Elsevier

Drug Discovery Today

Volume 20, Issue 6, June 2015, Pages 686-702
Drug Discovery Today

Review
Foundation
The impact of molecular dynamics on drug design: applications for the characterization of ligand–macromolecule complexes

https://doi.org/10.1016/j.drudis.2015.01.003Get rights and content

Highlights

  • Which molecular dynamics (MD) techniques are available for drug design?

  • How is MD used to investigate ligand–macromolecule complexes?

  • How are MD studies applied to human and non-human therapeutic targets?

  • What are the latest advances in the field of MD?

Among all tools available to design new drugs, molecular dynamics (MD) simulations have become an essential technique. Initially developed to investigate molecular models with a limited number of atoms, computers now enable investigations of large macromolecular systems with a simulation time reaching the microsecond range. The reviewed articles cover four years of research to give an overview on the actual impact of MD on the current medicinal chemistry landscape with a particular emphasis on studies of ligand–protein interactions. With a special focus on studies combining computational approaches with data gained from other techniques, this review shows how deeply embedded MD simulations are in drug design strategies and articulates what the future of this technique could be.

Introduction

Molecular dynamics (MD) has become a new major technique in the arsenal of tools developed to design novel bioactive molecules and investigate their mode of action. The main experimental technique to elucidate the molecular structures of macromolecules is the X-ray diffraction of crystallized protein. This technique has led to great achievements in the field of structural biology; but, in the best cases, X-ray crystallography can only provide a static snapshot of a fully functional state. In the past decade, NMR has become an increasingly important technique for protein structural investigations, giving access to the flexibility of a system by revealing an ensemble of conformations. But, despite giant steps achieved lately in the field, NMR spectroscopy remains challenging and time consuming for large protein complexes. Consequently, complementary tools have always been required to allow a dynamic insight into biological targets and ligand binding. MD is an in silico method that can utilize structural data gained experimentally to extrapolate the possible conformations of molecular systems and the different paths between each of them.

Computationally driven simulations of constrained molecular systems were initially used to investigate models with a limited number of atoms. But an impressively fast technological development made possible an astonishing increase of computational speed and data storage volume over the past decade. Today, investigations of larger systems, such as receptors fully incorporated within their biological environment, are routinely conducted. The simulation time being constantly extended, current studies report simulations in the microsecond range, enabling more relevant events to be observed in silico for the rationalization of experimental data.

The scientific literature published in the past four years led to a surprisingly high amount of studies reporting MD simulations (>750 articles – using the very restrictive keywords ‘drug’ and ‘molecular dynamics’). If this analysis of the current situation acknowledges that MD has become a major technique, it remains unclear how MD simulations are used and what unique information they give access to. To answer these questions, we have reviewed and discussed peer-reviewed articles published since 2011 in various fields of medicinal chemistry. In the next section, a short overview on the tools and methods available in the field of MD is given. Then, studies involving MD at a key step of the investigation of therapeutically relevant targets with a ligand are reported. The reviewed strategies are discussed and compared for each main class of enzyme structure investigated in the literature, including separated sections for nucleic acids and targets from pathogenic microorganisms. We finally close this review with a discussion on the weaknesses of the technique and exceptional studies focusing on the improvement of these specific points. This review aims at giving an overall picture of the actual impact of MD on the current landscape in medicinal chemistry and to anticipate what the future of this technique could be.

The most established computational tools available for classical MD simulations are the software suites GROMACS [1], NAMD [2], CHARMM [3] and AMBER [4] (Table 1). MD simulations are a numerical solution of Newton's equations of motion (Eq. (1)):fi=miai=V(rN)riwhere mi is the mass of atom i, ai the acceleration of atom i and fi is the force acting on atom i given by the partial, spatial derivative of the potential energy function V which is dependent on the positions rN = (r1, r2, …, rN) of all N particles in the system. MD simulations generate a time series of configurations of a system, which allows studying dynamic processes. The time series can also be viewed as a statistical-mechanical ensemble of configurations. By averaging over this ensemble, thermodynamic properties of a system can be accessed.

Four basic elements define the model that is simulated by MD: (i) resolution of the model; (ii) description of the interactions between atoms; (iii) generation of configurations; and (iv) definition of boundary conditions. The resolution defines the ‘elementary particles’ of the model [5]. These can be nuclei and electrons in quantum-mechanical (QM) simulations, atoms in classical MD simulations or coarse-grained (CG) beads where multiple atoms are described by a single particle. The computational demand decreases and thus the accessible time and length scale increases in this order. All three resolution levels and mixtures of them (i.e. QM/MM or hybrid atomistic/CG simulations) are relevant for computer-aided drug design. QM/MM simulations allow the study of enzymatic reactions, reaction pathways, energy barriers and proton and electron reshuffling in larger systems, where the active site is treated quantum-mechanically while the other parts of the system can be simulated at the atomic level 6, 7, 8, 9, 10. Coarse-graining of atomic models is a popular technique to access longer time scales in MD simulations. By embedding multiple atoms in a single CG bead, the number of particle–particle interactions to be calculated can be reduced considerably. However, several issues arise with this technique. First of all, coarse-graining involves a loss of information, therefore only degrees of freedom that are less important for the question of interest should be coarse-grained. Second, no general procedure to select the type and number of atoms per CG bead exists, which leads to a large number of CG models being proposed [11]. For drug design, where a high resolution of the ligand and active site is likely to be important, atomistic/CG hybrid or multiscale models are particularly interesting [12].

In classical MD simulations, the potential energy function commonly denoted as the ‘force-field’ describes the interactions between the atoms (or generally particles). It consists of so-called bonded terms between covalently bound atoms (i.e. bonds, angles, torsions) and nonbonded terms (i.e. van der Waals interactions and electrostatic interactions). The parameters of the force-field are typically fitted to reproduce data from higher-level calculations and/or experiments. For the majority of biomolecules such as proteins, DNA, lipids and sugars, a relatively small number of building blocks can be used, which have to be parameterized only once. Traditional biomolecular force-fields such as AMBER [13], OPLS [14], CHARMM [15] and GROMOS [16] have been continuously improved and tested over the years. The diversity of small organic molecules by contrast prevents the use of building blocks, and ligands have to be parameterized individually. Parameters for the bonded terms and the van der Waals interactions are typically taken from a generalized force-field, whereas the partial atomic charges are derived from a QM calculation. The force-fields most commonly used for this purpose are the general AMBER force-field (GAFF) [17], the OPLS all-atom force-field (OPLS-AA) [18], the CHARMM general force-field (CGenFF) 19, 20 and the GROMOS automatic topology builder [21].

The resulting MD trajectory of a ligand–macromolecule complex is analyzed to extract information such as distances and interactions between atoms or residues of interest. Additional values are measured to control the overall stability of the complex, such as the root-mean-square deviations (RMSD) from a reference configuration, the atom-positional root-mean-square fluctuations (RMSF) or the torsion angle distributions of the protein backbone. Additionally, an important application of MD simulations – especially in the context of drug design – is the estimation of free energy differences, ΔG. The free energy of binding, ΔGbind, is the difference between the free energy of the ligand free in solution and bound to the protein, and is directly correlated to the binding constant, Ki (Eq. (2)):ΔGbind=GcomplexGfree=RTln(Ki)

Because the direct estimation of ΔGbind from unbiased MD simulations is difficult, the difference in ΔGbind between two ligands A and B is often calculated. This quantity, ΔΔGBAbind, is accessible by MD simulations through so-called ‘alchemical’ transformations between ligand A and B. The rigorous methods for free energy calculation using MD such as thermodynamic integration (TI) [22], free energy perturbation (FEP) [23] and Bennett acceptance ratio (BAR) [24] belong to the most accurate, but also most time-consuming, techniques. Among these, TI is probably the most popular and robust method. In TI, the system is perturbed in small steps along an artificial ‘reaction coordinate’ λ from state (ligand) A to B, and the resulting curve is integrated to yield ΔGBA (Eq. (3)):ΔGBA=G(λB)G(λA)=λAλBH(λ)λλdλwhere H(λ) is the λ-dependent Hamiltonian of the system.

Alternative approaches that are more efficient but more approximate are the linear interaction energy (LIE) [25] and the combination of molecular mechanics (MM) calculations and continuum solvation models such as Poisson–Boltzmann surface area (PBSA) and generalized Born surface area (GBSA) 26, 27. In addition to their lower computational demands, MM/PBSA and MM/GBSA allow the direct estimation of ΔGbind. The free energy of the ligand, the protein and the protein–ligand complex are each estimated as a sum of four terms (Eq. (4)):G=EMM+Gsolv+GnonpolarTSMMwhere Gsolv is the polar solvation free energy (estimated by the PB or GB equation), Gnon-polar the nonpolar solvation free energy (estimated by the solvent accessible surface area), EMM the enthalpy and SMM the entropy of the system (estimated by molecular mechanics). In this equation, the entropy term, SMM, is the major limitation of this approach because it is approximated by normal mode analysis of the harmonic frequencies obtained from the MM calculation [28].

Because most MD methods assume a system in equilibrium, insufficient equilibration before the production simulation can have severe effects on the accuracy of the obtained results. This is especially important in the context of drug–receptor complexes, which tend to be large and can involve long-time motions. Similarly, sufficient sampling during the production simulation is crucial for high-quality results, leading to the development of a large variety of sampling enhancement techniques to overcome barriers and increase the sampling efficiency [29].

The different approaches discussed above can be used in computer-aided medicinal chemistry to understand drug–receptor interactions better and support the design of more-active or more-specific ligands. To gain an overall perspective on the subject, literature in the field of medicinal chemistry was surveyed for the past four years and articles using MD were collected. This analysis enables understanding of the variety of fields in which MD is used and highlights the key information this technique gives access to. In the following section, representative examples of MD are reported and discussed for studies of ligand–target complexes.

Section snippets

Ligand–macromolecule complexes

Generally, MD studies reported in the literature focus on drug–target complexes, where in silico investigations are undertaken to elucidate the binding mode of a particular ligand in a given biological target. But MD can offer many more solutions to study or predict the binding of a small molecule to a macromolecule. Different interesting approaches are presented in the following chapter, in which the literature is divided into three main categories: (i) human cytoplasmic proteins; (ii) human

Future challenges of MD

As shown in the previous sections, MD can be applied at different stages of a drug discovery project. Nevertheless, limitations of this technique can be encountered when it comes to small molecules binding to a biological target. In the following section, we discuss how scientists deal with limitations such as the size of the studied systems, the time scale of the simulations and the inaccuracy of the force-fields.

A major limitation of MD is the simulation time scale (generally from 10−9 s to 10

Concluding remarks

In this broad overview of applications for MD in medicinal chemistry, we showed that a therapeutically relevant target and its molecular binders can be investigated from different perspectives. From the point of view of the target, MD allows the investigation of its conformational space and its catalytic mechanism, which is key information to understand drug–target binding. The prediction of allosteric sites also generates valuable information for the design of novel inhibitors for target

Acknowledgment

Christin Rakers is grateful to the Ernst Schering Foundation for the award of a doctoral scholarship.

Jérémie Mortier is a postdoctoral fellow in Gerhard Wolber's computer-aided drug design group at the Free University of Berlin, Germany. His main field of research is at the interface of biological and medicinal chemistry, with a particular focus on the prediction and understanding of molecular systems, their structures and interactions. After a Master in Chemistry in 2006, he was first introduced to computational chemistry during his PhD in pharmaceutical and biomedical sciences at the

References (117)

  • A.A. Amato

    GQ-16, a novel peroxisome proliferator-activated receptor gamma (PPAR gamma) ligand, promotes insulin sensitization without weight gain

    J. Biol. Chem.

    (2012)
  • H. Lei

    Early stage intercalation of doxorubicin to DNA fragments observed in molecular dynamics binding simulations

    J. Mol. Graph. Model.

    (2012)
  • J. Li

    Identification of nonplanar small molecule for G-quadruplex grooves: molecular docking and molecular dynamic study

    Bioorg. Med. Chem. Lett.

    (2011)
  • M. Huang

    Binding modes of diketo-acid inhibitors of HIV-1 integrase: a comparative molecular dynamics simulation study

    J. Mol. Graph. Model.

    (2011)
  • C. Rakers

    Inhibitory potency of flavonoid derivatives on influenza virus neuraminidase

    Bioorg. Med. Chem. Lett.

    (2014)
  • A. Perdih et al.

    MurD ligase from Escherichia coli: C-terminal domain closing motion

    Comput. Theor. Chem.

    (2012)
  • J.C. Phillips

    Scalable molecular dynamics with NAMD

    J. Comput. Chem.

    (2005)
  • B.R. Brooks

    CHARMM: the biomolecular simulation program

    J. Comput. Chem.

    (2009)
  • D.A. Case

    The Amber biomolecular simulation programs

    J. Comput. Chem.

    (2005)
  • S. Riniker

    On developing coarse-grained models for biomolecular simulation: a review

    Phys. Chem. Chem. Phys.

    (2012)
  • J. Gao et al.

    Preface to proceedings of the symposium on methods and applications of combined quantum mechanical and molecular mechanical potentials (2001)

    Theor. Chem. Acc.

    (2003)
  • D. Bakowies et al.

    Hybrid models for combined quantum mechanical and molecular mechanical approaches

    J. Phys. Chem.

    (1996)
  • M.J. Field

    A combined quantum-mechanical and molecular mechanical potential for molecular-dynamics simulations

    J. Comput. Chem.

    (1990)
  • E. Brini

    Systematic coarse-graining methods for soft matter simulations – a review

    Soft Matter

    (2013)
  • S. Riniker

    Structural effects of an atomic-level layer of water molecules around proteins solvated in supra-molecular coarse-grained water

    J. Phys. Chem. B

    (2012)
  • W.D. Cornell

    A second generation force field for the simulation of proteins, nucleic acids, and organic molecules (vol. 117, pg. 5179, 1995)

    J. Am. Chem. Soc.

    (1996)
  • W.L. Jorgensen et al.

    The Opls potential functions for proteins–energy minimizations for crystals of cyclic-peptides and crambin

    J. Am. Chem. Soc.

    (1988)
  • A.D. MacKerell

    All-atom empirical potential for molecular modeling and dynamics studies of proteins

    J. Phys. Chem. B

    (1998)
  • C. Oostenbrink

    A biomolecular force field based on the free enthalpy of hydration and solvation: the GROMOS force-field parameter sets 53A5 and 53A6

    J. Comput. Chem.

    (2004)
  • J.M. Wang

    Development and testing of a general amber force field (vol. 25, pg. 1157, 2004)

    J. Comput. Chem.

    (2005)
  • W.L. Jorgensen

    Development and testing of the OPLS all-atom force field on conformational energetics and properties of organic liquids

    J. Am. Chem. Soc.

    (1996)
  • K. Vanommeslaeghe et al.

    Automation of the CHARMM general force field (CGenFF) I: bond perception and atom typing

    J. Chem. Inf. Model.

    (2012)
  • K. Vanommeslaeghe

    Automation of the CHARMM general force field (CGenFF) II: assignment of bonded parameters and partial atomic charges

    J. Chem. Inf. Model.

    (2012)
  • A.K. Malde

    An automated force field topology builder (ATB) and repository: version 1.0

    J. Chem. Theory Comput.

    (2011)
  • J.G. Kirkwood

    Statistical mechanics of fluid mixtures

    J. Chem. Phys.

    (1935)
  • R.W. Zwanzig

    Statistical mechanical theory of transport processes. 7. The coefficient of thermal conductivity of monatomic liquids

    J. Chem. Phys.

    (1954)
  • T. Hansson

    Ligand binding affinity prediction by linear interaction energy methods

    J. Comput. Aided Mol. Des.

    (1998)
  • J.M. Wang

    Recent advances in free energy calculations with a combination of molecular mechanics and continuum models

    Curr. Comput. Aided Drug Des.

    (2006)
  • P.A. Kollman

    Calculating structures and free energies of complex molecules: combining molecular mechanics and continuum models

    Acc. Chem. Res.

    (2000)
  • J. Kongsted et al.

    An improved method to predict the entropy term with the MM/PBSA approach

    J. Comput. Aided Mol. Des.

    (2009)
  • E. Muzzioli

    Assessing protein kinase selectivity with molecular dynamics and MM-PBSA binding free energy calculations

    Chem. Biol. Drug Des.

    (2011)
  • S.K. Tripathi

    Exploring the selectivity of a ligand complex with CDK2/CDK1: a molecular dynamics simulation approach

    J. Mol. Recognit.

    (2012)
  • J.S. Patel

    Steered molecular dynamics simulations for studying protein–ligand interaction in cyclin-dependent kinase 5

    J. Chem. Inf. Model.

    (2014)
  • N. Songtawee

    Computational study of EGFR inhibition: molecular dynamics studies on the active and inactive protein conformations

    J. Mol. Model.

    (2013)
  • Y. Li

    Conformational transition pathways of epidermal growth factor receptor kinase domain from multiple molecular dynamics simulations and Bayesian clustering

    J. Chem. Theory Comput.

    (2014)
  • D. Shukla

    Activation pathway of Src kinase reveals intermediate states as targets for drug design

    Nat. Commun.

    (2014)
  • P.M. Gasper

    Allosteric networks in thrombin distinguish procoagulant vs. anticoagulant activities

    Proc. Natl. Acad. Sci. U. S. A.

    (2012)
  • B. Fuglestad

    Correlated motions and residual frustration in thrombin

    J. Phys. Chem. B

    (2013)
  • C.B. Jalkute

    Molecular dynamics simulation and molecular docking studies of angiotensin converting enzyme with inhibitor lisinopril and amyloid beta peptide

    Protein J.

    (2013)
  • K. Natarajan et al.

    Understanding the basis of drug resistance of the mutants of alphabeta-tubulin dimer via molecular dynamics simulations

    PLoS ONE

    (2012)
  • Cited by (167)

    • Computational approaches in drug discovery and design

      2023, Computational Approaches in Drug Discovery, Development and Systems Pharmacology
    View all citing articles on Scopus

    Jérémie Mortier is a postdoctoral fellow in Gerhard Wolber's computer-aided drug design group at the Free University of Berlin, Germany. His main field of research is at the interface of biological and medicinal chemistry, with a particular focus on the prediction and understanding of molecular systems, their structures and interactions. After a Master in Chemistry in 2006, he was first introduced to computational chemistry during his PhD in pharmaceutical and biomedical sciences at the University of Namur, Belgium, in 2010. His position is currently funded by a fellowship from the Deutsche Forschung Gemeinschaft.

    Sereina Riniker received her PhD at ETH Zurich in the field of molecular dynamics simulations. In 2012, she moved on to take a postdoctoral position in cheminformatics at the Novartis Institutes for BioMedical Research in Basel as well as Cambridge, Massachusetts, before returning to ETH as Assistant Professor for Computational Chemistry in 2014. Her current research is focused on molecular dynamics and cheminformatics as well as synergies between the two disciplines.

    Gerhard Wolber is Professor for Pharmaceutical Chemistry at the Institute of Pharmacy at the Freie Universitaet, Berlin. After his studies of pharmacy at the University of Innsbruck and computer science at the Technical University of Vienna, he received his PhD in medicinal chemistry at the University of Innsbruck. In 2003 he founded the company Inte:Ligand, which successfully develops and markets computational drug development software. In 2010 he moved back to academia by accepting an appointment as professor at the Freie Universitaet Berlin. His research focuses on computational drug design and the development of new virtual screening tools and algorithms.

    View full text