Force field (chemistry)

Last updated
Part of force field of ethane for the C-C stretching bond. Bond stretching energy.png
Part of force field of ethane for the C-C stretching bond.

In the context of chemistry, molecular physics and physical chemistry and molecular modelling, a force field is a computational model that is used to describe the forces between atoms (or collections of atoms) within molecules or between molecules as well as in crystals. Force fields are a variety of interatomic potentials. More precisely, the force field refers to the functional form and parameter sets used to calculate the potential energy of a system of the atomistic level. Force fields are usually used in molecular dynamics or Monte Carlo simulations. The parameters for a chosen energy function may be derived from classical laboratory experiment data, calculations in quantum mechanics, or both. Force fields utilize the same concept as force fields in classical physics, with the main difference that the force field parameters in chemistry describe the energy landscape on the atomistic level. From a force field, the acting forces on every particle are derived as a gradient of the potential energy with respect to the particle coordinates. [1]

Contents

A large number of different force field types exist today, e.g. for organic molecules, ions, polymers, minerals and metals. Depending on the material, different functional forms are usually chosen for the force fields since different types of atomistic interactions dominate the material behavior.

There are various criteria that can be used for categorizing force field parametrization strategies. An important differentiation is 'component-specific' and 'transferable'. For a component-specific parametrization, the considered force field is developed solely for describing a single given substance, e.g. water. [2] For a transferable force field, all or some parameters are designed as building blocks and become transferable/ applicable for different substances, e.g. methyl groups in alkane transferable force fields. [3] A different important differentiation addresses the physical structure of the models: All-atom force fields provide parameters for every type of atom in a system, including hydrogen, while united-atom interatomic potentials treat the hydrogen and carbon atoms in methyl groups and methylene bridges as one interaction center. [4] [5] Coarse-grained potentials, which are often used in long-time simulations of macromolecules such as proteins, nucleic acids, and multi-component complexes, sacrifice chemical details for higher computing efficiency. [6]

Force fields for molecular systems

Molecular mechanics potential energy function with continuum solvent. MM PEF.png
Molecular mechanics potential energy function with continuum solvent.

The basic functional form of potential energy for modeling molecular systems includes intramolecular interaction terms for interactions of atoms that are linked by covalent bonds and intermolecular (i.e. nonbonded also termed noncovalent) terms that describe the long-range electrostatic and van der Waals forces. The specific decomposition of the terms depends on the force field, but a general form for the total energy in an additive force field can be written as

where the components of the covalent and noncovalent contributions are given by the following summations:

The bond and angle terms are usually modeled by quadratic energy functions that do not allow bond breaking. A more realistic description of a covalent bond at higher stretching is provided by the more expensive Morse potential. The functional form for dihedral energy is variable from one force field to another. Additional, "improper torsional" terms may be added to enforce the planarity of aromatic rings and other conjugated systems, and "cross-terms" that describe the coupling of different internal variables, such as angles and bond lengths. Some force fields also include explicit terms for hydrogen bonds.

The nonbonded terms are computationally most intensive. A popular choice is to limit interactions to pairwise energies. The van der Waals term is usually computed with a Lennard-Jones potential [7] or the Mie potential [8] and the electrostatic term with Coulomb's law. However, both can be buffered or scaled by a constant factor to account for electronic polarizability. A large number of force fields based on this or similar energy expressions have been proposed in the past decades for modeling different types of materials such as molecular substances, metals, glasses etc. - see below for a comprehensive list of force fields.

Bond stretching

As it is rare for bonds to deviate significantly from their equilibrium values, the most simplistic approaches utilize a Hooke's law formula:

where is the force constant, is the bond length and is the value for the bond length between atoms and when all other terms in the force field are set to 0. The term is at times differently defined/ taken at different thermodynamic conditions.

The bond stretching constant can be determined from the experimental Infrared spectrum, Raman spectrum, or high-level quantum mechanical calculations. [4] The constant determines vibrational frequencies in molecular dynamics simulations. The stronger the bond is between atoms, the higher is the value of the force constant, and the higher the wavenumber (energy) in the IR/Raman spectrum.

Though the formula of Hooke's law provides a reasonable level of accuracy at bond lengths near the equilibrium distance, it is less accurate as one moves away. In order to model the Morse curve better one could employ cubic and higher powers. [4] [9] However, for most practical applications these differences are negligible and inaccuracies in predictions of bond lengths are on the order of the thousandth of an angstrom, which is also the limit of reliability for common force fields. A Morse potential can be employed instead to enable bond breaking and higher accuracy, even though it is less efficient to compute. For reactive force fields, bond breaking and bond orders are additionally considered.

Electrostatic interactions

Electrostatic interactions are represented by a Coulomb energy, which utilizes atomic charges to represent chemical bonding ranging from covalent to polar covalent and ionic bonding. The typical formula is the Coulomb law:

Where is the distance between two atoms and . The total Coulomb energy is a summation over all pairwise combinations of atoms and usually excludes 1, 2 bonded atoms, 1, 3 bonded atoms, as well as 1, 4 bonded atoms. [10] [11] [12]

Atomic charges can make dominant contributions to the potential energy, especially for polar molecules and ionic compounds, and are critical to simulate the geometry, interaction energy, as well as the reactivity. The assignment of charges usually uses some heuristic approach, with different possible solutions.

Force fields for crystal systems

Atomistic interactions in crystal systems significantly deviate from those in molecular systems, [13] e.g. of organic molecules. For crystal systems, in particular multi-body interactions are important and cannot be neglected if a high accuracy of the force field is the aim. For crystal systems with covalent bonding, bond order potentials are usually used, e.g. Tersoff potentials. [14] For metal systems, usually embedded atom potentials [15] [16] are used. For metals, also so-called Drude model potentials have been developed, [17] which describe a form of attachment of electrons to nuclei. [18] [19]

Parameterization

In addition to the functional form of the potentials, a force fields consists of the parameters of these functions. Together, they specify the interactions on the atomistic level. The parametrization, i.e. determining of the parameter values, is crucial for the accuracy and reliability of the force field. Different parametrization procedures have been developed for the parametrization of different substances, e.g. metals, ions, and molecules. For different material types, usually different parametrization strategies are used. In general, two main types can be distinguished for the parametrization, either using data/ information from the atomistic level, e.g. from quantum mechanical calculations or spectroscopic data, or using data from macroscopic properties, e.g. the hardness or compressibility of a given material. Often a combination of these routes is used. Hence, one way or the other, the force field parameters are always determined in an empirical way. Nevertheless, the term 'empirical' is often used in the context of force field parameters when macroscopic material property data was used for the fitting. Experimental data (microscopic and macroscopic) included for the fit, for example, the enthalpy of vaporization, enthalpy of sublimation, dipole moments, and various spectroscopic properties such as vibrational frequencies. [20] [9] [21] Often, for molecular systems, quantum mechanical calculations in the gas phase are used for parametrizing intramolecular interactions and parametrizing intermolecular dispersive interactions by using macroscopic properties such as liquid densities. [3] [22] [23] The assignment of atomic charges often follows quantum mechanical protocols with some heuristics, which can lead to significant deviation in representing specific properties. [24] [25] [26]

A large number of workflows and parametrization procedures have been employed in the past decades using different data and optimization strategies for determining the force field parameters. They differ significantly, which is also due to different focuses of different developments. The parameters for molecular simulations of biological macromolecules such as proteins, DNA, and RNA were often derived/ transferred from observations for small organic molecules, which are more accessible for experimental studies and quantum calculations.

Atom types are defined for different elements as well as for the same elements in sufficiently different chemical environments. For example, oxygen atoms in water and an oxygen atoms in a carbonyl functional group are classified as different force field types. [21] Typical molecular force field parameter sets include values for atomic mass, atomic charge, Lennard-Jones parameters for every atom type, as well as equilibrium values of bond lengths, bond angles, and dihedral angles. [27] The bonded terms refer to pairs, triplets, and quadruplets of bonded atoms, and include values for the effective spring constant for each potential.

Heuristic force field parametrization procedures have been very successfully for many year, but recently criticized. [28] [29] since they are usually not fully automated and therefore subject to some subjectivity of the developers, which also brings problems regarding the reproducibility of the parametrization procedure.

Efforts to provide open source codes and methods include openMM and openMD. The use of semi-automation or full automation, without input from chemical knowledge, is likely to increase inconsistencies at the level of atomic charges, for the assignment of remaining parameters, and likely to dilute the interpretability and performance of parameters.

Force field databases

A large number of force fields has been published in the past decades - mostly in scientific publications. In recent years, some databases have attempted to collect, categorize and make force fields digitally available. Therein, different databases, focus on different types of force fields. For example, the openKim database focuses on interatomic functions describing the individual interactions between specific elements. [30] The TraPPE database focuses on transferable force fields of organic molecules (developed by the Siepmann group). [31] The MolMod database focuses on molecular and ionic force fields (both component-specific and transferable). [5] [32]

Transferability and mixing function types

Functional forms and parameter sets have been defined by the developers of interatomic potentials and feature variable degrees of self-consistency and transferability. When functional forms of the potential terms vary or are mixed, the parameters from one interatomic potential function can typically not be used together with another interatomic potential function. [33] In some cases, modifications can be made with minor effort, for example, between 9-6 Lennard-Jones potentials to 12-6 Lennard-Jones potentials. [12] Transfers from Buckingham potentials to harmonic potentials, or from Embedded Atom Models to harmonic potentials, on the contrary, would require many additional assumptions and may not be possible.

In many cases, force fields can be straight forwardly combined. Yet, often, additional specifications and assumptions are required.

Limitations

All interatomic potentials are based on approximations and experimental data, therefore often termed empirical. The performance varies from higher accuracy than density functional theory (DFT) calculations, with access to million times larger systems and time scales, to random guesses depending on the force field. [34] The use of accurate representations of chemical bonding, combined with reproducible experimental data and validation, can lead to lasting interatomic potentials of high quality with much fewer parameters and assumptions in comparison to DFT-level quantum methods. [35] [36]

Possible limitations include atomic charges, also called point charges. Most force fields rely on point charges to reproduce the electrostatic potential around molecules, which works less well for anisotropic charge distributions. [37] The remedy is that point charges have a clear interpretation [26] and virtual electrons can be added to capture essential features of the electronic structure, such additional polarizability in metallic systems to describe the image potential, internal multipole moments in π-conjugated systems, and lone pairs in water. [38] [39] [40] Electronic polarization of the environment may be better included by using polarizable force fields [41] [42] or using a macroscopic dielectric constant. However, application of one value of dielectric constant is a coarse approximation in the highly heterogeneous environments of proteins, biological membranes, minerals, or electrolytes. [43]

All types of van der Waals forces are also strongly environment-dependent because these forces originate from interactions of induced and "instantaneous" dipoles (see Intermolecular force). The original Fritz London theory of these forces applies only in a vacuum. A more general theory of van der Waals forces in condensed media was developed by A. D. McLachlan in 1963 and included the original London's approach as a special case. [44] The McLachlan theory predicts that van der Waals attractions in media are weaker than in vacuum and follow the like dissolves like rule, which means that different types of atoms interact more weakly than identical types of atoms. [45] This is in contrast to combinatorial rules or Slater-Kirkwood equation applied for development of the classical force fields. The combinatorial rules state that the interaction energy of two dissimilar atoms (e.g., C...N) is an average of the interaction energies of corresponding identical atom pairs (i.e., C...C and N...N). According to McLachlan's theory, the interactions of particles in media can even be fully repulsive, as observed for liquid helium, [44] however, the lack of vaporization and presence of a freezing point contradicts a theory of purely repulsive interactions. Measurements of attractive forces between different materials (Hamaker constant) have been explained by Jacob Israelachvili. [44] For example, "the interaction between hydrocarbons across water is about 10% of that across vacuum". [44] Such effects are represented in molecular dynamics through pairwise interactions that are spatially more dense in the condensed phase relative to the gas phase and reproduced once the parameters for all phases are validated to reproduce chemical bonding, density, and cohesive/surface energy.

Limitations have been strongly felt in protein structure refinement. The major underlying challenge is the huge conformation space of polymeric molecules, which grows beyond current computational feasibility when containing more than ~20 monomers. [46] Participants in Critical Assessment of protein Structure Prediction (CASP) did not try to refine their models to avoid "a central embarrassment of molecular mechanics, namely that energy minimization or molecular dynamics generally leads to a model that is less like the experimental structure". [47] Force fields have been applied successfully for protein structure refinement in different X-ray crystallography and NMR spectroscopy applications, especially using program XPLOR. [48] However, the refinement is driven mainly by a set of experimental constraints and the interatomic potentials serve mainly to remove interatomic hindrances. The results of calculations were practically the same with rigid sphere potentials implemented in program DYANA [49] (calculations from NMR data), or with programs for crystallographic refinement that use no energy functions at all. These shortcomings are related to interatomic potentials and to the inability to sample the conformation space of large molecules effectively. [50] Thereby also the development of parameters to tackle such large-scale problems requires new approaches. A specific problem area is homology modeling of proteins. [51] Meanwhile, alternative empirical scoring functions have been developed for ligand docking, [52] protein folding, [53] [54] [55] homology model refinement, [56] computational protein design, [57] [58] [59] and modeling of proteins in membranes. [60]

It was also argued that some protein force fields operate with energies that are irrelevant to protein folding or ligand binding. [41] The parameters of proteins force fields reproduce the enthalpy of sublimation, i.e., energy of evaporation of molecular crystals. However, protein folding and ligand binding are thermodynamically closer to crystallization, or liquid-solid transitions as these processes represent freezing of mobile molecules in condensed media. [61] [62] [63] Thus, free energy changes during protein folding or ligand binding are expected to represent a combination of an energy similar to heat of fusion (energy absorbed during melting of molecular crystals), a conformational entropy contribution, and solvation free energy. The heat of fusion is significantly smaller than enthalpy of sublimation. [44] Hence, the potentials describing protein folding or ligand binding need more consistent parameterization protocols, e.g., as described for IFF. Indeed, the energies of H-bonds in proteins are ~ -1.5 kcal/mol when estimated from protein engineering or alpha helix to coil transition data, [64] [65] but the same energies estimated from sublimation enthalpy of molecular crystals were -4 to -6 kcal/mol, [66] which is related to re-forming existing hydrogen bonds and not forming hydrogen bonds from scratch. The depths of modified Lennard-Jones potentials derived from protein engineering data were also smaller than in typical potential parameters and followed the like dissolves like rule, as predicted by McLachlan theory. [41]

Force fields available in literature

Different force fields are designed for different purposes:

Classical

Polarizable

Several force fields explicitly capture polarizability, where a particle's effective charge can be influenced by electrostatic interactions with its neighbors. Core-shell models are common, which consist of a positively charged core particle, representing the polarizable atom, and a negatively charged particle attached to the core atom through a spring-like harmonic oscillator potential. [86] [87] [88] Recent examples include polarizable models with virtual electrons that reproduce image charges in metals [76] and polarizable biomolecular force fields. [89]

Reactive

Coarse-grained

Machine learning

Water

The set of parameters used to model water or aqueous solutions (basically a force field for water) is called a water model. Many water models have been proposed; [5] some examples are TIP3P, TIP4P, [2] SPC, flexible simple point charge water model (flexible SPC), ST2, and mW. [128] Other solvents and methods of solvent representation are also applied within computational chemistry and physics; these are termed solvent models.

Modified amino acids

Other

See also

Related Research Articles

<span class="mw-page-title-main">Computational chemistry</span> Branch of chemistry

Computational chemistry is a branch of chemistry that uses computer simulations to assist in solving chemical problems. It uses methods of theoretical chemistry incorporated into computer programs to calculate the structures and properties of molecules, groups of molecules, and solids. The importance of this subject stems from the fact that, with the exception of some relatively recent findings related to the hydrogen molecular ion, achieving an accurate quantum mechanical depiction of chemical systems analytically, or in a closed form, is not feasible. The complexity inherent in the many-body problem exacerbates the challenge of providing detailed descriptions of quantum mechanical systems. While computational results normally complement information obtained by chemical experiments, it can occasionally predict unobserved chemical phenomena.

An intermolecular force (IMF) is the force that mediates interaction between molecules, including the electromagnetic forces of attraction or repulsion which act between atoms and other types of neighbouring particles, e.g. atoms or ions. Intermolecular forces are weak relative to intramolecular forces – the forces which hold a molecule together. For example, the covalent bond, involving sharing electron pairs between atoms, is much stronger than the forces present between neighboring molecules. Both sets of forces are essential parts of force fields frequently used in molecular mechanics.

<span class="mw-page-title-main">Van der Waals force</span> Interactions between groups of atoms that do not arise from chemical bonds

In molecular physics and chemistry, the van der Waals force is a distance-dependent interaction between atoms or molecules. Unlike ionic or covalent bonds, these attractions do not result from a chemical electronic bond; they are comparatively weak and therefore more susceptible to disturbance. The van der Waals force quickly vanishes at longer distances between interacting molecules.

<span class="mw-page-title-main">Molecular dynamics</span> Computer simulations to discover and understand chemical properties

Molecular dynamics (MD) is a computer simulation method for analyzing the physical movements of atoms and molecules. The atoms and molecules are allowed to interact for a fixed period of time, giving a view of the dynamic "evolution" of the system. In the most common version, the trajectories of atoms and molecules are determined by numerically solving Newton's equations of motion for a system of interacting particles, where forces between the particles and their potential energies are often calculated using interatomic potentials or molecular mechanical force fields. The method is applied mostly in chemical physics, materials science, and biophysics.

<span class="mw-page-title-main">Lennard-Jones potential</span> Model of intermolecular interactions

In computational chemistry, molecular physics, and physical chemistry the Lennard-Jones potential is an intermolecular pair potential. Out of all the intermolecular potentials, the Lennard-Jones potential is probably the one that has been the most extensively studied. It is considered an archetype model for simple yet realistic intermolecular interactions.

<span class="mw-page-title-main">AMBER</span>

Assisted Model Building with Energy Refinement (AMBER) is a family of force fields for molecular dynamics of biomolecules originally developed by Peter Kollman's group at the University of California, San Francisco.

Chemistry at Harvard Macromolecular Mechanics (CHARMM) is the name of a widely used set of force fields for molecular dynamics, and the name for the molecular dynamics simulation and analysis computer software package associated with them. The CHARMM Development Project involves a worldwide network of developers working with Martin Karplus and his group at Harvard to develop and maintain the CHARMM program. Licenses for this software are available, for a fee, to people and groups working in academia.

<span class="mw-page-title-main">Molecular mechanics</span> Use of classical mechanics to model molecular systems

Molecular mechanics uses classical mechanics to model molecular systems. The Born–Oppenheimer approximation is assumed valid and the potential energy of all systems is calculated as a function of the nuclear coordinates using force fields. Molecular mechanics can be used to study molecule systems ranging in size and complexity from small to large biological systems or material assemblies with many thousands to millions of atoms.

<span class="mw-page-title-main">Molecular modelling</span> Discovering chemical properties by physical simulations

Molecular modelling encompasses all methods, theoretical and computational, used to model or mimic the behaviour of molecules. The methods are used in the fields of computational chemistry, drug design, computational biology and materials science to study molecular systems ranging from small chemical systems to large biological molecules and material assemblies. The simplest calculations can be performed by hand, but inevitably computers are required to perform molecular modelling of any reasonably sized system. The common feature of molecular modelling methods is the atomistic level description of the molecular systems. This may include treating atoms as the smallest individual unit, or explicitly modelling protons and neutrons with its quarks, anti-quarks and gluons and electrons with its photons.

In quantum chemistry, the quantum theory of atoms in molecules (QTAIM), sometimes referred to as atoms in molecules (AIM), is a model of molecular and condensed matter electronic systems in which the principal objects of molecular structure - atoms and bonds - are natural expressions of a system's observable electron density distribution function. An electron density distribution of a molecule is a probability distribution that describes the average manner in which the electronic charge is distributed throughout real space in the attractive field exerted by the nuclei. According to QTAIM, molecular structure is revealed by the stationary points of the electron density together with the gradient paths of the electron density that originate and terminate at these points.

Tinker, previously stylized as TINKER, is a suite of computer software applications for molecular dynamics simulation. The codes provide a complete and general set of tools for molecular mechanics and molecular dynamics, with some special features for biomolecules. The core of the software is a modular set of callable routines which allow manipulating coordinates and evaluating potential energy and derivatives via straightforward means.

Implicit solvation is a method to represent solvent as a continuous medium instead of individual “explicit” solvent molecules, most often used in molecular dynamics simulations and in other applications of molecular mechanics. The method is often applied to estimate free energy of solute-solvent interactions in structural and chemical processes, such as folding or conformational transitions of proteins, DNA, RNA, and polysaccharides, association of biological macromolecules with ligands, or transport of drugs across biological membranes.

<span class="mw-page-title-main">Water model</span> Aspect of computational chemistry

In computational chemistry, a water model is used to simulate and thermodynamically calculate water clusters, liquid water, and aqueous solutions with explicit solvent. The models are determined from quantum mechanics, molecular mechanics, experimental results, and these combinations. To imitate a specific nature of molecules, many types of models have been developed. In general, these can be classified by the following three points; (i) the number of interaction points called site, (ii) whether the model is rigid or flexible, (iii) whether the model includes polarization effects.

Drude particles are model oscillators used to simulate the effects of electronic polarizability in the context of a classical molecular mechanics force field. They are inspired by the Drude model of mobile electrons and are used in the computational study of proteins, nucleic acids, and other biomolecules.

The hybrid QM/MM approach is a molecular simulation method that combines the strengths of ab initio QM calculations (accuracy) and MM (speed) approaches, thus allowing for the study of chemical processes in solution and in proteins. The QM/MM approach was introduced in the 1976 paper of Warshel and Levitt. They, along with Martin Karplus, won the 2013 Nobel Prize in Chemistry for "the development of multiscale models for complex chemical systems".

In chemistry, a halogen bond (XB) occurs when there is evidence of a net attractive interaction between an electrophilic region associated with a halogen atom in a molecular entity and a nucleophilic region in another, or the same, molecular entity. Like a hydrogen bond, the result is not a formal chemical bond, but rather a strong electrostatic attraction. Mathematically, the interaction can be decomposed in two terms: one describing an electrostatic, orbital-mixing charge-transfer and another describing electron-cloud dispersion. Halogen bonds find application in supramolecular chemistry; drug design and biochemistry; crystal engineering and liquid crystals; and organic catalysis.

<span class="mw-page-title-main">Interatomic potential</span> Functions for calculating potential energy

Interatomic potentials are mathematical functions to calculate the potential energy of a system of atoms with given positions in space. Interatomic potentials are widely used as the physical basis of molecular mechanics and molecular dynamics simulations in computational chemistry, computational physics and computational materials science to explain and predict materials properties. Examples of quantitative properties and qualitative phenomena that are explored with interatomic potentials include lattice parameters, surface energies, interfacial energies, adsorption, cohesion, thermal expansion, and elastic and plastic material behavior, as well as chemical reactions.

In computational chemistry, a solvent model is a computational method that accounts for the behavior of solvated condensed phases. Solvent models enable simulations and thermodynamic calculations applicable to reactions and processes which take place in solution. These include biological, chemical and environmental processes. Such calculations can lead to new predictions about the physical processes occurring by improved understanding.

In the context of chemistry and molecular modelling, the Interface force field (IFF) is a force field for classical molecular simulations of atoms, molecules, and assemblies up to the large nanometer scale, covering compounds from across the periodic table. It employs a consistent classical Hamiltonian energy function for metals, oxides, and organic compounds, linking biomolecular and materials simulation platforms into a single platform. The reliability is often higher than that of density functional theory calculations at more than a million times lower computational cost. IFF includes a physical-chemical interpretation for all parameters as well as a surface model database that covers different cleavage planes and surface chemistry of included compounds. The Interface Force Field is compatible with force fields for the simulation of primarily organic compounds and can be used with common molecular dynamics and Monte Carlo codes. Structures and energies of included chemical elements and compounds are rigorously validated and property predictions are up to a factor of 100 more accurate relative to earlier models.

Alexander D. MacKerell, Jr. is an American biophysicist who is the Grollman-Glick Professor of Pharmaceutical Sciences at the University of Maryland, Baltimore (UMB) and the Director of the Computer-Aided Drug Design (CADD) Center at UMB. He is also the Co-Founder and Chief Scientific Officer of the drug design tech company SilcsBio. In 2022, MacKerell was awarded the prestigious American Chemical Society Award for Computers in Chemical and Pharmaceutical Research.

References

  1. Frenkel D (2007). Understanding molecular simulation : from algorithms to applications. Academic Press. ISBN   978-0-12-267351-1. OCLC   254835355.
  2. 1 2 Abascal JL, Vega C (December 2005). "A general purpose model for the condensed phases of water: TIP4P/2005". The Journal of Chemical Physics. 123 (23): 234505. Bibcode:2005JChPh.123w4505A. doi:10.1063/1.2121687. PMID   16392929. S2CID   9757894.
  3. 1 2 Siu SW, Pluhackova K, Böckmann RA (April 2012). "Optimization of the OPLS-AA Force Field for Long Hydrocarbons". Journal of Chemical Theory and Computation. 8 (4): 1459–70. doi:10.1021/ct200908r. PMID   26596756.
  4. 1 2 3 Leach A (2001-01-30). Molecular Modelling: Principles and Applications (2nd ed.). Harlow: Prentice Hall. ISBN   9780582382107.
  5. 1 2 3 Stephan, Simon; Horsch, Martin T.; Vrabec, Jadran; Hasse, Hans (2019-07-03). "MolMod – an open access database of force fields for molecular simulations of fluids". Molecular Simulation. 45 (10): 806–814. arXiv: 1904.05206 . doi:10.1080/08927022.2019.1601191. ISSN   0892-7022. S2CID   119199372.
  6. 1 2 Marrink SJ, Risselada HJ, Yefimov S, Tieleman DP, de Vries AH (July 2007). "The MARTINI force field: coarse grained model for biomolecular simulations" (PDF). The Journal of Physical Chemistry B. 111 (27): 7812–24. doi:10.1021/jp071097f. PMID   17569554. S2CID   13874240.
  7. Lenhard, Johannes; Stephan, Simon; Hasse, Hans (February 2024). "A child of prediction. On the History, Ontology, and Computation of the Lennard-Jonesium". Studies in History and Philosophy of Science. 103: 105–113. doi:10.1016/j.shpsa.2023.11.007. ISSN   0039-3681. PMID   38128443. S2CID   266440296.
  8. Fischer, Johann; Wendland, Martin (October 2023). "On the history of key empirical intermolecular potentials". Fluid Phase Equilibria. 573: 113876. doi: 10.1016/j.fluid.2023.113876 . ISSN   0378-3812.
  9. 1 2 Sun H, Mumby SJ, Maple JR, Hagler AT (April 1994). "An ab Initio CFF93 All-Atom Force Field for Polycarbonates". Journal of the American Chemical Society. 116 (7): 2978–2987. doi:10.1021/ja00086a030. ISSN   0002-7863.
  10. Huang J, MacKerell AD (September 2013). "CHARMM36 all-atom additive protein force field: validation based on comparison to NMR data". Journal of Computational Chemistry. 34 (25): 2135–45. doi:10.1002/jcc.23354. PMC   3800559 . PMID   23832629.
  11. Wang J, Wolf RM, Caldwell JW, Kollman PA, Case DA (July 2004). "Development and testing of a general amber force field". Journal of Computational Chemistry. 25 (9): 1157–74. doi: 10.1002/jcc.20035 . PMID   15116359. S2CID   18734898.
  12. 1 2 Mishra RK, Fernández-Carrasco L, Flatt RJ, Heinz H (July 2014). "A force field for tricalcium aluminate to characterize surface properties, initial hydration, and organically modified interfaces in atomic resolution". Dalton Transactions. 43 (27): 10602–16. doi:10.1039/c4dt00438h. hdl: 2117/24209 . PMID   24828263.
  13. Ashcroft, N. W.; Mermin, N. D.; Smoluchowski, R. (1977-01-01). "Solid State Physics". Physics Today. 30 (1): 61–65. Bibcode:1977PhT....30R..61A. doi: 10.1063/1.3037370 . ISSN   0031-9228.
  14. Tersoff, J. (1988-04-15). "New empirical approach for the structure and energy of covalent systems". Physical Review B. 37 (12): 6991–7000. Bibcode:1988PhRvB..37.6991T. doi:10.1103/physrevb.37.6991. ISSN   0163-1829. PMID   9943969.
  15. Daw, Murray S.; Baskes, M. I. (1984-06-15). "Embedded-atom method: Derivation and application to impurities, surfaces, and other defects in metals". Physical Review B. 29 (12): 6443–6453. Bibcode:1984PhRvB..29.6443D. doi:10.1103/physrevb.29.6443. ISSN   0163-1829.
  16. Daw, Murray S.; Foiles, Stephen M.; Baskes, Michael I. (March 1993). "The embedded-atom method: a review of theory and applications". Materials Science Reports. 9 (7–8): 251–310. doi: 10.1016/0920-2307(93)90001-u . ISSN   0920-2307.
  17. Lemkul JA, Huang J, Roux B, MacKerell AD (May 2016). "An Empirical Polarizable Force Field Based on the Classical Drude Oscillator Model: Development History and Recent Applications". Chemical Reviews. 116 (9): 4983–5013. doi:10.1021/acs.chemrev.5b00505. PMC   4865892 . PMID   26815602.
  18. Lorentz HA (1905). "The Motion of Electrons in Metallic Bodies, I.". Proc. K. Ned. Akad. Wet. 7: 451. Bibcode:1904KNAB....7..438L.
  19. Drude P (1900). "Zur Elekronentheorie der Metalle. I. Teil". Annalen der Physik. 306 (3): 566–613. doi: 10.1002/andp.19003060312 .
  20. Jorgensen WL, Maxwell DS, Tirado-Rives J (January 1996). "Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids". Journal of the American Chemical Society. 118 (45): 11225–11236. doi:10.1021/ja9621760. ISSN   0002-7863.
  21. 1 2 3 Dauber-Osguthorpe P, Roberts VA, Osguthorpe DJ, Wolff J, Genest M, Hagler AT (1988). "Structure and energetics of ligand binding to proteins: Escherichia coli dihydrofolate reductase-trimethoprim, a drug-receptor system". Proteins. 4 (1): 31–47. doi:10.1002/prot.340040106. PMID   3054871. S2CID   2845395.
  22. Aduri R, Psciuk BT, Saro P, Taniga H, Schlegel HB, SantaLucia J (July 2007). "AMBER Force Field Parameters for the Naturally Occurring Modified Nucleosides in RNA". Journal of Chemical Theory and Computation. 3 (4): 1464–75. doi:10.1021/ct600329w. PMID   26633217.
  23. Kirschner KN, Lins RD, Maass A, Soares TA (November 2012). "A Glycam-Based Force Field for Simulations of Lipopolysaccharide Membranes: Parametrization and Validation". Journal of Chemical Theory and Computation. 8 (11): 4719–31. doi:10.1021/ct300534j. PMID   26605626.
  24. Gross KC, Seybold PG, Hadad CM (2002). "Comparison of different atomic charge schemes for predicting pKa variations in substituted anilines and phenols". International Journal of Quantum Chemistry. 90 (1): 445–458. doi:10.1002/qua.10108. ISSN   0020-7608.
  25. Wang B, Li SL, Truhlar DG (December 2014). "Modeling the Partial Atomic Charges in Inorganometallic Molecules and Solids and Charge Redistribution in Lithium-Ion Cathodes". Journal of Chemical Theory and Computation. 10 (12): 5640–50. doi: 10.1021/ct500790p . PMID   26583247.
  26. 1 2 Heinz H, Suter UW (November 2004). "Atomic Charges for Classical Simulations of Polar Systems". The Journal of Physical Chemistry B. 108 (47): 18341–18352. Bibcode:2004APS..MAR.Y8006H. doi:10.1021/jp048142t. ISSN   1520-6106.
  27. Dharmawardhana CC, Kanhaiya K, Lin TJ, Garley A, Knecht MR, Zhou J, Miao J, Heinz H (2017-06-19). "Reliable computational design of biological-inorganic materials to the large nanometer scale using Interface-FF". Molecular Simulation. 43 (13–16): 1394–1405. doi:10.1080/08927022.2017.1332414. ISSN   0892-7022. S2CID   36710284.
  28. Wang LP, Martinez TJ, Pande VS (June 2014). "Building Force Fields: An Automatic, Systematic, and Reproducible Approach". The Journal of Physical Chemistry Letters. 5 (11): 1885–91. doi:10.1021/jz500737m. PMC   9649520 . PMID   26273869.
  29. 1 2 McDonagh JL, Shkurti A, Bray DJ, Anderson RL, Pyzer-Knapp EO (October 2019). "Utilizing Machine Learning for Efficient Parameterization of Coarse Grained Molecular Force Fields". Journal of Chemical Information and Modeling. 59 (10): 4278–4288. doi:10.1021/acs.jcim.9b00646. PMID   31549507. S2CID   202745539.
  30. Tadmor, E. B.; Elliott, R. S.; Sethna, J. P.; Miller, R. E.; Becker, C. A. (July 2011). "The potential of atomistic simulations and the knowledgebase of interatomic models". JOM. 63 (7): 17. Bibcode:2011JOM....63g..17T. doi: 10.1007/s11837-011-0102-6 . ISSN   1047-4838.
  31. Eggimann, Becky L.; Sunnarborg, Amara J.; Stern, Hudson D.; Bliss, Andrew P.; Siepmann, J. Ilja (2014-01-02). "An online parameter and property database for the TraPPE force field". Molecular Simulation. 40 (1–3): 101–105. doi:10.1080/08927022.2013.842994. ISSN   0892-7022. S2CID   95716947.
  32. Schmitt, Sebastian; Kanagalingam, Gajanan; Fleckenstein, Florian; Froescher, Daniel; Hasse, Hans; Stephan, Simon (2023-11-27). "Extension of the MolMod Database to Transferable Force Fields". Journal of Chemical Information and Modeling. 63 (22): 7148–7158. doi:10.1021/acs.jcim.3c01484. ISSN   1549-9596. PMID   37947503. S2CID   265103133.
  33. Heinz H, Ramezani-Dakhel H (January 2016). "Simulations of inorganic-bioorganic interfaces to discover new materials: insights, comparisons to experiment, challenges, and opportunities". Chemical Society Reviews. 45 (2): 412–48. doi:10.1039/c5cs00890e. PMID   26750724.
  34. Emami FS, Puddu V, Berry RJ, Varshney V, Patwardhan SV, Perry CC, Heinz H (2014-04-22). "Force Field and a Surface Model Database for Silica to Simulate Interfacial Properties in Atomic Resolution" (PDF). Chemistry of Materials. 26 (8): 2647–2658. doi:10.1021/cm500365c. ISSN   0897-4756.
  35. Ruiz VG, Liu W, Tkatchenko A (2016-01-15). "Density-functional theory with screened van der Waals interactions applied to atomic and molecular adsorbates on close-packed and non-close-packed surfaces". Physical Review B. 93 (3): 035118. Bibcode:2016PhRvB..93c5118R. doi:10.1103/physrevb.93.035118. hdl: 11858/00-001M-0000-0029-3035-8 . ISSN   2469-9950.
  36. Ruiz VG, Liu W, Zojer E, Scheffler M, Tkatchenko A (April 2012). "Density-functional theory with screened van der Waals interactions for the modeling of hybrid inorganic-organic systems". Physical Review Letters. 108 (14): 146103. Bibcode:2012PhRvL.108n6103R. doi: 10.1103/physrevlett.108.146103 . hdl: 11858/00-001M-0000-000F-C6EA-3 . PMID   22540809.
  37. Kramer C, Spinn A, Liedl KR (October 2014). "Charge Anisotropy: Where Atomic Multipoles Matter Most". Journal of Chemical Theory and Computation. 10 (10): 4488–96. doi:10.1021/ct5005565. PMID   26588145.
  38. Mahoney MW, Jorgensen WL (2000-05-22). "A five-site model for liquid water and the reproduction of the density anomaly by rigid, nonpolarizable potential functions". The Journal of Chemical Physics. 112 (20): 8910–8922. Bibcode:2000JChPh.112.8910M. doi:10.1063/1.481505. ISSN   0021-9606.
  39. 1 2 Xu R, Chen CC, Wu L, Scott MC, Theis W, Ophus C, et al. (November 2015). "Three-dimensional coordinates of individual atoms in materials revealed by electrontomography". Nature Materials. 14 (11): 1099–103. arXiv: 1505.05938 . Bibcode:2015NatMa..14.1099X. doi:10.1038/nmat4426. PMID   26390325. S2CID   5455024.
  40. 1 2 Pramanik C, Gissinger JR, Kumar S, Heinz H (December 2017). "Carbon Nanotube Dispersion in Solvents and Polymer Solutions: Mechanisms, Assembly, and Preferences". ACS Nano. 11 (12): 12805–12816. doi:10.1021/acsnano.7b07684. PMID   29179536.
  41. 1 2 3 Lomize AL, Reibarkh MY, Pogozheva ID (August 2002). "Interatomic potentials and solvation parameters from protein engineering data for buried residues". Protein Science. 11 (8): 1984–2000. doi:10.1110/ps.0307002. PMC   2373680 . PMID   12142453.
  42. 1 2 Warshel A, Sharma PK, Kato M, Parson WW (November 2006). "Modeling electrostatic effects in proteins". Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 1764 (11): 1647–76. doi:10.1016/j.bbapap.2006.08.007. PMID   17049320.
  43. Schutz CN, Warshel A (September 2001). "What are the dielectric "constants" of proteins and how to validate electrostatic models?". Proteins. 44 (4): 400–17. doi:10.1002/prot.1106. PMID   11484218. S2CID   9912122.
  44. 1 2 3 4 5 Israelachvili JN (2011). "Intermolecular and Surface Forces". Elsevier. pp. iii. doi:10.1016/b978-0-12-391927-4.10024-6. ISBN   978-0-12-391927-4.{{cite book}}: Missing or empty |title= (help)
  45. Leckband D, Israelachvili J (May 2001). "Intermolecular forces in biology". Quarterly Reviews of Biophysics. 34 (2): 105–267. doi:10.1017/S0033583501003687. PMID   11771120. S2CID   8401242.
  46. Pramanik C, Jamil T, Gissinger JR, Guittet D, Arias-Monje PJ, Kumar S, Heinz H (2019-10-03). "Polyacrylonitrile Interactions with Carbon Nanotubes in Solution: Conformations and Binding as a Function of Solvent, Temperature, and Concentration". Advanced Functional Materials. 29 (50): 1905247. doi:10.1002/adfm.201905247. ISSN   1616-301X. S2CID   208700020.
  47. Koehl P, Levitt M (February 1999). "A brighter future for protein structure prediction". Nature Structural Biology. 6 (2): 108–11. doi:10.1038/5794. PMID   10048917. S2CID   3162636.
  48. Brunger AT, Adams PD (June 2002). "Molecular dynamics applied to X-ray structure refinement". Accounts of Chemical Research. 35 (6): 404–12. doi:10.1021/ar010034r. PMID   12069625.
  49. Güntert P (May 1998). "Structure calculation of biological macromolecules from NMR data". Quarterly Reviews of Biophysics. 31 (2): 145–237. doi:10.1017/S0033583598003436. PMID   9794034. S2CID   43575627.
  50. Ostermeir K, Zacharias M (January 2013). "163 Enhanced sampling of peptides and proteins with a new biasing replica exchange method". Journal of Biomolecular Structure and Dynamics. 31 (sup1): 106. doi: 10.1080/07391102.2013.786405 . ISSN   0739-1102. S2CID   98441607.
  51. Tramontano A, Morea V (2003). "Assessment of homology-based predictions in CASP5". Proteins. 53 (Suppl 6): 352–68. doi: 10.1002/prot.10543 . PMID   14579324. S2CID   33186021.
  52. Gohlke H, Klebe G (August 2002). "Approaches to the description and prediction of the binding affinity of small-molecule ligands to macromolecular receptors". Angewandte Chemie. 41 (15): 2644–76. doi:10.1002/1521-3773(20020802)41:15<2644::AID-ANIE2644>3.0.CO;2-O. PMID   12203463.
  53. Edgcomb SP, Murphy KP (February 2000). "Structural energetics of protein folding and binding". Current Opinion in Biotechnology. 11 (1): 62–6. doi:10.1016/s0958-1669(99)00055-5. PMID   10679345.
  54. Lazaridis T, Karplus M (April 2000). "Effective energy functions for protein structure prediction". Current Opinion in Structural Biology. 10 (2): 139–45. doi:10.1016/s0959-440x(00)00063-4. PMID   10753811.
  55. Javidpour L (2012). "Computer Simulations of Protein Folding". Computing in Science & Engineering. 14 (2): 97–103. Bibcode:2012CSE....14b..97J. doi:10.1109/MCSE.2012.21. S2CID   17613729.
  56. Krieger E, Joo K, Lee J, Lee J, Raman S, Thompson J, et al. (2009). "Improving physical realism, stereochemistry, and side-chain accuracy in homology modeling: Four approaches that performed well in CASP8". Proteins. 77 (Suppl 9): 114–22. doi:10.1002/prot.22570. PMC   2922016 . PMID   19768677.
  57. Gordon DB, Marshall SA, Mayo SL (August 1999). "Energy functions for protein design". Current Opinion in Structural Biology. 9 (4): 509–13. doi:10.1016/S0959-440X(99)80072-4. PMID   10449371.
  58. Mendes J, Guerois R, Serrano L (August 2002). "Energy estimation in protein design". Current Opinion in Structural Biology. 12 (4): 441–6. doi:10.1016/s0959-440x(02)00345-7. PMID   12163065.
  59. Rohl CA, Strauss CE, Misura KM, Baker D (2004). "Protein Structure Prediction Using Rosetta". Numerical Computer Methods, Part D. Methods in Enzymology. Vol. 383. pp. 66–93. doi:10.1016/S0076-6879(04)83004-0. ISBN   9780121827885. PMID   15063647.
  60. Lomize AL, Pogozheva ID, Lomize MA, Mosberg HI (June 2006). "Positioning of proteins in membranes: a computational approach". Protein Science. 15 (6): 1318–33. doi:10.1110/ps.062126106. PMC   2242528 . PMID   16731967.
  61. Murphy KP, Gill SJ (December 1991). "Solid model compounds and the thermodynamics of protein unfolding". Journal of Molecular Biology. 222 (3): 699–709. doi:10.1016/0022-2836(91)90506-2. PMID   1660931.
  62. Shakhnovich EI, Finkelstein AV (October 1989). "Theory of cooperative transitions in protein molecules. I. Why denaturation of globular protein is a first-order phase transition". Biopolymers. 28 (10): 1667–80. doi:10.1002/bip.360281003. PMID   2597723. S2CID   26981215.
  63. Graziano G, Catanzano F, Del Vecchio P, Giancola C, Barone G (1996). "Thermodynamic stability of globular proteins: a reliable model from small molecule studies". Gazetta Chim. Italiana. 126: 559–567.
  64. Myers JK, Pace CN (October 1996). "Hydrogen bonding stabilizes globular proteins". Biophysical Journal. 71 (4): 2033–9. Bibcode:1996BpJ....71.2033M. doi:10.1016/S0006-3495(96)79401-8. PMC   1233669 . PMID   8889177.
  65. Scholtz JM, Marqusee S, Baldwin RL, York EJ, Stewart JM, Santoro M, Bolen DW (April 1991). "Calorimetric determination of the enthalpy change for the alpha-helix to coil transition of an alanine peptide in water". Proceedings of the National Academy of Sciences of the United States of America. 88 (7): 2854–8. Bibcode:1991PNAS...88.2854S. doi: 10.1073/pnas.88.7.2854 . PMC   51338 . PMID   2011594.
  66. Gavezzotti A, Filippini G (May 1994). "Geometry of the intermolecular XH. cntdot.. cntdot.. cntdot. Y (X, Y= N, O) hydrogen bond and the calibration of empirical hydrogen-bond potentials". The Journal of Physical Chemistry. 98 (18): 4831–7. doi:10.1021/j100069a010.
  67. Möllhoff M, Sternberg U (May 2001). "Molecular mechanics with fluctuating atomic charges–a new force field with a semi-empirical charge calculation". Molecular Modeling Annual. 7 (4): 90–102. doi:10.1007/s008940100008. S2CID   91705326.
  68. "ECEPP". biohpc.cornell.edu.
  69. Momany FA, McGuire RF, Burgess AW, Scheraga HA (October 1975). "Energy parameters in polypeptides. VII. Geometric parameters, partial atomic charges, nonbonded interactions, hydrogen bond interactions, and intrinsic torsional potentials for the naturally occurring amino acids". The Journal of Physical Chemistry. 79 (22): 2361–81. doi:10.1021/j100589a006.
  70. Arnautova YA, Jagielska A, Scheraga HA (March 2006). "A new force field (ECEPP-05) for peptides, proteins, and organic molecules". The Journal of Physical Chemistry B. 110 (10): 5025–44. doi:10.1021/jp054994x. PMID   16526746.
  71. Schaumann T, Braun W, Wüthrich K (March 1990). "The program FANTOM for energy refinement of polypeptides and proteins using a Newton–Raphson minimizer in torsion angle space". Biopolymers. 29 (4–5): 679–94. doi:10.1002/bip.360290403. S2CID   94519023.
  72. "GROMOS". www.igc.ethz.ch.
  73. 1 2 Heinz H, Lin TJ, Mishra RK, Emami FS (February 2013). "Thermodynamically consistent force fields for the assembly of inorganic, organic, and biological nanostructures: the INTERFACE force field". Langmuir. 29 (6): 1754–65. doi:10.1021/la3038846. PMID   23276161.
  74. "Interface Force Field (IFF)". Heinz Laboratory. 2 February 2016.
  75. 1 2 Mishra RK, Mohamed AK, Geissbühler D, Manzano H, Jamil T, Shahsavari R, Kalinichev AG, Galmarini S, Tao L, Heinz H, Pellenq R (December 2017). "A force field database for cementitious materials including validations, applications and opportunities" (PDF). Cement and Concrete Research. 102: 68–89. doi:10.1016/j.cemconres.2017.09.003.
  76. 1 2 Geada IL, Ramezani-Dakhel H, Jamil T, Sulpizi M, Heinz H (February 2018). "Insight into induced charges at metal surfaces and biointerfaces using a polarizable Lennard-Jones potential". Nature Communications. 9 (1): 716. Bibcode:2018NatCo...9..716G. doi:10.1038/s41467-018-03137-8. PMC   5818522 . PMID   29459638.
  77. Allinger NL (December 1977). "Conformational analysis. 130. MM2. A hydrocarbon force field utilizing V1 and V2 torsional terms". Journal of the American Chemical Society. 99 (25): 8127–34. doi:10.1021/ja00467a001.
  78. "MM2 and MM3 home page". Archived from the original on 2009-01-23.
  79. Allinger NL, Yuh YH, Lii JH (November 1989). "Molecular mechanics. The MM3 force field for hydrocarbons. 1". Journal of the American Chemical Society. 111 (23): 8551–66. doi:10.1021/ja00205a001.
  80. Lii JH, Allinger NL (November 1989). "Molecular mechanics. The MM3 force field for hydrocarbons. 2. Vibrational frequencies and thermodynamics". Journal of the American Chemical Society. 111 (23): 8566–75. doi:10.1021/ja00205a002.
  81. Lii JH, Allinger NL (November 1989). "Molecular mechanics. The MM3 force field for hydrocarbons. 3. The van der Waals' potentials and crystal data for aliphatic and aromatic hydrocarbons". Journal of the American Chemical Society. 111 (23): 8576–82. doi:10.1021/ja00205a003.
  82. Warshel A (1973). "Quantum mechanical consistent force field (QCFF/PI) method: Calculations of energies, conformations and vibronic interactions of ground and excited states of conjugated molecules". Israel Journal of Chemistry. 11 (5): 709–17. doi:10.1002/ijch.197300067.
  83. Warshel A, Levitt M (1974). QCFF/PI: A Program for the Consistent Force Field Evaluation of Equilibrium Geometries and Vibrational Frequencies of Molecules (Report). Indiana University: Quantum Chemistry Program Exchange. p. QCPE 247.
  84. Rappé AK, Casewit CJ, Colwell KS, Goddard III WA, Skiff WM (December 1992). "UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations". Journal of the American Chemical Society. 114 (25): 10024–10035. doi:10.1021/ja00051a040. ISSN   0002-7863.
  85. Heinz H, Koerner H, Anderson KL, Vaia RA, Farmer BL (November 2005). "Force Field for Mica-Type Silicates and Dynamics of Octadecylammonium Chains Grafted to Montmorillonite". Chemistry of Materials. 17 (23): 5658–5669. doi:10.1021/cm0509328. ISSN   0897-4756.
  86. Dick BG, Overhauser AW (1958-10-01). "Theory of the Dielectric Constants of Alkali Halide Crystals". Physical Review. 112 (1): 90–103. Bibcode:1958PhRv..112...90D. doi:10.1103/physrev.112.90. ISSN   0031-899X.
  87. Mitchell PJ, Fincham D (1993-02-22). "Shell model simulations by adiabatic dynamics". Journal of Physics: Condensed Matter. 5 (8): 1031–1038. Bibcode:1993JPCM....5.1031M. doi:10.1088/0953-8984/5/8/006. ISSN   0953-8984. S2CID   250752417.
  88. Yu H, van Gunsteren WF (November 2005). "Accounting for polarization in molecular simulation". Computer Physics Communications. 172 (2): 69–85. Bibcode:2005CoPhC.172...69Y. doi:10.1016/j.cpc.2005.01.022. ISSN   0010-4655.
  89. 1 2 Patel S, Brooks CL (January 2004). "CHARMM fluctuating charge force field for proteins: I parameterization and application to bulk organic liquid simulations". Journal of Computational Chemistry. 25 (1): 1–15. doi:10.1002/jcc.10355. PMID   14634989. S2CID   39320318.
  90. Yang L, Tan CH, Hsieh MJ, Wang J, Duan Y, Cieplak P, Caldwell J, Kollman PA, Luo R (July 2006). "New-generation amber united-atom force field". The Journal of Physical Chemistry B. 110 (26): 13166–76. doi:10.1021/jp060163v. PMID   16805629.
  91. "Tinker Molecular Modeling Package". dasher.wustl.edu.
  92. Liu C, Piquemal JP, Ren P (January 2020). "Implementation of Geometry-Dependent Charge Flux into the Polarizable AMOEBA+ Potential". The Journal of Physical Chemistry Letters. 11 (2): 419–426. doi:10.1021/acs.jpclett.9b03489. PMC   7384396 . PMID   31865706.
  93. Liu C, Piquemal JP, Ren P (July 2019). "AMOEBA+ Classical Potential for Modeling Molecular Interactions". Journal of Chemical Theory and Computation. 15 (7): 4122–4139. doi:10.1021/acs.jctc.9b00261. PMC   6615954 . PMID   31136175.
  94. Patel S, Mackerell AD, Brooks CL (September 2004). "CHARMM fluctuating charge force field for proteins: II protein/solvent properties from molecular dynamics simulations using a nonadditive electrostatic model". Journal of Computational Chemistry. 25 (12): 1504–14. doi: 10.1002/jcc.20077 . PMID   15224394. S2CID   16741310.
  95. Anisimov VM, Lamoureux G, Vorobyov IV, Huang N, Roux B, MacKerell AD (January 2005). "Determination of Electrostatic Parameters for a Polarizable Force Field Based on the Classical Drude Oscillator". Journal of Chemical Theory and Computation. 1 (1): 153–68. doi:10.1021/ct049930p. PMID   26641126.
  96. Yu H, Whitfield TW, Harder E, Lamoureux G, Vorobyov I, Anisimov VM, Mackerell AD, Roux B (2010). "Simulating Monovalent and Divalent Ions in Aqueous Solution Using a Drude Polarizable Force Field". Journal of Chemical Theory and Computation. 6 (3): 774–786. doi:10.1021/ct900576a. PMC   2838399 . PMID   20300554.
  97. Warshel A, Levitt M (May 1976). "Theoretical studies of enzymic reactions: dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme". Journal of Molecular Biology. 103 (2): 227–49. doi:10.1016/0022-2836(76)90311-9. PMID   985660.
  98. Sternberg U, Koch FT, Möllhoff M (May 1994). "New approach to the semiempirical calculation of atomic charges for polypeptides and large molecular systems". Journal of Computational Chemistry. 15 (5): 524–31. doi:10.1002/jcc.540150505. S2CID   5227353.
  99. Swart M, van Duijnen PT (May 2006). "DRF90: a polarizable force field". Molecular Simulation. 32 (6): 471–84. doi:10.1080/08927020600631270. S2CID   96616243.
  100. Engkvist O, Astrand PO, Karlström G (November 2000). "Accurate Intermolecular Potentials Obtained from Molecular Wave Functions: Bridging the Gap between Quantum Chemistry and Molecular Simulations". Chemical Reviews. 100 (11): 4087–108. doi:10.1021/cr9900477. PMID   11749341.
  101. Gao J, Habibollazadeh D, Shao L (November 1995). "A polarizable intermolecular potential function for simulation of liquid alcohols". The Journal of Physical Chemistry. 99 (44): 16460–7. doi:10.1021/j100044a039.
  102. Xie W, Pu J, Mackerell AD, Gao J (2007). "Development of a polarizable intermolecular potential function (PIPF) for liquid amides and alkanes". Journal of Chemical Theory and Computation. 3 (6): 1878–1889. doi:10.1021/ct700146x. PMC   2572772 . PMID   18958290.
  103. Maple JR, Cao Y, Damm W, Halgren TA, Kaminski GA, Zhang LY, Friesner RA (July 2005). "A Polarizable Force Field and Continuum Solvation Methodology for Modeling of Protein-Ligand Interactions". Journal of Chemical Theory and Computation. 1 (4): 694–715. doi:10.1021/ct049855i. PMID   26641692.
  104. Chelli R, Procacci P (November 2002). "A transferable polarizable electrostatic force field for molecular mechanics based on the chemical potential equalization principle". The Journal of Chemical Physics. 117 (20): 9175–89. Bibcode:2002JChPh.117.9175C. doi:10.1063/1.1515773.
  105. Cioce CR, McLaughlin K, Belof JL, Space B (December 2013). "A Polarizable and Transferable PHAST N2 Potential for Use in Materials Simulation". Journal of Chemical Theory and Computation. 9 (12): 5550–7. doi:10.1021/ct400526a. PMID   26592288.
  106. "Anthony Stone: Computer programs". www-stone.ch.cam.ac.uk.
  107. Gresh N, Cisneros GA, Darden TA, Piquemal JP (November 2007). "Anisotropic, Polarizable Molecular Mechanics Studies of Inter- and Intramolecular Interactions and Ligand-Macromolecule Complexes. A Bottom-Up Strategy". Journal of Chemical Theory and Computation. 3 (6): 1960–1986. doi:10.1021/ct700134r. PMC   2367138 . PMID   18978934.
  108. Piquemal JP, Cisneros GA, Reinhardt P, Gresh N, Darden TA (March 2006). "Towards a force field based on density fitting". The Journal of Chemical Physics. 124 (10): 104101. Bibcode:2006JChPh.124j4101P. doi:10.1063/1.2173256. PMC   2080832 . PMID   16542062.
  109. Cisneros GA, Piquemal JP, Darden TA (November 2006). "Generalization of the Gaussian electrostatic model: extension to arbitrary angular momentum, distributed multipoles, and speedup with reciprocal space methods". The Journal of Chemical Physics. 125 (18): 184101. Bibcode:2006JChPh.125r4101C. doi:10.1063/1.2363374. PMC   2080839 . PMID   17115732.
  110. Borodin O (August 2009). "Polarizable force field development and molecular dynamics simulations of ionic liquids". The Journal of Physical Chemistry B. 113 (33): 11463–78. doi:10.1021/jp905220k. PMID   19637900.
  111. Spicher S, Grimme S (September 2020). "Robust Atomistic Modeling of Materials, Organometallic, and Biochemical Systems". Angewandte Chemie International Edition. 59 (36): 15665–15673. doi: 10.1002/anie.202004239 . PMC   7267649 . PMID   32343883.
  112. Starovoytov ON (September 2021). "Development of Polarizable Force Field for Molecular Dynamics Simulation of Lithium-Ion Battery Electrolytes: Sulfonate Based Solvents and Lithium Salts". The Journal of Physical Chemistry B. 125 (40): 11242–11255. doi:10.1021/acs.jpcb.1c05744. PMID   34586817. S2CID   238230196.
  113. van Duin AC, Dasgupta S, Lorant F, Goddard WA (2001). "ReaxFF: A Reactive Force Field for Hydrocarbons" (PDF). The Journal of Physical Chemistry A . 105 (41): 9396–9409. Bibcode:2001JPCA..105.9396V. CiteSeerX   10.1.1.507.6992 . doi:10.1021/jp004368u. Archived from the original (PDF) on 2018-03-21. Retrieved 2015-08-29.
  114. Hoogerbrugge PJ, Koelman JM (1992). "Simulating Microscopic Hydrodynamic Phenomena with Dissipative Particle Dynamics". Europhysics Letters (EPL). 19 (3): 155–160. Bibcode:1992EL.....19..155H. doi:10.1209/0295-5075/19/3/001. ISSN   0295-5075. S2CID   250796817.
  115. Koelman JM, Hoogerbrugge PJ (1993). "Dynamic Simulations of Hard-Sphere Suspensions Under Steady Shear". Europhysics Letters (EPL). 21 (3): 363–368. Bibcode:1993EL.....21..363K. doi:10.1209/0295-5075/21/3/018. ISSN   0295-5075. S2CID   250913111.
  116. Español P, Warren P (1995). "Statistical Mechanics of Dissipative Particle Dynamics". Europhysics Letters (EPL). 30 (4): 191–196. Bibcode:1995EL.....30..191E. doi:10.1209/0295-5075/30/4/001. ISSN   0295-5075. S2CID   14385201.
  117. Dissipative Particle Dynamics: Addressing deficiencies and establishing new frontiers Archived 2010-07-15 at the Wayback Machine , CECAM workshop, July 16–18, 2008, Lausanne, Switzerland.
  118. "SAFT".
  119. Korkut A, Hendrickson WA (September 2009). "A force field for virtual atom molecular mechanics of proteins". Proceedings of the National Academy of Sciences of the United States of America. 106 (37): 15667–72. Bibcode:2009PNAS..10615667K. doi: 10.1073/pnas.0907674106 . PMC   2734882 . PMID   19717427.
  120. Batatia I, Kovacs DP, Simm G, Ortner C, Csanyi G (2022). "MACE: Higher order equivariant message passing neural networks for fast and accurate force fields". Advances in Neural Information Processing Systems. 35: 11423–11436. arXiv: 2206.07697 .
  121. Smith JS, Isayev O, Roitberg AE (April 2017). "ANI-1: an extensible neural network potential with DFT accuracy at force field computational cost". Chemical Science. 8 (4): 3192–3203. doi:10.1039/C6SC05720A. PMC   5414547 . PMID   28507695.
  122. Hughes ZE, Thacker JC, Wilson AL, Popelier PL (January 2019). "Description of Potential Energy Surfaces of Molecules Using FFLUX Machine Learning Models". Journal of Chemical Theory and Computation. 15 (1): 116–126. doi:10.1021/acs.jctc.8b00806. hdl: 10454/16776 . PMID   30507180. S2CID   54524604.
  123. Fletcher TL, Popelier PL (June 2016). "Multipolar Electrostatic Energy Prediction for all 20 Natural Amino Acids Using Kriging Machine Learning". Journal of Chemical Theory and Computation. 12 (6): 2742–51. doi: 10.1021/acs.jctc.6b00457 . PMID   27224739.
  124. McDonagh JL, Silva AF, Vincent MA, Popelier PL (January 2018). "Machine Learning of Dynamic Electron Correlation Energies from Topological Atoms". Journal of Chemical Theory and Computation. 14 (1): 216–224. doi: 10.1021/acs.jctc.7b01157 . PMID   29211469.
  125. Ramakrishnan R, Dral PO, Rupp M, von Lilienfeld OA (May 2015). "Big Data Meets Quantum Chemistry Approximations: The Δ-Machine Learning Approach". Journal of Chemical Theory and Computation. 11 (5): 2087–96. arXiv: 1503.04987 . Bibcode:2015arXiv150304987R. doi:10.1021/acs.jctc.5b00099. PMID   26574412. S2CID   28672393.
  126. Schütt KT, Sauceda HE, Kindermans PJ, Tkatchenko A, Müller KR (June 2018). "SchNet - A deep learning architecture for molecules and materials". The Journal of Chemical Physics. 148 (24): 241722. arXiv: 1712.06113 . Bibcode:2018JChPh.148x1722S. doi:10.1063/1.5019779. PMID   29960322. S2CID   4897444.
  127. O. T. Unke and M. Meuwly (2019). "PhysNet: A Neural Network for Predicting Energies, Forces, Dipole Moments, and Partial Charges". J. Chem. Theo. Chem. 15 (6): 3678–3693. arXiv: 1902.08408 . doi:10.1021/acs.jctc.9b00181. PMID   31042390. S2CID   85543050.
  128. Molinero V, Moore EB (April 2009). "Water modeled as an intermediate element between carbon and silicon". The Journal of Physical Chemistry B. 113 (13): 4008–16. arXiv: 0809.2811 . doi:10.1021/jp805227c. PMID   18956896. S2CID   20782587.
  129. Khoury GA, Thompson JP, Smadbeck J, Kieslich CA, Floudas CA (December 2013). "Ab Initio Charge and AMBER Forcefield Parameters for Frequently Occurring Post-Translational Modifications". Journal of Chemical Theory and Computation. 9 (12): 5653–5674. doi:10.1021/ct400556v. PMC   3904396 . PMID   24489522.
  130. Khoury GA, Smadbeck J, Tamamis P, Vandris AC, Kieslich CA, Floudas CA (December 2014). "Forcefield_NCAA: ab initio charge parameters to aid in the discovery and design of therapeutic proteins and peptides with unnatural amino acids and their application to complement inhibitors of the compstatin family". ACS Synthetic Biology. 3 (12): 855–69. doi:10.1021/sb400168u. PMC   4277759 . PMID   24932669.
  131. Khoury GA, Bhatia N, Floudas CA (2014). "Hydration free energies calculated using the AMBER ff03 charge model for natural and unnatural amino acids and multiple water models". Computers & Chemical Engineering. 71: 745–752. doi:10.1016/j.compchemeng.2014.07.017.
  132. Deeth RJ (2001). "The ligand field molecular mechanics model and the stereoelectronic effects of d and s electrons". Coordination Chemistry Reviews. 212 (212): 11–34. doi:10.1016/S0010-8545(00)00354-4.
  133. Foscato M, Deeth RJ, Jensen VR (June 2015). "Integration of Ligand Field Molecular Mechanics in Tinker". Journal of Chemical Information and Modeling. 55 (6): 1282–90. doi:10.1021/acs.jcim.5b00098. hdl: 1956/10456 . PMID   25970002.

Further reading