Chirality (chemistry)


In chemistry, a molecule or ion is called chiral if it cannot be superposed on its mirror image by any combination of rotations and translations. This geometric property is called chirality. The terms are derived from Ancient Greek χείρ, meaning "hand"; which is the canonical example of an object with this property.
A chiral molecule or ion exists in two stereoisomers that are mirror images of each other, called enantiomers; they are often distinguished as either "right-handed" or "left-handed" by their absolute configuration or some other criterion. The two enantiomers have the same chemical properties, except when reacting with other chiral compounds. They also have the same physical properties, except that they often have opposite optical activities. A homogeneous mixture of the two enantiomers in equal parts is said to be racemic, and it usually differs chemically and physically from the pure enantiomers.
A chiral molecule or ion must have at least one chiral center or stereocenter. When that center coincides with an atom, the substance is said to have point chirality. In chiral organic compounds, a stereocenter is often an asymmetric carbon. Multiple stereocenters may give rise to additional stereoisomers. On the other hand, a molecule with an even number of stereocenters may have one or more stereoisomers which are not chiral.
Chirality is an important concept for stereochemistry and biochemistry. Most substances relevant to biology are chiral, such as carbohydrates, the amino acids that are the building blocks of proteins, and the nucleic acids. In living organisms, one typically finds only one of the two enantiomers or a chiral compound. For that reason, organisms that consume a chiral compound usually can metabolize only one of its enantiomers. For the same reason, the two enantiomers of a chiral pharmaceutical usually have vastly different potencies and/or effects.

Definition

Chirality is based on molecular symmetry. Specifically, a chiral compound can contain no improper axis of rotation, which includes planes of symmetry and inversion center. Chiral molecules are always dissymmetric but not always asymmetric. Asymmetric molecules are always chiral.

Stereogenic centers

Many chiral molecules have point chirality, namely a single stereogenic center that coincides with an atom. This stereogenic center usually has four or more bonds to different groups, and may be carbon, silicon, or a metal. However, a stereogenic center can also be a trivalent atom whose bonds are not in the same plane, such as phosphorus in P-chiral phosphines and sulfur in S-chiral sulfoxides.
Chirality can also arise from isotopic differences between atoms, such as in the deuterated benzyl alcohol PhCHDOH; which is chiral and optically active, even though the non-deuterated compound PhCH2OH is not.
If two enantiomers easily interconvert, the pure enantiomers may be practically impossible to separate, and only the racemic mixture is observable. This is the case, for example, of most amines with three different substituents, because of the low energy barrier for nitrogen inversion.
While the presence of a stereogenic center describes the great majority of chiral molecules, many variations and exceptions exist. For instance it is not necessary for the chiral substance to have a stereogenic center. Examples include 1-bromo-3-chloro-5-fluoroadamantane, methylethylphenyltetrahedrane, certain calixarenes and fullerenes, which have inherent chirality. The C2-symmetric species 1,1′-bi-2-naphthol, 1,3-dichloroallene have axial chirality. -cyclooctene and many ferrocenes have planar chirality.
When the optical rotation for an enantiomer is too low for practical measurement, the species is said to exhibit cryptochirality.

Manifestations of chirality

Many biologically active molecules are chiral, including the naturally occurring amino acids and sugars.
The origin of this homochirality in biology is the subject of much debate. Most scientists believe that Earth life's "choice" of chirality was purely random, and that if carbon-based life forms exist elsewhere in the universe, their chemistry could theoretically have opposite chirality. However, there is some suggestion that early amino acids could have formed in comet dust. In this case, circularly polarised radiation could have caused the selective destruction of one chirality of amino acids, leading to a selection bias which ultimately resulted in all life on Earth being homochiral.
Enzymes, which are chiral, often distinguish between the two enantiomers of a chiral substrate. One could imagine an enzyme as having a glove-like cavity that binds a substrate. If this glove is right-handed, then one enantiomer will fit inside and be bound, whereas the other enantiomer will have a poor fit and is unlikely to bind.
-forms of amino acids tend to be tasteless, whereas -forms tend to taste sweet. Spearmint leaves contain the -enantiomer of the chemical carvone or R--carvone and caraway seeds contain the -enantiomer or S--carvone. The two smell different to most people because our olfactory receptors are chiral.
Chirality is important in context of ordered phases as well, for example the addition of a small amount of an optically active molecule to a nematic phase transforms that phase to a chiral nematic phase. Chirality in context of such phases in polymeric fluids has also been studied in this context.

In inorganic chemistry

Chirality is a symmetry property, not a property of any part of the periodic table. Thus many inorganic materials, molecules, and ions are chiral. Quartz is an example from the mineral kingdom. Such noncentric materials are of interest for applications in nonlinear optics.
In the areas of coordination chemistry and organometallic chemistry, chirality is pervasive and of practical importance. A famous example is trisruthenium complex in which the three bipyridine ligands adopt a chiral propeller-like arrangement. The two enantiomers of complexes such as 2+ may be designated as Λ for a left-handed twist of the propeller described by the ligands, and Δ for a right-handed twist. Also cf. dextro- and levo-.
Chiral ligands confer chirality to a metal complex, as illustrated by metal-amino acid complexes. If the metal exhibits catalytic properties, its combination with a chiral ligand is the basis of asymmetric catalysis.

Methods and practices

The term optical activity is derived from the interaction of chiral materials with polarized light. In a solution, the -form, or levorotatory form, of an optical isomer rotates the plane of a beam of linearly polarized light counterclockwise. The -form, or dextrorotatory form, of an optical isomer does the opposite. The rotation of light is measured using a polarimeter and is expressed as the optical rotation.

Miscellaneous nomenclature

The rotation of plane polarized light by chiral substances was first observed by Jean-Baptiste Biot in 1815, and gained considerable importance in the sugar industry, analytical chemistry, and pharmaceuticals. Louis Pasteur deduced in 1848 that this phenomenon has a molecular basis. The term chirality itself was coined by Lord Kelvin in 1894. Different enantiomers or diastereomers of a compound were formerly called optical isomers due to their different optical properties. At one time, chirality was thought to be associated with organic chemistry, but this misconception was overthrown by the resolution of a purely inorganic compound, hexol, by Alfred Werner.