimising the energy of the system. In a biological macromolecule, the potential energy surface is a complicated one, in which there are many local energy minima as well as a single overall energy minimum. All the energy minimisation algorithms commonly used have a marked tendency to locate only a local energy minimum that is close to the starting conformation. For a biological macromolecule, the number of conformations that have to be searched rises exponentially with the size of the molecule; hence, systematic searching is not a practical method for large molecules.

Molecular dynamics (MD) is a conformation space search procedure in which the atoms of a biological macromolecule are given an initial velocity and are then allowed to evolve in time according to the laws of Newtonian mechanics [van Gunsteren & Berendsen, 1977]. Depending on the simulated temperature of the system, the macromolecule can then overcome barriers at the potential energy surface in a way that is not possible with a minimisation procedure. One useful combination of molecular dynamics and minimisation schemes is a method known as simulated annealing [Kirkpatrick et al, 1983, Černý, 1985]. This method uses a molecular dynamics calculation in which the system temperature is raised to a high value to allow for a widespread exploration of the available conformational space. The system temperature is then gradually decreased as further dynamics are performed. Finally, a minimisation phase may be used to select a minimum energy molecular conformation.

One of the most important applications of molecular modelling techniques in structural biology is the simulation of the docking of a ligand molecule onto a receptor. These methods often search to identify the location of the ligand binding site and the geometry of the ligand in the active site, to get the correct ranking when considering a series of related ligands in terms of their affinity, or to evaluate the absolute binding free energy as accurately as possible. To select a force field and the adequate modelling methodology for a given task, it is important to appreciate the range of molecular systems to which it is applicable and the types of simulations that can be performed.

2.2. Most used existing force fields

AMBER (Assisted Model Building with Energy Refinement) developed by Kollman et al. [http://ambermd.org/] was originally parameterised specifically for proteins and nucleic acids [Weiner et al., 1984, 1986; Cornell et al., 1995], using 5 bonding and non-bonding terms along with a sophisticated electrostatic treatment and with no cross terms included. The results obtained with this method can be very good for proteins and nucleic acids, but less so for other systems, although parameters that enable the simulation of other systems have been published [for examples see Doshi & Hamelberg, 2009; Zgarbová et al., 2011].

CHARMM (Chemistry at HARward Macromolecular Mechanics) developed by Karplus et al. [http://www.charmm.org] was originally devised for proteins and nucleic acids [Brooks et al., 1983], and is now used for a range of macromolecules, molecular dynamics, solvation, crystal packing, vibrational analysis and QM/MM (quantum mechanics/molecular mechanics) studies. It uses five valence terms, one of which is electrostatic and is a basis for other force fields (e.g., MOIL [Elber et al., 1995]).

GROMOS (Groningen Molecular Simulation) developed at the University of Groningen and the ETH (Eidgenössische Technische Hochschule) of Zurich [http://www.igc.ethz. ch/GROMOS/index] is quite popular for predicting the dynamical motion of molecules and bulk liquids, also being used for modelling biomolecules. It uses five valence terms, one of which is electrostatic [van Gunsteren and Berendsen., 1977]. Its parameters are currently being updated [Horta et al., 2011].

MM1-4 (Molecular Mechanics) developed by Allinger [1976] are general purpose force fields for monofunctional organic molecules. The first version of this method was the MM1 [Allinger, 1976]. MM2 was parameterised for a lot of functional groups while MM3 [Allinger & Durkin, 2000; Allinger & Yan, 1993] is probably one of the most accurate ways of modelling hydrocarbons. MM4 is the latest version with several improvements [Allinger et al., 1996].

MMFF (Merck Molecular Force Field) developed by Halgren [1996] is also a general purpose force field mainly for organic molecules. MMFF94 [Halgren, 1996] was originally designed for molecular dynamics simulations, but has also been widely used for geometrical optimisation. It uses five valence terms, one of which is electrostatic and another is a cross term. MMFF was parameterised based on high level ab initio calculations. MMFF94 contains parameters for a wide variety of functional groups that arise in Organic and Medicinal Chemistry.

OPLS (Optimized Potential for Liquid Simulations) developed by Jorgensen at Yale [http://zarbi.chem.yale.edu] was designed for modelling bulk liquids [Jorgensen & Tirado-Rives, 1996] and has been extensively used for modelling the molecular dynamics of biomolecules. It uses five valence terms, one of which is an electrostatic term and none of them is a cross term.

TRIPOS (Sybil force field) is a commercial method designed for modelling organics and biomolecules. It is often used for CoMFA analysis and uses five valence terms, one of which is an electrostatic term.

CVFF (Consistent Valence Force Field) developed by Dauber-Osguthorpe is a method parameterised for small organic (amides and carboxylic acids, among others) crystals and gas phase structures [Dauber-Osguthorpe et al., 2004]. It handles peptides, proteins and a wide range of organic systems. It was primarily intended for studies of structures and binding energies, although it predicts vibrational frequencies and conformational energies reasonably well.

2.3. Popular docking programs

One of the most important and useful areas of application of molecular modelling is the approach of docking a protein onto a second molecule, typically a small ligand. This is of interest because it models the possible interactions between the protein and the ligand in the formation of a biologically important protein-ligand complex. To perform a computational docking, experimental or model 3D structures of both the protein and ligand molecules are required together with the charge distribution for each molecule.

There are several software programs that are available for carrying out docking calculations, only some of them will be considered here. The DOCK program suite [Kuntz, 1992] is one of the best known. First of all, a set of overlapping spheres are used in the program to construct a negative image of a specified site on the protein or another macromolecule, and the negative image is then matched against structures of potential ligands. Matches can be scored in this program by the quality of the geometric fit, as well as by the molecular mechanics interaction energy [Meng et al., 1992] and can lead to protein-binding ligands that have micromolecular levels of binding affinity [Kuntz et al., 1994]. It has also been used for modelling protein-protein docking [Shoichet & Kuntz, 1996].

The program GRID [Goodford, 1985] identifies likely protein binding sites for ligands [Reynolds et al., 1989; Cruciani & Goodford, 1994] using a 3D grid around the protein.

The program AutoDock developed by Morris et al.; [http://www.scripps.edu/pub/olson-web/doc/autodock/] uses a grid-based scheme for energies of individual atoms, allowing a quick computation of the interaction energy of the protein-ligand complex as the interaction between the ligand and the grid.

GLIDE software [Friesner et al., 2004, 2006; Halgren et al., 2004] also uses a grid-based scheme to represent the shape and properties of the receptor and then uses a systematic search algorithm to produce a set of initial conformations, using a OPLS-AA force field for ligand minimisation in the field of the receptor.

SURFLEX [Jain, 2003, 2007] is a fully automatic flexible molecular docking algorithm that presents results evaluated for reliability and accuracy in comparison with crystallographic experimental results on 81 protein/ligand pairs of substantial structural diversity.

In a recent study, comparison of seven popular docking programs [Plewczynski et al., 2011] clearly showed that the ligand binding conformation could be identified in most cases by using the existing software. Yet, there is still the lack of universal scoring function for all types of molecules and protein families. One can always hope that incremental improvements in current techniques will gradually lead to major advances in this field.

Advertisement

3. The solvent and how to model it

Solvation plays an important role in ligand-protein association and has a strong impact on comparisons of binding energies for dissimilar molecules. The binding affinity of a ligand for a receptor (ΔGbind) depends on the interaction free energy of the two molecules relative to their free energy in solution:

ΔGbind = ΔGinteract – ΔGsolv,L - ΔGsolv,R

where ΔGinteract is the interaction free energy of the complex, ΔGsolv,L is the free energy of desolvating the ligand, and ΔGsolv,R is the free energy of occluding the receptor site from the solvent. Various methods have been proposed to evaluate or estimate these terms. The problem is difficult because the energy of each component on the right hand side of Equation 1 is large while the difference between them is small.

An accurate way to calculate relative binding energies is with free-energy perturbation techniques, although they are usually restricted to calculating the differential binding of similar compounds and require extensive computation, making it impractical as an initial screen, but quite useful sometimes [Buch et al., 2011; Reddy & Erion, 2007]. Several authors have described force fields that consider the bound and solvated states [see for example Chen et al., 2008; Moon & Howe, 1991], successfully predicting new ligands and also the structures of ligand-receptor complexes [Wilson et al., 1991].

When calculating interactions in congeneric series, the cost in electrostatic free energy of desolvating both the enzyme binding site and the burial part of the ligand (ΔGdesolv) is roughly constant within the series. This is particularly true when the calculation is done partitioning the electrostatic free energy contributions into a van der Waals term from the molecular mechanics force field, and an electrostatic contribution computed using a continuum method [Checa et al., 1997]. For that reason, it has been proposed to neglect ΔGdesolv in earlier studies.

The binding energy between ligand and receptor is approximated to the interaction enthalpy calculated by means of empirical energy functions that represent van der Waals repulsion, dispersion interactions by a Lennard-Jones term, and electrostatic interactions in the form of a Coulomb term that uses atom-centred point charges [Ajay & Murcko, 1995]. In most cases, these calculations of molecular mechanics are performed on a structure that is taken to represent the ensemble average of each complex. Entropy contributions are usually ignored although solvation terms are sometimes added to the scoring function by calculating changes in buried nonpolar surface area [Viswanadhan et al., 1999] or differences in the ease of desolvation of both the ligand and the binding site upon complex formation [Checa et al., 1997]. Molecular mechanics-based QSAR studies on ligand-receptor complexes can benefit greatly from proper incorporation of solvation effects into a COMBINE framework based on residue-based interaction energy decomposition [Pérez et al., 1998].

The relevance of solvation in modulating the biological activity of drugs is well known [Orozco & Luque, 2000]. In the last years, theoretical methods have been developed to calculate fragment contributions to the solvation free energy, particularly in the framework of quantum mechanical (QM) continuum solvation methods [Klamt et al., 2009]. Thus, fractional methods based on GB/SA methods have been developed [Cramer & Truhlar, 2008], as well as those based on the MST(Miertus-Scrocco-Tomasi) solvation method model [Soteras et al., 2004].

An explicit solvent model includes individual solvent molecules and calculates the free energy of solvation by simulating solute-solvent interactions. It requires an empirical interaction potential between the solvent and the solute, and between the solvent molecules, usually involving Monte Carlo (MC) calculations and/or molecular dynamics. MC calculations can be used to compute free energy differences and radial distribution functions, among others, and cannot be used to compute time-dependent properties such as diffusion coefficients or viscosity. MD simulations, on the other hand, can be used to compute free energies and time-dependent properties, transport properties, correlation functions, and others.

An implicit solvent model treats solvent as a polarisable continuum with a dielectric constant, ε, instead of explicit solvent molecules. The charge distribution of the solute polarises the solvent, producing a reaction potential that alters the solute. This interaction is represented by a solvent reaction potential introduced into the Hamiltonian. As interactions should be self consistently computed, they are also known as self-consistent reaction field (SCRF) methods [Onsager, 1936]. These models are significantly easier than explicit solvent models, but cannot model specific interactions such as hydrogen bonds.

Changes in hydration free energy during complex formation are a crucial element of binding free energies [Gilson & Zhou, 2007]. With the use of methods to predict binding free energies becoming common-place in the field of drug design, there is still a need for solvation methods that are both quick and accurate [Mancera, 2007], although much research has been carried out on the improvement of existing methods and development of new solvation models at many levels of theory [Chambers et al., 1996; Gallicchio et al., 2002; Palmer et al., 2011].

Explicit solvation models such as free energy perturbation (FEP), thermodynamic integration (TI) [Gilson & Zhou, 2007; Khavertskii & Wallquist, 2010] and the faster linear interaction energy (LIE) [Aqvist et al., 1994; Carlson & Jorgensen, 1995] offer detail on the distinct nature of water around the solute and are transferable across a wide range of data sets, although there is a lack of throughput in the field of drug design.

Implicit solvation models offer a faster alternative to explicit models by replacing the individual water molecules with a continuous medium [Baker, 2005; Chen et al., 2008], combining the hydration free energy density and group contribution [Jäger & Kast, 2001], or, more recently, calculating solvation free energy directly from the molecular structure [Delgado & Jaña, 2009]. For small organic molecules, the loss of molecular detail of the solvent results in relatively small differences between hydration free energy prediction accuracies calculated with explicit solvent models relative to the explicit treatment [Mobley et al., 2009; Nicholls et al., 2009]. To cope with some of these pitfalls, the variational implicit solvent model (VISM) has been proposed for calculating the solute/water interface where established models fail [Dzubiella et al., 2006a, 2006b].

It is sometimes possible to get quite accurate results with very simple models, such as the case of the molecular modellisation of phenethylamine carriers conducted in our lab. Calculations were carried out using chloride anion to mimic the picrate anion used in experimental measurements and with no explicit solvent molecules. The chloroform environment was simulated by a constant dielectric factor, as this solvent has a low dielectric constant and thus, interactions should not end quickly with the distance [Campayo et al., 2005]. When complexation takes place in water as the solvent, the environment is simulated by a distance-dependent dielectric factor, as it takes into account the fact that the intermolecular electrostatic interactions should vanish with distance faster than in the gas phase. This assumption proves to work as it gives theoretical results in good agreement with experimental transportation values [Miranda et al., 2004; Reviriego et al., 2008]. Results for theoretical interactions have been supported by NMR experiments.

When applied to complex biomolecular systems, this loss of detail may become problematic in locations where water does not behave as a continuum medium, for example, the individual water molecules occurring in concave pockets at the surfaces of proteins [Li & Lazaridis, 2007]. The ELSCA (Energy by Linear Superposition of Corrections Approximation) method [Cerutti et al., 2005] has also been proposed for the rapid estimation of solvation energies. This procedure calculates the electrostatic and apolar solvation energy of bringing two proteins into close proximity or into contact compatible with the AMBER ff99 parameter set. The method is most useful in macromolecular docking and protein association simulations.

Solvent treatment is also of considerable interest in MD simulations as the solvent molecules (usually water, sometimes co-solvent and counterions/buffer or salt for electrolyte solutions) enter pockets and inner cavities of the proteins through their conformational changes. This is a very slow process and nearly as difficult to model as protein solving. One solution to this problem is using an efficient coupling of molecular dynamics simulation with the 3D molecular theory of solvation (3D-RISM-KH), contracting the solvent degrees of freedom [Luchko et al., 2010] or using free energy perturbation and OPLS force field together with molecular dynamics [Shivakumer et al., 2010].

Advertisement

4. Molecule-molecule or ion-molecule interactions in active molecule design

Non-covalent interactions are central to biological structure and function. In considering potential interactions of molecules and/or ions and their receptor, the focus has been on hydrophobic interactions, hydrogen bonding and ion pairing. Although hydrogen bonds are by far the most important interactions in biological recognition processes, the cation-π interaction is a general, strong, non-covalent binding force that occurs throughout nature, being energetically comparable or stronger than a typical hydrogen bond.

Cooperativity in multiple weak bonds (hydrogen bond and ion-π interactions among others) has been considered and studied at the MP2/6-311++ G(d,p) computational level [Alkorta et al., 2010]. Due to the presence of a great number of aromatic rings containing heteroatoms in biological systems, this effect might be important and help to understand some biological processes where the interplay between both interactions may exist.

Computer-assisted drug design (CADD) has contributed to the successful discovery of numerous novel enzyme inhibitors, having been used to predict the binding affinity of an inhibitor designed from a lead compound prior to synthesis [Reddy & Erion, 2005]. A free energy simulation technique known as the thermodynamic cycle perturbation (TCP) approach [Reddy et al., 2007], used together with calculations of molecular dynamics, offers a theoretically precise method to determine the binding free energy differences of related inhibitors.

Many small molecules are transported across cell membranes by large integral membrane proteins, which are referred generically as transporters. Selection among competing alternatives is always interesting and cation-π interactions are strongly involved in substrate recognition by many transporters. Drug transporters are able to carry small molecules or ions across membranes, being an important target for pharmaceutical development [Zacharias & Dougherty, 2002].

The regulation of metal ions plays a major role in enzymes, allowing to catalyse a range of biological reactions. Identification and characterisation of the metal ion binding sites and their selectivity have received immense attention over the past few decades [Ma & Dougherty, 1997]. It is evident from earlier studies that metal ions can bind to aromatic groups in a covalent as well as non-covalent fashion. Non-covalent interactions between metal ions and an aromatic ring, which are considered strong cation-aromatic interactions, are increasingly being recognised as an important binding force relevant to structural biology [Meyar et al., 2003; Elguero et al., 2009]. However, in many cases, the cation is the side chain protonated nitrogen of a basic amino acid. Reddy et al. have made available a web-based cation-aromatic database (CAD) including metal ions and basic amino acids [Reddy et al., 2007b].

Macrocyclic entities that act as ion receptors and carriers exhibit a large number of conformations in crystals and solutions, depending on the nature of their environments and of the complexed ion. To ensure the formation of the most favourable cavity for a given ion, as well as to enhance the binding and release of the ion during transport at interfaces, flexibility in the ligand structure is of utmost importance.

Complexation studies of ions with macrocycles are well documented in the literature. Some representative trends in these studies would include the following: taking into account the existence of hydrogen-bonded water molecules [Hill & Feller, 2000; Durand et al., 2000; Fantoni, 2003] and sometimes using molecular dynamics and free energy perturbation studies [Varnek et al., 1999]. The complexation phenomena have also been studied in cases where the ligand can exist as different conformers able to complex the cation [Hashimoto & Ikuta, 1999].

One of the most important aspects of ion complexation is ion selectivity, which is considered in terms of the more or less favourable binding energies. The binding energy (BE) is defined as the difference between the energy of the complex and the energy of the free ligand and ion:

BE = Ecomplex – (Eion + Eligand)

Metal ion affinity is enhanced if the host molecule has a unique conformation that is optimal for complexation, that is, with all the binding sites positioned to structurally complement the metal ion [Lumetta et al., 2002].

Density functional theory based on electronic structure calculations is computationally affordable. It has very good predictability power for various structural and thermodynamic properties of a molecular system, and has therefore been used to model M+-crown ether complexes [Ali et al., 2008] and collarenes acting as ionophores and receptors [Choi et al., 1998]. In both cases, the most stable equilibrium structure for complexes are estimated based on PM3 semi-empirical calculations followed by B3LYP calculations using the G-311++G(d,p) basis set of functions. Currently, the Protein Data Bank (PDB) contains over 25,000 structures that contain a metal ion. Thus, methodologies to incorporate metal ions into the AMBER force field have been developed there [Hoops et al., 1991; Reichert et al., 2001; Peters et al., 2010].

Molecular modellisation of Cu(II) and Zn(II) coordination complexes has been studied by our research group, among many others. The parameters used by us were checked against a known X-ray structure and the data obtained agreed quite well with similar deviations published for other theoretical results [Miranda et al., 2005]. The models obtained were useful in explaining the differences observed among the complexes obtained in different environments. Our cation metal parameters have also been used to help in the data elucidation of coordination metal complex structures [Rodríguez-Ciria, 2000; Rodríguez-Ciria et al., 2002].

Coordination and complexation of ions by aromatic moieties have been studied, taking into account the different characteristics of the electronic charge distribution on the aromatic frame as an addition of coulombic potentials [Albertí et al., 2010]. This type of interaction is quite common in biology as signalling in the nervous system is generally mediated by the binding of small molecules (neurotransmitters) to the appropriate receptors, which usually contain a cationic group at physiological pH.

An organic ammonium ion never exists as a sole cation; an anion is always associated with it. Depending on the polarity and hydrogen donor/acceptor abilities of the solvent, the association strength is different [Marcus & Hefter, 2006]. Strongly coordinating counter ions such as chloride generally lead to weaker binding constants upon recognition of the associated cation, when compared to weakly coordinating counterions such as iodide or perchlorate [Gevorkyan et al., 2001].

Advertisement

5. Selective formation of complexes

There are several examples of molecular modelling studies on complexes between cyclic receptors and ammonium ions, calixarenes [Choe & Chang, 2002] and crown ethers being the most used. As an example, it is noteworthy to mention the theoretical studies on calix[4]crown-5 and a series of alkyl ammonium ions [Park et al., 2007], having shown that the energy of complex formation depends on the number of amine groups in the alkyl chain as well as on the number of methylene groups between the primary and secondary amine groups, results that agree with experimental measurements. Although the calculations are performed under quite different conditions of vacuum compared with the experimental conditions of the phase system of chloroform-water, the binding properties of calixarene-type compounds towards alkyl ammonium ions have been successfully simulated, providing general and useful explanations for the molecular recognition behaviour.

Complex formation of compounds containing benzene rings with ammonium cations has also been theoretically studied using many computational techniques, including ab initio calculations [Kim et al., 2000]. It has been shown that two types of NH-aromatic π and CH-aromatic π interactions, which are important in biological systems, are responsible for binding, and that charged hydrogen bonds versus cation-π interaction is the origin of the high affinity and selectivity of novel receptors for NH4+ over K+ ions [Oh et al., 2000]. Organic molecules complexed with metal cations have also been studied by MM2 molecular modelling [Mishra, 2010]. The search for metal ion selectivity is of interest in the field of biomimetic models of metalloenzymes and molecular modelling helps in the design of new ligands with this purpose [Kaye, 2011].

Molecular modelling has been used to suggest possible contributions of carrier effectivity and selectivity to complex formation in accordance with experimental results [Chipot et al., 1996; Ilioudis et al., 2005]. Our research group has evaluated the possible cation-receptor interactions involved in the complexes with ammonium and metal cations of selective carriers using the Amber force field with appropriate parameters developed by us. The complexation energies obtained are in reasonable agreement with experimental values, taking into account that complexation/decomplexation processes have a great influence on transport rates and are not equally favoured in cyclic and acyclic carriers [Campayo et al., 2004].

Both binding and selectivity in binding can be understood through the combined efforts of several non-covalent interactions, such as hydrogen bonding, electrostatic interactions, hydrophobic interactions, cation-π interactions, π-π stacking interactions and steric complementarity [Späth & König, 2010]. Formation of complexes is also possible in the case of neutral ligands. For example, the interactions between cholesterol and cyclodextrins have been theoretically studied to investigate their 1:1 and 1:2 complexes [Castagne et al., 2010], while the formation of stable complexes between trehalose and benzene compounds have been investigated by the general Amber force field (GAFF) and Gaussian 03 for MP2/6G-31G** calculation of atomic charges [Sakakura et al., 2011].

Docking of a ligand into a receptor may occur via an automated procedure [Subramanian et al., 2000] or manually [Filizola et al., 1999]. In both cases, docking is a combination of two components: a search strategy and a scoring function [Taylor et al., 2002]. The computational method MOLINE (Molecular Interaction Evaluation) was created to study complexes in an unbiased fashion [Alcaro et al., 2000]. It is based on a systematic, automatic and quasi-flexible docking approach that prevents the influence of the chemist`s intuition on generating the configuration. This method has been used with acceptable results in studying inclusion complexes [Alcaro et al., 2004].

It would be adequate at this point to remember that testing the `drug-receptor complexation’ for a receptor model against available experimental data usually involves the use of site-directed mutagenesis experiments. This fact provides information on the amino acids involved in ligand binding and receptor activation. However, it should be noted that the results of mutagenesis studies are not necessarily related to receptor-ligand interactions. In fact, mutations can also alter the 3D structure of a receptor and therefore, modify the binding profile of a ligand by this mechanism. Besides that, efficient binding to a receptor does not guarantee that a ligand will produce a pharmacological action, given that the ligand may act as an agonist or antagonist.

Advertisement

6. Interaction of molecules with DNA

Anthracycline antibiotics such as doxorubicin and its analogues have been in common use as anticancer drugs for around half a century. There has been intense interest in the DNA-binding sequence specificity of these compounds in recent years, with the hope of identifying a compound that can modulate gene expression or exhibit reduced toxicity. Cashman and Kellog have studied models of binding for doxorubicin and derivatives [Cashman & Kellog, 2004], looking for sequence specificity and the effects of adding aromatic or aliphatic ring substituents or additional amino or hydroxyl groups. They performed a hydropathic interaction analysis using the HINT program (a Sybyl program module, Tripos Inc.) and four double base pair combinations. Interaction of some intercalators with two double DNA base pairs have also been studied with the density functional based tight binding (DFTB) method [Riahi et al., 2010], despite DFT methods being known to be inherently deficient in calculating stacking interactions, and the Amber force field and then AM1 to dock the intercalator between DNA base pairs [Miri et al., 2004].

Studies on sequence-selectivity of DNA minor groove binding ligands have shown that the most reliable results for AT-rich DNA sequences are obtained when MD simulations are performed in explicit solvent, when the data are processed using the MM-PB/SA approach, and when normal mode analysis is used to estimate configurational entropy changes [Shaikh et al., 2004; Wang & Laughton, 2009]. Use of the GB/SE model with a suitable choice of parameters adequately reproduces the structural and dynamic characteristics in explicitly solvated simulations in approximately a quarter of the computational time, although limitations become apparent when the thermodynamic properties are evaluated [Sands & Laughton, 2004]. Water molecules taking part in the complexation have been studied using the MMX force field and then PMB for gas phase optimisation, followed by re-optimisation in aqueous phase with the PM3 method using the AMSOL package [Silva & Jayasundera, 2002]. Optimisation geometries with AM1 and the use of implicit solvent have been taken into account when considering intercalation versus insertion into the minor or major groove [Bendic & Volanschi, 2006].

Calculating the curvature radius of molecular DNA structures has been reported [Slickers et al., 1998] as a new method for understanding the dependence of binding affinity on ligand structure, assuming that strong binders should have a shape complementary to the DNA minor groove. A method for predicting sequence selectivity and minor groove binding, based on MD simulations on DNA sequences with and without the bound ligand, to obtain an approximate free energy of binding has been proposed [Wang & Laughton, 2010].

Amber force field, developing the necessary parameters, has also been used together with electrostatic potential-derived (ESP) charges and explicit solvent molecules to study bisintercalation into DNA. The targeted molecular dynamics (tMD) approach has been considered for comparing the relative energetic cost involved in creating the intercalation sites and also studying the mechanisms of action [Braña et al., 2004]. It has been found that the electrostatic contribution is a critical characteristic of binding selectivity [Marco et al., 2005]. Reports on duplex and triplex formation of oligonucleotides by stacking aromatic moieties in the major groove, using Amber force field and the GB/SA solvation model in molecular dynamic simulations, can be found in the literature [Andersen et al., 2011]. Studies on docking using GOLD [Kiselev et al., 2010] to optimise the starting structures with the MMFF94 force field have also been performed.

Most of the published molecular modelling studies use two double base pairs or more than eight double base pairs to represent DNA. In our opinion, molecular modelling of DNA intercalation complexes should be done using at least the two base pairs of the intercalation site and an additional base pair at the two strand ends to maintain DNA shape and avoid distortion leading to inaccurate results. That means four base pairs for monointercalation studies and five or six base pairs for bisintercalation ones should be used. Using these DNA models, our studies on the mono and bisintercalation of benzo[g]phthalazine derivatives strongly suggest the possibility of bisintercalation and the important role played by an N-methyl group in stabilising the DNA complex of one of the compounds, throwing some light over the experimental results obtained [Rodríguez-Ciria et al., 2003]. The possibility of bisintercalation for a 1,4-disubstituted piperazine has been studied on duplexes of five and six base pairs, obtaining much better results in the case of five base pairs, in accordance with the theoretical calculations of binding mode not conforming to the neighbouring exclusion principle proposed by different authors [Veal et al., 1990].

Advertisement

7. Interaction of small molecules with enzymes

The potential of molecular simulations to enhance our understanding of drug behaviour and resistance relies ultimately on their ability to achieve an accurate ranking of drug binding affinities at clinically relevant time scales. Several computational approaches exist to estimate ligand binding affinities and selectivities, with various levels of accuracy and computational expense: free energy perturbation (FEP), thermodynamic integration (TI), lineal response (LR), and molecular mechanics Poisson-Boltzman surface area (MM/PBSA). Identification of conformational preferences and binding site residues, as well as structural and energetic characterisation, is possible using MD simulations [Anzini et al., 2011; Dastidor et al., 2008; Stoika et al., 2008]. It is also possible to estimate conformational energy penalties for adopting the bioactive conformation identified by using a pharmacophore model [Frølund et al., 2005].

A model based on van der Waals intermolecular contribution from Amber and electrostatic interactions derived from the Poisson-Boltzman equation has been used to predict the change in the apparent dissociation constant for a series of six enzyme-substrate complexes during COMBINE analysis [Kmunicek et al., 2001]. In COMBINE analysis, binding energies are calculated for the set of enzyme-substrate complexes using the molecular mechanics force field. The total binding energy, ΔU, may be assumed to be the sum of five terms: the intermolecular interaction energies between the substrate and each enzyme residue, EinterES, the change in the intramolecular energy of the substrate upon binding to the enzyme, ΔES, the change in the intramolecular energy of the enzyme upon binding, ΔEE, the desolvation energy of the substrate, EdesolvS, and the desolvation energy of the enzyme, EdesolvE.

ΔU = EinterES + ΔES + ΔEE + EdesolvS + EdesolvE

When the substrate is a rather small molecule, there is no evidence for large differences in the structure of the enzyme when different substrates are bound and so the second and third are neglected. This method identifies the amino acid residues responsible for modulating enzyme activity [Kmunicek et al., 2005].

Molecular modelling of proteins is sometimes directed towards homology modelling, enabling progress in understanding the mechanisms of action despite the lack of detailed information on the 3D structure of a protein. Molecular dynamic simulations are usually used to test the stability of the complete structure derived from homology modelling [Srinivas et al., 2006].

Molecular docking examples can be used to compare relative stabilities of the complexes, but not calculate binding affinities, since changes in entropy and solvation effects are not taken into account [Pastorin et al., 2006; Tschammer et al., 2011]. In any case, docking calculations are common studies on novel drugs, Autodock being one of the most used docking programs [see for example Venskutonyte et al., 2011]. Docking programs treat enzymes and substrates as rigid entities, but flexible docking is also possible, if several different protein conformations extracted from molecular dynamic simulations are used [Roumen et al., 2010].

In our laboratory, molecular modelling has been tentatively used to study the trypanosomicidal activity of some phthalazine derivatives. Results obtained with Amber force field implemented in HyperChem 8.0 plus our own necessary parameters, and with AutoDock 4.2 using the PDB structure for T. cruzi Fe-SOD enzyme, were in accordance with experimental data, helping to explain the experimental results obtained. However, if there is no PDB structure for the desired enzyme and only a model of the active site, as for Leishmania Fe-SOD enzyme, results obtained with our calculations do not agree with the experimental ones when compared to the T. cruzi ones. This indicates that the interaction with the external part of the enzyme plays an important role as it might collaborate in, or make access to the active site difficult, since the enzyme shape and conformation plays a crucial role in its activity [Sanchez-Moreno et al., 2011; Yunta, unpublished results].

Advertisement

8. Conclusion

Modern molecular modelling techniques are remarkable tools in the search for potentially novel active agents by helping to understand and predict the behaviour of molecular systems, having assumed an important role in the development and optimisation of leading compounds. Moreover, current information on the 3D structure of proteins and their functions provide a possibility of understanding the relevant molecular interactions between a ligand and a target macromolecule. Although improvements are still needed in the techniques used, they have been shown to be invaluable in structure–activity relationship research.

On the basis of the current improved level of understanding of molecular recognition and the widespread availability of target structures, it is reasonable to assume that computational methods will continue to aid not only the design and interpretation of hypothesis-driven experiments in disease research, but also the fast generation of new hypotheses.

Acknowledgement

Financial support from the Spanish MEC project (CGL2008-0367-E/BOS) and the MCINN projects (CTQ2009-14288-C04-01 and CONSOLIDER INGENIO 2010 CSD2010-00065) are gratefully acknowledged.

References

  1. 1. AjayMurcko, M.A. (1995Computational methods to predict binding free energy in ligand-receptor complexes.J. Med. Chem., 3826December 1995), 495349670022-2623
  2. 2. AlbertíMAguilarALucasJ. MPiraniF2010A generalized formulation of ion-π electron interactions: role of the nonelectrostatic component and probe of the potential parameter transferability. J. Phys. Chem. A, 11444November 2010), 11964119701089-5639
  3. 3. AlcaroSBattagliaDOrtusoF2004Molecular modeling of β-cyclodextrin inclusions complexes with pharmaceutical compounds. Arkivoc, 20045February 2004), 1071171424-6376
  4. 4. AlcaroSGasparriniFIncaniOMecucciSMisitiDPieriniMVillaniC2000AQuasi-flexibleautomatic docking processing for sudying stereoselective recognition mechanisms. Part 1. Protocol validation. J. Comput. Chem., 217May 2000), 5155300109-6987X
  5. 5. AliSk.M., Mainly, D.K., De, S. & Shenoi, M.R.K. (2008Ligands for selective metal ion extraction: a molecular modeling approachDesalination2321-3November 2008), 1811900011-9164
  6. 6. AlkortaIBlancoFDeyaP. MElgueroJEstarellasCFronteraAQuinoneroD2010Cooperativity in multiple unusual weak bonds. Theor. Chem. Acc., 1261-2May 2010), 1141432-2234
  7. 7. AllingerN. L1976Calculation of molecular structure and energy by force field methods, In: Advances in Physical Organic Chemistry, 13Gold, V & Bethell, D. (Eds.), 182Elsevier, 978-0-12033-513-8Amsterdam
  8. 8. AllingerN. LChenKLiiJ. H1996An improved force field (MM4) for saturated hydrocarbonsJ. Comput. Chem., 175-6April 1996), 6426680109-6987X
  9. 9. AllingerN. LDurkinK. A2000Van der Waals effects between hydrogen and first row atoms in molecular mechanics (MM3/MM4)J. Comput. Chem., 2114November 2000), 122912420109-6987X
  10. 10. AllingerN. LYanQ. L1993Molecular mechanics (MM3)- calculations of vinyl ethers, and related compounds. J. Am. Chem. Soc., 115199311918119250002-7863
  11. 11. AndersenN. KDossingHJensenFVesterBNielsenP2011Duplex and triplex formation of mixed pyrimidine oligonucleotides with stacking of phenyl-triazole moieties in the major groove. J. Org. Chem., 7615August 2011), 617761870022-3263
  12. 12. AnziniMValentiSBraileCCappelliAVomeroSAlcaroSOrtusoFMarinelliLLimongelliVNovellinoEBettiLGiannacciniGLucacchiniADanieleSMartiniCGhelardiniCMannelliL. D. CGiorgiGMasciaM. PBiggioG2011New insight into the central benzodiazepine receptor-ligand interactions: design, synthesis, biological evaluation, and molecular modeling of 3-substituted 6-phenyl-4H-imidazo[1,5-a][1,4]benzodiazepines and related compounds. JMed. Chem., 5416August 2011), 569457110022-2623
  13. 13. AqvistJMedinaCSamuelsonJ. E1994A new method for predicting binding affinity in computer aided drug design.Protein Eng. 73March 1994), 3853910269-2139
  14. 14. BakerN. A2005Improving implicit solvent simulations: a poisson-centric viewCurr. Opin. Struct. Biol., 152April 2005), 1371430095-9440X
  15. 15. BendicCVolanschiE2006Molecular modeling of the interaction of some phenoxazone-antitumoral drugs with DNA. Int. Elect. J. Mol. Des., 56June 2006), 3203301538-6414
  16. 16. BissantzCKuhnBStahlM2010A medicinal chemist’s guide to molecular interactionsJ. Med. Chem, 5314July 2010), 506150840022-2623
  17. 17. BranaM. FCachoMGarcíaM. APascual-teresaBRamosADominguezM. TPozueloJ. MAbradeloCRey-stolleM. FYusteMBanez-coronelMLacalJ. C2004New analogues of amonafide and elinafide, containing aromatic heterocycles: synthesis, antitumor activity, molecular modeling, and DNA binding properties. J. Med. Chem., 476March 2004), 139113990022-2623
  18. 18. BrooksB. RBruccoleriR. EOlafsonB. DStalesD. JSwaminathanSKarplusM1983CHARMM: A program for macromolecular energy, minimization and dynamic calculations.J. Comput. Chem., 42July 1983), 1872170109-6987X
  19. 19. BuchISadigS. KDefabritiisG2011Optimized potential of mean force calculations for standard binding free energiesJ. Chem. Theory Comp., 76June 2011), 176517721549-9626
  20. 20. CampayoLCalzadoFCanoM. CYuntaM. J. RPardoMNavarroPJimenoM. LGómez-contrerasFSanzA. M2005New acyclic receptors containing pyridazine units. The influence of π-stacking on the selective transport of lipophilic phenethylamines. Tetrahedron, 6150December 2005), 11965119750040-4020
  21. 21. CampayoLPardoMCotillasAJaúreguiOYuntaM. J. RCanoMGómez-contrerasFNavarroPSanzA. M2004A new series of heteroaromatic receptors containing the 1,3-bis(6-oxopyridazin-1-yl)propane unit: their selective transport ability towards NH4+ in relation to Na+, K+ and Ca2+Tetrahedron604January 2004), 9799860040-4020
  22. 22. CarlsonH. AJorgensenW. L1995An extended linear response method for determining free energies of hydrationJ. Phys. Chem., 9926June 1995), 10667106730022-3654
  23. 23. CashmanD. JKelloggG. E2004A computational model for anthracycline binding to DNA: Tuning groove-binding intercalators for specific sequences.J. Med. Chem., 476March 2004), 136013740022-2623
  24. 24. CastagneDDiveGEvradBFrédérichMPielG2010Spectroscopic studies and molecular modeling for understanding the interactions between cholesterol and cyclodextrinsJ. Pharm. Pharmaceut. Sci., 133July 2010), 3623671482-1826
  25. 25. ČernýV1985Thermodynamical approach to the traveling salesman problem: An efficient simulation algorithm. J. Optimization Theory and Applications, 451January 1985): 41510022-3239
  26. 26. CeruttiD. STen Eyck, L.F. & McCammon, J.A. (2005Rapid estimation of solvation energy for simulations of protein-protein association. J. Chem. Theory Comput., 11January 2005), 1431521549-9626
  27. 27. ChambersC. CHawkinsG. DCramerC. JTruhlarD. G1996Model for aqueous solvation based on class IV atomic charges and first solvation shell effectsJ. Phys. Chem., 10040October 1996), 16385163980022-3654
  28. 28. ChecaAOrtizA. RPascual-teresaBGagoF1997Assessment of solvation effects on calculated binding affinity differences: Trypsin inhibition by flavonoids as a model system for congeneric seriesJ. Med. Chem., 4025December 1997), 413641450022-2623
  29. 29. ChenJBrooksC. LIII & Khandogin, J. (2008Recent advances in implicit solvent-based methods for biomolecular simulationsCurr. Opin. Struct. Biol., 182April 2008), 1401480095-9440X
  30. 30. ChipotCMaigretBPearlmanD. AKollmanP. A1996Molecular dynamics potential of mean force calculations: a study of the toluene-ammonium π-cation interactions. J. Am. Chem. Soc., 11812March 1996), 299830050002-7863
  31. 31. ChoeJ. IChangS. K2002Molecular modeling of complexation behavior of p-tert-butylcalix[5]arene derivative toward butylammonium ions. Bull. Korean Chem. Soc., 231January 2002), 48520253-2964
  32. 32. ChoiH. SSuhS. BChoS. JKimK. S1998Ionophores and receptors using cation-π interactions: collarenes. Proc. Natl. Acad. Sci. USA, 9521October 1998), 1209412099ISNN: 1091-6490
  33. 33. CornellW. DCieplakPBaylyC. IGouldI. RMerzK. MJr., Ferguson, D.M., Spellmeyer, D.C., Fox, T., Caldwell, J.W. & Kollman, P.A. (1995A second generation force field for thr simulation of proteins and nucleic acids. J. Am. Chem. Soc., 11719May 1995), 517951970002-7863
  34. 34. CramerC. JTruhlarD. G2008A universal approach to solvation modeling.Acc. Chem. Res., 416June 2008), 7607680001-4842
  35. 35. CramerR. DPattersonD. EBunceJ. D1988Comparative molecular field analysis (CoMFA). 1. Effect of shape on binding of steroids to carrier proteins. J. Am. Chem. Soc., 11018August 1998), 595959670002-7863
  36. 36. CruzianiGGoodfordP. J1994A research for specificity in DNA-drug interactions. J. Mol. Graph., 122June 1994), 1161290263-7855
  37. 37. DastiderS. GLaneD. PVermaC. S2008Multiple peptide conformations give rise to similar binding affinities: molecular simulations of p53-MDM2. J. Am. Chem. Soc., 13041October 2008), 13514135150002-7863
  38. 38. Dauber-osguthorpePRobertsV. AOsguthorpeD. JWolffJGenestMHaglerA. T2004Structure and energetics of ligand binding to proteins: E. coli dihydrofolate reductase- trimethoprim, a drug-receptor system. Proteins, 41February 2004), 31470887-3585
  39. 39. DelgadoE. JJanaG. A2009Quantitative prediction of solvation free energy in octanol of organic compounds. Int. J. Mol. Sci., 103March 2009), 103110441422-0067
  40. 40. DoshiUHamelbergD2009Reoptimization of the AMBER force field parameters for peptide bond (omega) torsions using accelerated molecular dynamicsJ. Phys. Chem. B, 11352December 2009), 16590165951089-5647
  41. 41. DurandSDognonJ. PGuibandPRabbeCWipffG2000Lanthanide and alkaline-earth complexes of EDTA in water: a molecular dynamics study of structures and binding selectivitiesJ. Chem. Soc., Perkin Trans. 2, 20004April 2000), 7057141364-5471
  42. 42. DzubiellaJSwansonJ. M. JMccammonJ. A2006aCoupling nonpolar and polar solvation free energies in implicit solvent models.J. Chem. Phys., 124February 2006), 0849050021-9606
  43. 43. DzubiellaJSwansonJ. M. JMccammonJ. A2006bCoupling hydrophobicity dispersion, and electrostatics in continuum solvent models.Phys. Rev. Lett., 968March 2006), 0878021079-7114
  44. 44. ElberRRotbergASimmerlingCGoldsteinRLiHVerkhivkerGKeasarCZhangJUlitskyA1995MOIL: A program for simulations of macromoleculesComp. Phys. Comm., 911-3September 1995), 1591891815-2406
  45. 45. ElgueroJAlkortaIClaramuntR. MLópezCSanzDSanta María, D. (2009Theoretical calculations of a model of NOS indazole inhibitors: Interaction of aromatic compounds with Zn-porphyrinsBioorg. Med. Chem., 1723December 2009), 802780310968-0896
  46. 46. FantoniA. C2003Molecular dynamics study of geometrical isomers of a pyridinocalix[4]arene in methanol solution: solvation and alkali metal cation binding propertiesJ. Mol. Struct. (Theochem), 6931August 2003), 160166-1280
  47. 47. FilizolaMCarteri-farinaMPerezJ. J1999Molecular modeling study of the differential ligand-receptor interaction at the μ, δ and κ opioid receptors. J. Comput. Aid. Mol. Des., 134July 1999), 3974071573-4951
  48. 48. FriesnerR. ABanksJ. LMurphyR. BHalgrenT. AKlicicJ. JMainzD. TRepaskyM. PKnollE. HShelleyMPerryJ. KShawD. EFrancisPShenkinP. S2004Glide: A new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracyJ. Med. Chem., 477March 2004), 173917490022-2623
  49. 49. FriesnerR. AMurphyR. BRepaskyM. PFryeL. LGreenwoodJ. RHalgrenT. ASanschagrinP. CMainzD. T2006Extra precision Glide: Docking and scoring incorporating a model of hydrophobic enclosure for protein-ligand complexes.J. Med. Chem., 4921October 2006), 617761960022-2623
  50. 50. FrolundBJensenL. SGuandaliniLCanilloCVestergarardH. TKristiansenUNielsenBStensbolT. BMadsenCKrogsgaard-larsenPLiljeforsT2005Potent 4-aryl- or 4-arylalkyl-substituted 3-isoxazolol GABAA antagonists: synthesis, pharmacology, and molecular modeling. J. Med. Chem., 482January 2005), 4274390022-2623
  51. 51. GalisteoJNavarroPCampayoLYuntaM. J. RGómez-contrerasFVilla-pulgarinJ. ASierraB. GMollinedoFGonzalezJGarcía-espanaE2010Synthesis and cytotoxic activity of a new potential bisintercalator: 1,4-bis{3-[N-(4-chlorobenzo[g]phthalazin-1-yl)aminopropil]}piperazine. Bioorg. Med. Chem., 1814July 2010), 530153090968-0896
  52. 52. GallicchioELevyR. M2004AGBNP: An analytical implicit solvent model suitable for molecular dynamics simulations and high-resolution modeling. J. Comput. Chem., 254March 2004), 4794990109-6987X
  53. 53. GallicchioEZhangL. YLevyR. M2002The SGB/NP hydration free energy model based on the surface generalized born solvent reaction field and novel nonpolar hydration free energy estimators.J. Comput. Chem., 235April 2002), 5175290109-6987X
  54. 54. GevorkyanA. AArakelyanA. SEsayanV. APetrosyanK. ATorosyanG. O2001Ionic character of the ammonium-counterion bond and catalytic activity of ammonium salts in elimination reactionsGen. Chem., 718August 2001), 132713281070-3632
  55. 55. GilsonM. KZhouH. X2007Calculation of protein-ligand binding affinities.Annu. Rev. Biophys. Biomol. Struct., 36June 2007), 21421056-8700
  56. 56. GoodfordP. J1985A computational procedure for determining energetically favorable binding sites on biologically important macromolecules.J. Med. Chem., 188August 1985), 8498570022-2623
  57. 57. HalgrenT. A1996Merck molecular force field. 1. Basis, form, scope, parameterización, and performanceof MMFF9J. Comput. Chem., 175-6April 1996), 4905190109-6987X
  58. 58. HalgrenT. AMurphyR. BFriesnerR. ABeardH. SFrieL. LPollardW. TBanksJ. L2004Glide: A new approach for rapid, accurate docking and scoring. 2. Enrichment factors in database screening.J. Med. Chem., 477March 2004), 175017590022-2623
  59. 59. HashimotoSIkutaS1999A theoretical study on the conformations, energetics, and solvation effects on the cation-π interaction between monovalent ions Li+, Na+ and K+ and naphthalene molecules. J. Mol. Struct. (Theochem), 4681-2August 1999), 85940166-1280
  60. 60. HillS. EFellerD2000Theoretical study of cation/ether complexes: 15-crown-5 and its alkali metal complexesInt. J. of Mass Spect., 2011-3December 2000), 41581387-3806
  61. 61. HoopsS. CAndersonK. WMerzK. MJr. (1991Force field design for metalloproteinsJ. Am. Chem. Soc., 11322October 1991), 826282700002-7863
  62. 62. HortaB. A. CFuchsP. F. JVan GunsterenW. FHunenbergerP. H2011New interaction parameters for oxygen compounds in the GROMOS force field: improved pure-liquid and solvation properties for alcohols, ethers, aldehydes, ketones, carboxylic acids and esters.J. Chem. Theory Comp., 74April 2011), 101610311549-9626
  63. 63. IlioudisC. ABearparkM. JSteadJ. W2005Hydrogen bonds between ammonium ions and aromatic rings exist and have key consequences on solid-state and solution phase propertiesNew J. Chem., 291January 2005), 64671144-0546
  64. 64. JagerRKastS. M2001Fast prediction of hydration free energies from molecular interaction fields.J. Mol. Graphs. Mod., 201February 2001), 1231311093-3263
  65. 65. JainA. N2003Surflex: Fully automatic flexible molecular docking using a molecular similarity-based search engine.J. Med. Chem., 464February 2003), 4995110022-2623
  66. 66. JainA. N2007Surflex-Dock 2.1: Robust performance from ligand energetic modeling, ring flexibility, and knowledge-based searchJ. Comput. Aid. Mol. Des., 215May 2007), 2813061573-4951
  67. 67. JorgensenW. LTirado-rivesJ1996Monte Carlo vs molecular dynamics for conformational samplingJ. Phys. Chem., 10034August 1996), 14508145130022-3654
  68. 68. KayeP. T2011Designer ligands: The search for the metal ion selectivity. S. Afr. J. Sci., 1073-4March 2011), 4394460038-2353
  69. 69. KhavretskiiI. VWallquistA2010Computing relative energies of solvation using single reference thermodynamic integration augmented with Hamiltonian replica exchange. J. Chem. Theor. Comput., 611342734411549-9626
  70. 70. KimK. SLeeJ. YTarakeshwarP2000Molecular clusters of π-systems: Theoretical studies of structures, spectra, and origin of interaction energies. Chem. Rev., 10011November, 2000), 414541860009-2665
  71. 71. KirkpatrickSGelattC. DVecchiM. P1983Optimization by Simulated AnnealingScience, 2204598May 1983) 6716800036-8075
  72. 72. KiselevEDexheimerTPommierYCushmanM2010Design, synthesis, and evaluation of dibenzo[c,h][1,6]naphthyridines as topoisomerase I inhibitors and potential anticancer agents. J. Med. Chem., 5324December 2010), 871687260022-2623
  73. 73. KlamtAMennucciBTomasiJBaroneVCurutchetCOrozcoMLuqueF. J2009On the performance of continuum solvation methods. A comment on ‘universal approaches to solvation modeling’Acc. Chem. Res., 424April 2009), 4894920001-4842
  74. 74. KmunicekJLuengoSGagoFOrtízA. RWadeR. CDamborskýJ2001Comparative binding energy analysis of the substrate specificity of haloalkane dehalogenase from Xanthobacter autotrophicus GJ10.Biochem., 4030July 2001), 890589170006-2960
  75. 75. KmunicekJHyncováKJedlickaTNagataYNegriAGagoFWadeR. CDamborskýJ2005Quantitative analysis of substrate specificity of haloalcane dehalogenase Lin B from Sphingomonas paucimobilis UT26. Biochem., 4410March 2005), 339034010006-2960
  76. 76. KollmanP. AMassovaIReyesCKuhnBHuoS2000Calculating structures and free energies of complex molecules: combining molecular mechanics and continuum models.Acc. Chem. Res., 3312December 2000), 8898970001-4842
  77. 77. KuntzI. D1992Structure based strategies for drug design and discovery. Science, 2578August 1992), 107810820036-8075
  78. 78. KuntzI. DMengE. CShoichetB. K1994Structure based molecular design. Acc. Chem. Res., 275May 1994), 1171230001-4842
  79. 79. LiZLazaridisT2007Water at biomolecular binding interfacesPhys. Chem. Chem. Phys., 95February 2007), 5735811463-9076
  80. 80. LinFWangR2010Systematic derivation of AMBER force field parameters applicable to zinc-containing systemsJ. Chem. Theory Comput., 66June 2010), 185218701549-9626
  81. 81. LuchkoTGusarovSRoeD. RSimmerlingCCaseD. ATuszynskiJKovalenkoA2010Three-diemnsional molecular theory of solvation coupled with molecular dynamics in Amber. J. Chem. Theory Comput., 63March 2010), 6076241549-9626
  82. 82. LumettaG. JRapkoB. MGarzaP. AHayB. P2002Deliberate design of ligand architecture yields dramatic enhancement of metal ion affinity.J. Am. Chem. Soc., 12420May 2002), 564456450002-7863
  83. 83. MaJ. CDoughertyD. A1997The cation-π interaction. Chem. Rev., 975August 1997), 130313240009-2665
  84. 84. ManceraR. L2007Molecular modeling of hydration in drug design.Curr. Opin. Drug. Discov. Dev., 103May 2007), 2752801367-6733
  85. 85. MarcoENegriALuqueF. JGagoF2005Role of staking interactions in the binding sequence preferences of DNA bis-intercalators: insight from thermodynamic integration free energy simulations. Nuc. Ac. Res., 3319November 2005), 621462240305-1048
  86. 86. MarcusYHefterG2006Ion pairing. Chem. Rev., 10611November 2006), 458546210009-2665
  87. 87. MengE. CShoichetB. KKuntzI. D1992Automated docking with grid-based energy evaluationJ. Comput. Chem., 134May 1992), 5055240109-6987X
  88. 88. MeyarE. ACastellanoR. KDiederichF2003Interactions with aromatic rings in chemical and biological recognition. Angew. Chem. Int. Ed., 4211March 2003), 121012501433-7851
  89. 89. MirandaCEscartíFLamarqueLYuntaM. J. RNavarroPGarcía-espanaEJimenoM. L2004New 1H-pyrazole-containing polyamine receptor sable to complex L-glutamate in wáter at physiological pH values. J. Am. Chem. Soc., 1263January 2004), 8238330002-7863
  90. 90. Miranda, C., Escartí, F., Lamarque, L., García-España, E., Navarro, P., Latorre, L., lloret, F., Jimenez, H.R. & Yunta, M.J.R. (2005). CuII and ZnII coordination chemistry of pyrazole-containing poliamine receptors- Influence of the hydrocarbon side chain length on the metal coordination. Eur. J. Inorg. Chem., Vol. 2005, No. 1, (January 2005), pp. 189-208. ISSN: 1434-1948
  91. 91. MiriRJavidniaKHemmateenejadBAzarpiraAAmirghofranZ2004Synthesis, cytotoxicity, QSAR, and intercalation study of new diindenopyridine derivatives. Bioorg. Med. Chem., 1210May 2004), 252925360968-0896
  92. 92. MishraP2010Biocoordination and computational modeling of streptomycin with Co(II), Ni(II), In(II) and inorganic Sn(II). Int. J. Pharm. Sci., 22June 2010), 87970097-6044X
  93. 93. MobleyD. LBaylyC. ICooperM. DShirts & Dill, K.A. (2009Small molecule hydration free energies in explicit solvent: an extensive test of fixed-cahrge atomistic simulations. J. Chem. Theory Comput., 52February 2009), 3503581549-9626
  94. 94. MoonJ. BHoweW. J1991Computer design of bioactive molecules: a method for receptor-based de novo ligands design. Proteins, 114December 1991), 3143280887-3585
  95. 95. NichollsAWlodekSGrantJ. A2009The SAMP1 solvation challenge: Further lessons regarding the pitfalls of parameterization. J. Phys. Chem. B, 11314April 2009), 452145321089-5647
  96. 96. OhK. SLeeC. WChoiH. SLeeS. JKimK. S2000Origin of the high affinity and selectivity of novel receptors for NH4+ over K+: Charged hydrogen bonds vs cation-π interaction.Org. Lett., 217August 2000), 267926811523-7052
  97. 97. OnsagerL1936Electric moments of molecules in liquids, J. Am. Chem. Soc., 588August 1936), 148614930002-7863
  98. 98. OrozcoMLuqueF. J2000Theoretical methods for the description of the solvent effect in biomolecular systemsChem. Rev., 10011November 2000), 418742260009-2665
  99. 99. PalmerD. SFrolovA. IRatkovaE. LFedorovM. V2011Toward a universal model to calculate the solvation thermodynamics of druglaike molecules: The importance of new experimental databases. Mol. Pharm., 84August 2011), 142314290002-6895X
  100. 100. ParkJ. YKimB. CParkS. M2007Molecular recognition of protonated polyamines at calix[4]crown-5 self-assembled monolayer modified electrodes by impedance measurements.Anal. Chem., 795March 2007), 189018960003-2700
  101. 101. PastorinGDa Ros, T., Bolcato, C., Montopoli, C., Moro, S., Cacciari, B., Baraldi, P.G., varani, K., Borea, P.A. & Spalluto, G. (2006Synthesis and biological studies of a new series of 5-heteroarylcarbamoylaminopirazolo[4,3-e]1,2,4-triazolo[1,5-c]pyrimidines as human A3 adenosine receptor antagonists. Influence of the heteroaryl substituent on binding affinity and molecular modeling investigations. J. Med. Chem., 495March 2006), 172017290022-2623
  102. 102. PerezCPastorMOrtizA. RGagoF1998Comparative binding energy analysis of HIV-1 protease inhibitors: incorporation of solvent effects and validation as a powerful tool in receptor-based drug design.J. Med. Chem., 416March 1998), 8368520022-2623
  103. 103. PetersM. BYangYWangBFüsti-molnárLWeaverM. NMerzK. MJr. (2010Structural survey of zinc-containing proteins and development of the zinc AMBER forcefield (ZAFF). J. Chem.Theory Comput. 69September 2010), 293529471549-9626
  104. 104. PlewczynskiDLazniewskiMAugustyniacRGinalskiK2011Can we trust docking results? Evaluation of seven commonly used programs on PDB bind database. J. Comput. Chem., 324March 2011), 7427550109-6987X
  105. 105. ReddyM. RErionM. DAgaewalA2000Use of free energy calculations in drug design, In: Reviews in computational chemistry 2. K.B. Lipkowitz & D.B. Boyd (Eds.), 217304978-0-47118-810-0
  106. 106. ReddyM. RErionM. D2005Computer aided drug design strategies used in the discovery of fructose 1,6-biphosphate inhibitors. Curr. Pharm. Des., 113February 2005), 2832941381-6128
  107. 107. ReddyA. SSastryG. MSastryG. N2007Cation-aromatic database.Proteins674March 2007), 41580887-3585
  108. 108. ReddyM. RErionM. D2007Relative binding affinities of fructose-1,6-bisphosphatase inhibitors calculated using a quantum mechanics-based free energy perturbation method. J. Am. Chem. Soc., 12930August 2007), 929692970002-7863
  109. 109. ReichertD. ENorrbyP-OWelchM. J2001Molecular modeling of bifunctional chelate peptide conjugates. 1. Copper and indium parameters for the Amber force field.Inorg. Chem., 4020September 2001), 522352300020-1669
  110. 110. ReviriegoFNavarroPGarcía-espanaEAlbeldaM. TFriasJ. CDomenechAYuntaM. J. RCostaROrtíE2008Diazatetraester 1H-pyrazole crowns as fluorescent chemosensors for AMPH, METH, MDMA(Ecstasy) and dopamine. Org. Lett., 1022November 2008), 509951021523-7060
  111. 111. ReynoldsC. AWadeR. CGoodfordP. J1989Identifying targets for bioreductive agents: using GRID to predict selective binding regions of proteins.J. Mol. Graph., 72June 1989), 1031080263-7855
  112. 112. RiahiSEynollahiSGanjaliM. RNorouziP2010Computational modeling of interaction between Camphothecin and DNA base pairs. Int. J. Electrochem. Sci., 58August 2010), 115111631452-3981
  113. 113. Rodríguez-ciriaM2000Síntesis De 1-aminoy14diamino derivados de benzo[g]ftalazina, evaluación de sus propiedades complejantes frente a cationes metálicos y catecolaminas involucrados en mecanismos de neurotrnasmisión. Ph.D. Thesis, Universidad Complutense, Madrid
  114. 114. Rodríguez-ciriaMSanzA. MGómez-contrerasFNavarroPPardoMYuntaM. J. RCastineirasACanoM. C2002Benzo[g]phthalazine ligands as tyrosinase mimetics: the influence of the polyaminic side-chains size and nature on the complexation of Cu(II). Proceedings of 8th International symposium on the chemistry and pharmacology of pyridazines, Ferrara (Italy), October 2002
  115. 115. Rodríguez-ciriaMSanzA. MYuntaM. J. RGómez-contrerasFNavarroPFernándezIPardoMCanoM2003Synthesis and cytostatic activity of N,N-bis-{3-[N-(4-chlorobenzo[g]-phthalazin-1-yl]aminopropil}-N-methylamine: a new potential DNA bisintercalator. Bioorg. Med. Chem., 1110May 2003), 214321480968-0896
  116. 116. RoumenLPeetersJ. WEmmenJ. M. ABengelsI. P. ECustersE. M. GDe GooyerMPlateRPieterseKHilbersP. A. JSmitsJ. F. MVekemansJ. A. JLeysenDOttenheijmH. C. JJanssenH. MHermansJ. J. R2010Synthesis, biological evaluation, and molecular modeling of 1-benzyl-1H-imidazoles as selective inhibitors of aldosterone synthase (CYP11B2J. Med. Chem., 534February 2010), 171217250022-2623
  117. 117. SakakuraKOkabeAOkuKSakuraiM2011Experimental and theoretical study on the intermolecular complex formation between trehalose and benzene compounds in aqueous solutionJ. Phys. Chem. B, 11532August 2011), 982398301089-5647
  118. 118. Sanchez-morenoMSanzA. MGómez-contrerasFNavarroPMarínCRamírez-macíasIRosalesM. JOlmoFGarcía-arandaICampayoLCanoCArrebolaFYuntaM. J. R2011In vivo Trypanosomicidal activity of imidazole-or pyrazole-based venzo[g]phthalazine derivatives against acute and chronic phases of chagas disease. J. Med. Chem., 544February 2011), 9709790223-5234
  119. 119. SandsZ. ALaughtonC. A2004Molecular dynamics simulations of DNA using the generalized Born solvation model: quantitative comparisons with explicit solvation resultsJ. Phys. Chem. B, 10828July 2004), 10113101191089-5647
  120. 120. ShaikhS. AAhmedS. RJayaramB2004A molecular thermodynamic view of DNA-drug interactions: A case study of 25 minor-groove binders.Arch. Biochem. Biophys., 4291September 2004), 81990003-9861
  121. 121. ShivakumarDWilliamsJWuYDammWShelleyJShermanW2010Prediction of absolute solvation free energies using molecular dynamics free energy perturbation and the OPLS force fieldJ. Chem. Theory Comput., 65May 2010), 150915191549-9626
  122. 122. ShoichetB. KKuntzI. D1996Predicting the structure of protein complexes: a step in the right direction.Chem. and Biol., 33March 1996), 1511561074-5521
  123. 123. SilvaS. JJayasunderaK2002Quantitative structure activity relationships for guanidiniothiazole carboxamides using theoretically calculated molecular descriptors. J. Natn. Sci. Found. Sri Lanka, 303-4December 2002), 1711841391-4588
  124. 124. SimonsonT2001Macromolecular electrostatics: continuum models and their growing painsCurr. Opin. Struct. Biol., 112April 2001), 2432520095-9440X
  125. 125. SlickersPHillebrandMKittlerLLöberGSühnelJ1998Molecular modeling and footprinting studies of DNA minor groove binders: bisquaternary ammonium heterocyclic compounds.Anti-Cancer Drug Des., 135September 1998), 4634880266-9536
  126. 126. SoterasIMorrealeALópezJ. MOrozcoMLuqueF. J2004Group contributions to the solvation free energy from MST continuum calculationsBraz. J. Phys., 341March 2004), 48571678-4448
  127. 127. SpäthAKönigB2010Molecular recognition of organic ammonium ions in solution using synthetic receptors.Beilstein J. Org. Chem., 632April 2010), 11111860-5397
  128. 128. SrinivasEMurthyJ. NRaoA. R. RSastryG. N2006Recent advances in molecular modeling and medicinal chemistry aspects of phosphor-glycoprotein. Curr. Drug Metabol., 72February 2006), 2052171389-2002
  129. 129. StoikaISadiqS. KCoveneyP. V2008Rapid and accurate prediction of binding free energies for saquinavir-bound HIV-1 proteases. J. Am. Chem. Soc., 1308February 2008), 263926480002-7863
  130. 130. SubramanianGPaterliniM. GPortogheseP. SFergusonD. M2000Molecular docking reveals a novel binding site model for fentanyl at the µ-opioid receptor. J. Med. Chem., 433February 2000), 3813910022-2623
  131. 131. TaylorR. DJewsburyP. JEssexJ. W2002A review of protein-small molecule docking methods.. J. Comput. Aid. Mol. Des., 163March 2002), 1511661573-4951
  132. 132. TóthJRemkoMNagyM2005The ability of molecular modeling methods to reproduce the structure of flavonoids. Acta Facul. Pharm. Univ. Comenianae, Vol. LII, (2005), 2182250301-2298
  133. 133. TschammerNElsnerJGoetzAEhrlichKSchusterSRubergMKühhornJThompsonDWhistlerJHübnerHGmeinerP2011Highly potent 5-aminotetrahydropyrazolopyridines: enantioselective dopamine D3 receptor binding, functional selectivity, and analysis of receptor-ligand interactions. J. Med. Chem., 547April 2011), 247724910022-2623
  134. 134. Van GunsterenW. FBerendsenH. J. C1977Algorithms for macromolecular dynamics and constraint dynamicsMol. Phys., 345August 2006), 131113271362-3028
  135. 135. VarnekAWipffGBilykAHarrowfieldJ. M1999Molecular dynamics and free energy perturbation studies of Ca2+/Sr2+ complexation selectivities of the macrocyclic ionophores DOTA and TETA in waterJ. Chem. Soc. Dalton Trans., 199923December 1999), 415541641472-7773
  136. 136. VealJ. MLiXZimmermanS. CLambenmonC. RCoryMZonGWilsonW. D1990Interaction of a macrocyclic bisacridine with DNA.Biochem., 2949December 1990), 10918109270006-2960
  137. 137. VenskutonyteRButiniSCocconeS. SGemmaSBrindisiMKumorVGuarinoEMaramaiSAmirAValadesE. AFrydenvangKKastrupJ. SNovellinoECampianiGPickeringD. S2011Selective kainite receptor (Gluk1) ligands structurally based upon 1H-cyclopentapyrimidin-2,4(1H,3H)-dione: Synthesis, molecular modeling, and pharmacological and biostructural characterization. J. Med. Chem., 5413July 2011), 479348050022-2623
  138. 138. ViswanadhanV. NGhoseA. KSingU. CWendoloskiJ. J1999Prediction of solvation free energies of small organic molecules:additive-constitutive models based on molecular fingerprints and atomic constants.J. Chem. Inf. Comput. Sci., 392March 1999), 4054120095-2338
  139. 139. WangJKangXKuntzI. DKollmanP. A2005Hierarchical database screenings for HIV-1 reverse transcriptase using a pharmacophore model, rigid docking and MM-PB/SA. J. Med. Chem.,488April 2005), 243224440022-2623
  140. 140. WangHLaughtonC. A2009Evaluation of molecular modeling methods to predict the secuence-selectivity of DNA minor groove binding ligands. Phys. Chem. Chem. Phys.. 1145December 2009), 10722107281463-9076
  141. 141. WangHLaughtonC. A2010Molecular modeling mrthods to quantitative drud-DNA interactions, In: Drug-DNA interaction protocols, Methods in molecular biology, 6131931Humana Press, Germany. 978-1-60327-417-3
  142. 142. WeinerS. JKollmanP. ACaseD. ASinghV. CGhioCAlagonaGProfetaSJr. & Weiner, P. (1984A new force field for molecular mechanical simulation of nucleic acids and proteinsJ. Am. Chem. Soc., 1063February 1984), 7657840002-7863
  143. 143. WeinerS. JKollmanP. ANguyenD. TCaseD. A1986An all atom force field for simulations of proteins and nucleic acidsJ. Comput. Chem., 72April 1986), 2302520109-6987X
  144. 144. WilsonCMaceJ. EAgardD. A1991A computational method for designing enzymes with altered substrate specifity. J. Mol. Biol., 2202July 1991), 4955060022-2836
  145. 145. WoodsR. JDwekR. AEdgeC. JFraser-reidB1995Molecular mechanical and molecular dynamical simulations of glycoproteins and oligosaccharides. 1. GLYCAM_93 parameter developmentJ. Phys. Chem., 9911March 1995), 383238460022-3654
  146. 146. YangLTanCHsiehM-JWangJDuanYCieplakPCaldwellJKollmanP. ALuoR2006New generation Amber united-atom force field. J. Phys. Chem. B, 11026July 2006), 13166131761089-5647
  147. 147. ZachariasNDoughertyD. A2002Cation-π interactions in ligand recognition and catalysis. Trends Pharm. Sci., 236June 2002), 2812870165-6147
  148. 148. ZgarbováMOtyepkaMSponerJMládekABanésPCheathamT. EIII & Jurecka, P. (2011Refinement of the Cornell et al. nucleic acids force field based on reference quantum chemical calculations of glycosidic torsion profilesJ. Chem. Theory Comp., 79September 2011), 288629021549-9626

Written By

María J. R. Yunta

Submitted: 03 October 2011 Published: 28 November 2012