Space-filling model

Last updated
A space-filling model of n-octane, the straight chain (normal) hydrocarbon composed of 8 carbons and 18 hydrogens, formulae: CH3CH2(CH2)4CH2CH3 or C
8H
18. Note, the representative shown is of a single conformational "pose" of a population of molecules, which, because of low Gibbs energy barriers to rotation about its carbon-carbon bonds (giving the carbon "chain" great flexibility), normally is composed of a very large number of different such conformations (e.g., in solution). Octane.qutemol.png
A space-filling model of n-octane, the straight chain (normal) hydrocarbon composed of 8 carbons and 18 hydrogens, formulae: CH3CH2(CH2)4CH2CH3orC
8
H
18
. Note, the representative shown is of a single conformational "pose" of a population of molecules, which, because of low Gibbs energy barriers to rotation about its carbon-carbon bonds (giving the carbon "chain" great flexibility), normally is composed of a very large number of different such conformations (e.g., in solution).
An example of a three-dimensional, space-filling model of a complex molecule, THC, the active agent in marijuana. Thc.pdb.gif
An example of a three-dimensional, space-filling model of a complex molecule, THC, the active agent in marijuana.

In chemistry, a space-filling model, also known as a calotte model, is a type of three-dimensional (3D) molecular model where the atoms are represented by spheres whose radii are proportional to the radii of the atoms and whose center-to-center distances are proportional to the distances between the atomic nuclei, all in the same scale. Atoms of different chemical elements are usually represented by spheres of different colors.

Contents

Space-filling calotte models are also referred to as CPK models after the chemists Robert Corey, Linus Pauling, and Walter Koltun, who over a span of time developed the modeling concept into a useful form. [1] They are distinguished from other 3D representations, such as the ball-and-stick and skeletal models, by the use of the "full size" space-filling spheres for the atoms. They are useful for visualizing the effective shape and relative dimensions of the molecule, and the shapes of surface a given static conformer might present. On the other hand, these models mask the chemical bonds between the atoms, and make it difficult to see the structure of the molecule that is obscured by the atoms nearest to the viewer in a particular pose. For this reason, such models are of greater utility if they can be used dynamically, especially when used with complex molecules (e.g., see the greater understanding of the molecules shape given when the THC model is clicked on to rotate).

History

Space-filling models arise out of a desire to represent molecules in ways that reflect the electronic surfaces that molecules present, that dictate how they interact, one with another (or with surfaces, or macromolecules such as enzymes, etc.). Crystallographic data are the starting point for understanding static molecular structure, and these data contain the information rigorously required to generate space-filling representations (e.g., see these crystallographic models); most often, however, crystallographers present the locations of atoms derived from crystallography via "thermal ellipsoids" whose cut-off parameters are set for convenience both to show the atom locations (with anisotropies), and to allow representation of the covalent bonds or other interactions between atoms as lines. In short, for reasons of utility, crystallographic data historically have appeared in presentations closer to ball-and-stick models. Hence, while crystallographic data contain the information to create space-filling models, it remained for individuals interested in modeling an effective static shape of a molecule, and the space it occupied, and the ways in which it might present a surface to another molecule, to develop the formalism shown above.

In 1952, Robert Corey and Linus Pauling described accurate scale models of molecules which they had built at Caltech. [1] In their models, they envisioned the surface of the molecule as being determined by the van der Waals radius of each atom of the molecule, and crafted atoms as hardwood spheres of diameter proportional to each atom's van der Waals radius, in the scale 1 inch = 1 Å. To allow bonds between atoms a portion of each sphere was cut away to create a pair of matching flat faces, with the cuts dimensioned so that the distance between sphere centers was proportional to the lengths of standard types of chemical bonds. [1] A connector was designed—a metal bushing that threaded into each sphere at the center of each flat face. The two spheres were then firmly held together by a metal rod inserted into the pair of opposing bushing (with fastening by screws). The models also had special features to allow representation of hydrogen bonds. [1] [ verification needed ] [2]

An example of a 3D, space-filling model of a simple molecule, sulfur dioxide, SO2, showing the electrostatic potential surface, computed for the molecule using the Spartan software suite of computational chemistry tools. It is shaded from blue for electropositive areas to red for electronegative areas. The surface was generated by calculating the energy of interaction of a spherical point positive charge (e.g., a proton, H ) with the molecule's atoms and bonding electrons, in a series of discrete computational steps. Here, the electrostatic surface emphasizes the electron deficiency of the sulfur atom, suggesting interactions in which it might engage, and chemical reactions it might undergo. Sulfur-dioxide-elpot.png
An example of a 3D, space-filling model of a simple molecule, sulfur dioxide, SO2, showing the electrostatic potential surface, computed for the molecule using the Spartan software suite of computational chemistry tools. It is shaded from blue for electropositive areas to red for electronegative areas. The surface was generated by calculating the energy of interaction of a spherical point positive charge (e.g., a proton, H ) with the molecule's atoms and bonding electrons, in a series of discrete computational steps. Here, the electrostatic surface emphasizes the electron deficiency of the sulfur atom, suggesting interactions in which it might engage, and chemical reactions it might undergo.
An example of a 3D, space-filling model of a very complex macromolecule, a protein, the cell membrane-spanning b2 adrenoreceptor, a G protein-coupled receptor, in this image, viewed as if looking down onto the extracellular surface. The electrostatic potential surface was applied to a model with atom positions determined by crystallography (PDB code 2RH1); the electrostatic surface was computed using Adaptive Poisson-Boltzmann Solver (APBS) freeware. It is again shaded blue for electropositive areas to red for electronegative areas. Somewhat apparent, in stick representation in yellow, red and blue, in a groove at the top of the receptor, is a small molecule ligand bound to it, the agent carazolol, a partial inverse agonist which, through this binding, antagonizes binding of the normal ligand, the neurotransmitter/hormone epinephrine. In response to binding epinephrine, this receptor, in conjunction with an L-type calcium channel, mediates physiologic responses such as smooth muscle relaxation and bronchodilation. All of such binding interactions and the function of the receptor in signal transduction are mediated by electrostatic effects, and in modern structure work they are often studied using similar space filling models. Beta-2-adrenergic-receptor-electrostatic-top.png
An example of a 3D, space-filling model of a very complex macromolecule, a protein, the cell membrane-spanning β2 adrenoreceptor, a G protein-coupled receptor, in this image, viewed as if looking down onto the extracellular surface. The electrostatic potential surface was applied to a model with atom positions determined by crystallography (PDB code 2RH1); the electrostatic surface was computed using Adaptive Poisson-Boltzmann Solver (APBS) freeware. It is again shaded blue for electropositive areas to red for electronegative areas. Somewhat apparent, in stick representation in yellow, red and blue, in a groove at the top of the receptor, is a small molecule ligand bound to it, the agent carazolol, a partial inverse agonist which, through this binding, antagonizes binding of the normal ligand, the neurotransmitter/hormone epinephrine. In response to binding epinephrine, this receptor, in conjunction with an L-type calcium channel, mediates physiologic responses such as smooth muscle relaxation and bronchodilation. All of such binding interactions and the function of the receptor in signal transduction are mediated by electrostatic effects, and in modern structure work they are often studied using similar space filling models.

In 1965, Walter L. Koltun designed and patented a simplified system with molded plastic atoms of various colours, which were joined by specially designed snap connectors; this simpler system accomplished essentially the same ends as the Corey-Pauling system, [4] [ non-primary source needed ][ better source needed ] and allowed for the development of the models as a popular way of working with molecules in training and research environments. Such colour-coded, bond length-defined, van der Waal's-type space-filling models are now commonly known as CPK models, after these three developers of the specific concept.

In modern research efforts, attention returned to use of data-rich crystallographic models in combination with traditional and new computational methods to provide space-filling models of molecules, both simple and complex, where added information such as which portions of the surface of the molecule were readily accessible to solvent, or how the electrostatic characteristics of a space-filling representation—which in the CPK case is almost fully left to the imagination—could be added to the visual models created. The two closing images give examples of the latter type of calculation and representation, and its utility.

See also

Related Research Articles

Chemical bond Lasting attraction between atoms that enables the formation of chemical compounds

A chemical bond is a lasting attraction between atoms, ions or molecules that enables the formation of chemical compounds. The bond may result from the electrostatic force between oppositely charged ions as in ionic bonds or through the sharing of electrons as in covalent bonds. The strength of chemical bonds varies considerably; there are "strong bonds" or "primary bonds" such as covalent, ionic and metallic bonds, and "weak bonds" or "secondary bonds" such as dipole–dipole interactions, the London dispersion force and hydrogen bonding.

Hydrogen bond Hydrogen Partial intermolecular bonding interaction

A hydrogen bond is a primarily electrostatic force of attraction between a hydrogen (H) atom which is covalently bound to a more electronegative "donor" atom or group, and another electronegative atom bearing a lone pair of electrons—the hydrogen bond acceptor (Ac). Such an interacting system is generally denoted Dn–H···Ac, where the solid line denotes a polar covalent bond, and the dotted or dashed line indicates the hydrogen bond. The most frequent donor and acceptor atoms are the second-row elements nitrogen (N), oxygen (O), and fluorine (F).

Ionic bonding Chemical bonding involving attraction between ions

Ionic bonding is a type of chemical bonding that involves the electrostatic attraction between oppositely charged ions, or between two atoms with sharply different electronegativities, and is the primary interaction occurring in ionic compounds. It is one of the main types of bonding along with covalent bonding and metallic bonding. Ions are atoms with an electrostatic charge. Atoms that gain electrons make negatively charged ions. Atoms that lose electrons make positively charged ions. This transfer of electrons is known as electrovalence in contrast to covalence. In the simplest case, the cation is a metal atom and the anion is a nonmetal atom, but these ions can be of a more complex nature, e.g. molecular ions like NH+
4
or SO2−
4
. In simpler words, an ionic bond results from the transfer of electrons from a metal to a non-metal in order to obtain a full valence shell for both atoms.

Quantum chemistry, also called molecular quantum mechanics, is a branch of physical chemistry focused on the application of quantum mechanics to chemical systems, particularly towards the quantum-mechanical calculation of electronic contributions to physical and chemical properties of molecules, materials, and solutions at the atomic level. These calculations include systematically applied approximations intended to make calculations computationally feasible while still capturing as much information about important contributions to the computed wave functions as well as to observable properties such as structures, spectra, and thermodynamic properties. Quantum chemistry is also concerned with the computation of quantum effects on molecular dynamics and chemical kinetics.

The van der Waals radius, rw, of an atom is the radius of an imaginary hard sphere representing the distance of closest approach for another atom. It is named after Johannes Diderik van der Waals, winner of the 1910 Nobel Prize in Physics, as he was the first to recognise that atoms were not simply points and to demonstrate the physical consequences of their size through the van der Waals equation of state.

Van der Waals force Interactions between groups of atoms that dont arise from chemical bonds

In molecular physics, the van der Waals force, named after Dutch physicist Johannes Diderik van der Waals, 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.

Molecular mechanics 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.

Molecular modelling 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 chemistry, valence bond (VB) theory is one of the two basic theories, along with molecular orbital (MO) theory, that were developed to use the methods of quantum mechanics to explain chemical bonding. It focuses on how the atomic orbitals of the dissociated atoms combine to give individual chemical bonds when a molecule is formed. In contrast, molecular orbital theory has orbitals that cover the whole molecule.

Molecular geometry Study of the 3D shapes of molecules

Molecular geometry is the three-dimensional arrangement of the atoms that constitute a molecule. It includes the general shape of the molecule as well as bond lengths, bond angles, torsional angles and any other geometrical parameters that determine the position of each atom.

In chemistry, orbital hybridisation is the concept of mixing atomic orbitals to form new hybrid orbitals suitable for the pairing of electrons to form chemical bonds in valence bond theory. For example, in a carbon atom which forms four single bonds the valence-shell s orbital combines with three valence-shell p orbitals to form four equivalent sp3 mixtures in a tetrahedral arrangement around the carbon to bond to four different atoms. Hybrid orbitals are useful in the explanation of molecular geometry and atomic bonding properties and are symmetrically disposed in space. Usually hybrid orbitals are formed by mixing atomic orbitals of comparable energies.

A non-covalent interaction differs from a covalent bond in that it does not involve the sharing of electrons, but rather involves more dispersed variations of electromagnetic interactions between molecules or within a molecule. The chemical energy released in the formation of non-covalent interactions is typically on the order of 1–5 kcal/mol (1000–5000 calories per 6.02 × 1023 molecules). Non-covalent interactions can be classified into different categories, such as electrostatic, π-effects, van der Waals forces, and hydrophobic effects.

Force field (chemistry) Concept on molecular modeling

In the context of chemistry and molecular modelling, a force field is a computational method that is used to estimate the forces between atoms within molecules and also between molecules. More precisely, the force field refers to the functional form and parameter sets used to calculate the potential energy of a system of atoms or coarse-grained particles in molecular mechanics, molecular dynamics, or Monte Carlo simulations. The parameters for a chosen energy function may be derived from experiments in physics and chemistry, calculations in quantum mechanics, or both. Force fields are interatomic potentials and utilize the same concept as force fields in classical physics, with the difference that the force field parameters in chemistry describe the energy landscape, from which the acting forces on every particle are derived as a gradient of the potential energy with respect to the particle coordinates.

CPK coloring Colour convention for differentiating atoms

In chemistry, the CPK coloring is a popular color convention for distinguishing atoms of different chemical elements in molecular models.

Molecular graphics is the discipline and philosophy of studying molecules and their properties through graphical representation. IUPAC limits the definition to representations on a "graphical display device". Ever since Dalton's atoms and Kekulé's benzene, there has been a rich history of hand-drawn atoms and molecules, and these representations have had an important influence on modern molecular graphics.

A molecular model is a physical model of an atomistic system that represents molecules and their processes. They play an important role in understanding chemistry and generating and testing hypotheses. The creation of mathematical models of molecular properties and behavior is referred to as molecular modeling, and their graphical depiction is referred to as molecular graphics.

History of molecular theory Aspect of history

In chemistry, the history of molecular theory traces the origins of the concept or idea of the existence of strong chemical bonds between two or more atoms.

Ball-and-stick model

In chemistry, the ball-and-stick model is a molecular model of a chemical substance which is to display both the three-dimensional position of the atoms and the bonds between them. The atoms are typically represented by spheres, connected by rods which represent the bonds. Double and triple bonds are usually represented by two or three curved rods, respectively, or alternately by correctly positioned sticks for the sigma and pi bonds. In a good model, the angles between the rods should be the same as the angles between the bonds, and the distances between the centers of the spheres should be proportional to the distances between the corresponding atomic nuclei. The chemical element of each atom is often indicated by the sphere's color.

The Van der Waals surface of a molecule is an abstract representation or model of that molecule, illustrating where, in very rough terms, a surface might reside for the molecule based on the hard cutoffs of Van der Waals radii for individual atoms, and it represents a surface through which the molecule might be conceived as interacting with other molecules. Also referred to as a Van der Waals envelope, the Van der Waals surface is named for Johannes Diderik van der Waals, a Dutch theoretical physicist and thermodynamicist who developed theory to provide a liquid-gas equation of state that accounted for the non-zero volume of atoms and molecules, and on their exhibiting an attractive force when they interacted. Van der Waals surfaces are therefore a tool used in the abstract representations of molecules, whether accessed, as they were originally, via hand calculation, or via physical wood/plastic models, or now digitally, via computational chemistry software. Practically speaking, CPK models, developed by and named for Robert Corey, Linus Pauling, and Walter Koltun, were the first widely used physical molecular models based on Van der Waals radii, and allowed broad pedagogical and research use of a model showing the Van der Waals surfaces of molecules.

Pauling's principle of electroneutrality states that each atom in a stable substance has a charge close to zero. It was formulated by Linus Pauling in 1948 and later revised. The principle has been used to predict which of a set of molecular resonance structures would be the most significant, to explain the stability of inorganic complexes and to explain the existence of π-bonding in compounds and polyatomic anions containing silicon, phosphorus or sulfur bonded to oxygen; it is still invoked in the context of coordination complexes. However, modern computational techniques indicate many stable compounds have a greater charge distribution than the principle predicts.

References

  1. 1 2 3 4 Corey, Robert B.; Pauling, Linus (1953). "Molecular models of amino acids, peptides, and proteins" (PDF). Review of Scientific Instruments. 8 (24): 621–627. Bibcode:1953RScI...24..621C. doi:10.1063/1.1770803 . Retrieved 9 March 2020.
  2. In the same paper Corey and Pauling also briefly describe a much simpler but less accurate type of model, with rubber-like polyvinyl plastic spheres in the scale 1 inch = 2Å and connected by snap fasteners. See Corey & Pauling, 1953, op. cit.
  3. Baker, N.A., Sept, D., Joseph, S., Holst, M.J. & McCammon, J.A., 2001, "Electrostatics of nanosystems: Application to microtubules and the ribosome," Proc. Natl. Acad. Sci. U.S.A.98: pp. 10037-10041, see , and "Calculating Electrostatics". Archived from the original on 2015-06-24. Retrieved 2015-06-23., and , accessed 23 June 2015.
  4. Walter L. Koltun (1965), Space filling atomic units and connectors for molecular models. U. S. Patent 3170246.[ non-primary source needed ][ better source needed ]
A space-filling model of cyclohexane C
6H
12. Carbon atoms, partially masked, are in grey, and hydrogen atoms are presented as white spheres. Cyclohexanmolekulmodell.png
A space-filling model of cyclohexane C
6
H
12
. Carbon atoms, partially masked, are in grey, and hydrogen atoms are presented as white spheres.