Chirality (chemistry)

"L-form" redirects here. For the bacterial strains, see L-form bacteria.
Two enantiomers of a generic amino acid that is chiral
(S)-Alanine (left) and (R)-alanine (right) in zwitterionic form at neutral pH

Chirality /kˈrælɪti/ is a geometric property of some molecules and ions. A chiral molecule/ion is non-superposable on its mirror image. The presence of an asymmetric carbon atom is one of several structural features that induce chirality in organic and inorganic molecules.[1] [2] [3][4] The term chirality is derived from the Greek word for hand, χειρ (kheir).

The mirror images of a chiral molecule/ion are called enantiomers or optical isomers. Individual enantiomers are often designated as either "right-" or "left-handed". Chirality is an essential consideration when discussing the stereochemistry in inorganic chemistry and organic chemistry. The concept is of great practical importance because most biomolecules and pharmaceuticals are chiral.

Chiral molecules and ions are described by various ways of designating their ″absolute configuration″ which codifies either the entity's geometry or its ability to rotate plane-polarized light, a common technique in studying chirality. Various naming conventions used to describe a given chiral entity's absolute configuration are explained here.

Definition

Chirality is based on molecular symmetry elements. Specifically, a chiral compound can contain no improper axis of rotation (Sn), which includes planes of symmetry and inversion center. Chiral molecules are always dysymmetric (lacking Sn) but not always asymmetric (lacking all symmetry elements except the trivial identity). Asymmetric molecules are always chiral.[5]

Molecular symmetry and chirality
Rotational
axis (Cn)
Improper rotational elements (Sn)
  Chiral
no Sn
Achiral
mirror plane
S1 = σ
Achiral
inversion centre
S2 = i
C1
C2

Stereogenic centers

Main article: Stereogenic center

In general, chiral molecules have point chirality at a single stereogenic atom, which has four different substituents. The two enantiomers of such compounds are said to have different absolute configurations at this center. This center is thus stereogenic (i.e., a grouping within a molecular entity that may be considered a focus of stereoisomerism).

Normally, when a tetrahedral atom has four different substituents it is chiral. However, in rare cases, if two of the ligands differ from each other by being mirror images of each other, the mirror image of the molecule is identical to the original, and the molecule is achiral. This is called pseudochirality.

A molecule can have multiple stereogenic centers without being chiral overall if there is a symmetry between the two (or more) stereocenters themselves. Such a molecule is called a meso compound.

It is also possible for a molecule to be chiral without having actual point chirality. Common examples include 1,1'-bi-2-naphthol (BINOL), 1,3-dichloro-allene, and BINAP, which have axial chirality, (E)-cyclooctene, which has planar chirality, and certain calixarenes and fullerenes, which have inherent chirality.

A form of point chirality can also occur if a molecule contains a tetrahedral subunit which cannot easily rearrange, for instance 1-bromo-1-chloro-1-fluoroadamantane and methylethylphenyltetrahedrane.

It is important to keep in mind that molecules have considerable flexibility and thus, depending on the medium, may adopt a variety of different conformations. These various conformations are themselves almost always chiral. When assessing chirality, a time-averaged structure is considered and for routine compounds, one should refer to the most symmetric possible conformation.

When the optical rotation for an enantiomer is too low for practical measurement, it is said to exhibit cryptochirality.

Even isotopic differences must be considered when examining chirality. Replacing one of the two 1H atoms at the CH2 position of benzyl alcohol with a deuterium (2H) makes that carbon a stereocenter. The resulting benzyl-α-d alcohol exists as two distinct enantiomers, which can be assigned by the usual stereochemical naming conventions. The S enantiomer has [α]D = +0.715°.[6]

The identity of the stereogenic atom

The stereogenic atom in chiral molecules is usually carbon, as in many biological molecules. However chirality can exist in any atom, including metals (as in many chiral coordination compounds), phosphorus, or sulfur. Chiral nitrogen is equally possible, although the effects of nitrogen inversion can make many of these compound impossible to isolate.

The chiral atom Carbon Nitrogen Phosphorus (phosphates) Phosphorus (phosphines) Sulfur Metal (type of metal)
1 stereogenic center Serine, glyceraldehyde Cyclanoline Sarin, VX Esomeprazole, armodafinil Tris(bipyridine)ruthenium(II) (ruthenium), cis-Dichlorobis(ethylenediamine)cobalt(III) (cobalt), hexol (cobalt)
2 stereogenic centers Threonine, isoleucine Tröger's base Adenosine triphosphate DIPAMP Dithionous acid
3 or more stereogenic centers Met-enkephalin, leu-enkephalin DNA

In biochemistry

Many biologically active molecules are chiral, including the naturally occurring amino acids (the building blocks of proteins) and sugars. In biological systems, most of these compounds are of the same chirality: most amino acids are levorotatory (L) and sugars are dextrorotatory (D). Typical naturally occurring proteins, made of L amino acids, are known as left-handed proteins, whereas D amino acids produce right-handed proteins . D-amino acids are very rare in nature and have only been found in small peptides attached to bacteria cell walls.

The origin of this homochirality in biology is the subject of much debate.[7] 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 (which makes up 17% of stellar 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.[8]

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.

L-forms of amino acids tend to be tasteless, whereas D-forms tend to taste sweet.[7] Spearmint leaves contain the L-enantiomer of the chemical carvone or R-()-carvone and caraway seeds contain the D-enantiomer or S-(+)-carvone.[9] These smell different to most people because our olfactory receptors also contain chiral molecules that behave differently in the presence of different enantiomers.

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 (a phase that has long range orientational order of molecules) transforms that phase to a chiral nematic phase (or cholesteric phase). Chirality in context of such phases in polymeric fluids has also been studied in this context.[10]

Inorganic chemistry

Delta-ruthenium-tris(bipyridine) cation

Chirality is a symmetry property, not a characteristic 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 tris(bipyridine)ruthenium(II) complex in which the three bipyridine ligands adopt a chiral propeller-like arrangement.[11] The two enantiomers of complexes such as [Ru(2,2′-bipyridine)3]2+ may be designated as Λ (capital lambda, the Greek version of "L") for a left-handed twist of the propeller described by the ligands, and Δ (capital delta, Greek "D") for a right-handed twist (pictured).

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.[12]

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

History

The rotation of plane polarized light by chiral substances was first observed by Jean-Baptiste Biot in 1815,[17] and gained considerable importance in the sugar industry, analytical chemistry, and pharmaceuticals. Louis Pasteur deduced in 1848 that this phenomenon has a molecular basis.[18][19] The term chirality itself was coined by Lord Kelvin in 1894.[20] Different enantiomers or diastereomers of a compound were formerly called optical isomers due to their different optical properties.[21] 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.

Further reading

The following are principle sources for the developing content in this article, that contain further information that may be of interest to readers. In the books listed, the material on chirality appears throughout, and so page numbers should be given for specific ideas and quotations drawn from them that are introduced above.

See also

References

  1. Organic Chemistry (4th Edition) Paula Y. Bruice.
  2. Organic Chemistry (3rd Edition) Marye Anne Fox ,James K. Whitesell.
  3. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "Chirality".
  4. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "Superposability".
  5. Cotton, F. A., "Chemical Applications of Group Theory," John Wiley & Sons: New York, 1990.
  6. ^ Streitwieser, A., Jr.; Wolfe, J. R., Jr.; Schaeffer, W. D. (1959). "Stereochemistry of the Primary Carbon. X. Stereochemical Configurations of Some Optically Active Deuterium Compounds". Tetrahedron 6 (4): 338–344. doi:10.1016/0040-4020(59)80014-4.
  7. 1 2 Meierhenrich, Uwe J. (2008). Amino acids and the Asymmetry of Life. Berlin, GER: Springer. ISBN 3540768858.
  8. McKee, Maggie (2005-08-24). "Space radiation may select amino acids for life". New Scientist. Retrieved 2016-02-05.
  9. Theodore J. Leitereg, Dante G. Guadagni, Jean Harris, Thomas R. Mon, and Roy Teranishi (1971). "Chemical and sensory data supporting the difference between the odors of the enantiomeric carvones". J. Agric. Food Chem. 19 (4): 785–787. doi:10.1021/jf60176a035.
  10. Srinivasarao, M. (1999). "Chirality and Polymers". Current Opinion in Colloid and Interface Science 4 (5): 369–376.
  11. von Zelewsky, A. (1995). Stereochemistry of Coordination Compounds. Chichester: John Wiley.. ISBN 047195599X.
  12. Hartwig, J. F. Organotransition Metal Chemistry, from Bonding to Catalysis; University Science Books: New York, 2010. ISBN 189138953X
  13. Eliel, E.L. (1997). "Infelicitous Stereochemical Nomenclatures". Chirality 9: 428–430. Retrieved 5 February 2016.
  14. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "asymmetric synthesis".
  15. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "enantiomerically enriched (enantioenriched)".
  16. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "enantiomer excess (enantiomeric excess)".
  17. Lakhtakia, A. (ed.) (1990). Selected Papers on Natural Optical Activity (SPIE Milestone Volume 15). SPIE.
  18. Pasteur, L. (1848). "Researches on the molecular asymmetry of natural organic products, English translation of French original, published by Alembic Club Reprints (Vol. 14, pp. 1–46) in 1905, facsimile reproduction by SPIE in a 1990 book".
  19. Eliel, Ernest Ludwig; Wilen, Samuel H. & Mander, Lewis N. (1994). "Chirality in Molecules Devoid of Chiral Centers (Chapter 14)". Stereochemistry of Organic Compounds (1st ed.). New York, NY, USA: Wiley & Sons. ISBN 0471016705. Retrieved 2 February 2016. For a further but less stable source of the same text that provides access to the relevant material, see , same access date.
  20. Bentley, Ronald (1995). "From Optical Activity in Quartz to Chiral Drugs: Molecular Handedness in Biology and Medicine.". Perspect. Biol. Med. 38 (2): 188–229. doi:10.1353/pbm.1995.0069. PMID 7899056. Retrieved 2 February 2016.
  21. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:  (2006) "Optical isomers".

External links

Wikimedia Commons has media related to Chirality.
This article is issued from Wikipedia - version of the Monday, April 25, 2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.