You are on page 1of 7

Chirality (chemistry)

Chirality /kaɪˈrælɪti/ 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 center 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 Ancient
Greek word for hand, χείρ (cheir).

The mirror images of a chiral molecule or ion are called


enantiomers or optical isomers. Individual enantiomers are often
designated as either right-handed or left-handed. Chirality is an
essential consideration when discussing the stereochemistry in
organic and inorganic chemistry. The concept is of great practical Two enantiomers of a generic amino acid that are
chiral
importance because most biomolecules and pharmaceuticals are
chiral.

Chiral molecules and ions are described by various ways of


designating their absolute configuration, which codify either the
entity's geometry or its ability to rotate plane-polarized light, a
common technique in studying chirality.

(S)-Alanine (left) and (R)-alanine (right) in


zwitterionic form at neutral pH
Contents
Definition
Stereogenic centers
Manifestations of chirality
In biochemistry
In inorganic chemistry
Methods and practices
Miscellaneous nomenclature
History
See also
References
Further reading
External links

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 dissymmetric (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
Improper rotational elements (Sn)
axis (Cn)

Achiral Achiral
Chiral
mirror plane inversion centre
no Sn
S1 = σ S2 = i

C1

C2

Stereogenic centers
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). The stereogenic atom (also known as
the stereocenter) is usually carbon, as in many biological molecules. However a stereocenter can coincide with any atom, including
metals (as in many chiral coordination compounds), phosphorus, or sulfur. The low barrier of nitrogen inversion make most N-chiral
amines (NRR′R″) impossible to resolve, but P-chiral phosphines (PRR′R″) as well as S-chiral sulfoxides (OSRR′) are optically
stable.

While the presence of a stereogenic atom 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 atom. 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 (BINOL), 1,3-dichloroallene have axial chirality.
(E)-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. 1,1′-Bi-2-naphthol is an
example of a molecule
Even isotopic differences must be considered when examining chirality. Illustrative is the lacking point chirality.
derivative of benzyl alcohol PhCHDOH, which is chiral. The S enantiomer has [α]D = +0.715°.[6]

Manifestations of chirality
Flavour: the artificial sweetener aspartame has two enantiomers. L-aspartame tastes sweet whereas D-aspartame
is tasteless.[7]
Odor: R-(–)-carvone smells like spearmint whereas S-(+)-carvone smells like caraway.[8]
Drug effectiveness: the antidepressant drug Citalopram is sold as a racemic mixture. However, studies have
shown that only the (S)-(+) enantiomer is responsible for the drug's beneficial effects.[9][10]
Drug safety: D‑penicillamine is used in chelation therapy and for the treatment of rheumatoid arthritis whereas
L‑penicillamine is toxic as it inhibits the action of pyridoxine, an essential B vitamin.[11]

In biochemistry
Many biologically active molecules are chiral, including the naturally occurring amino acids (the building blocks of proteins) and
sugars.

The origin of this homochirality in biology is the subject of much debate.[12] 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.[13][14]

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.[12] Spearmint leaves contain the L-enantiomer of
the chemical carvone or R-(−)-carvone and caraway seeds contain the D-enantiomer or S-(+)-carvone.[15] 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 (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.[16]

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
tris(bipyridine)ruthenium(II) complex in which the three bipyridine ligands adopt
a chiral propeller-like arrangement.[17] 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). Also cf.
dextro- and levo- (laevo-).

Chiral ligands confer chirality to a metal complex, as illustrated by metal-amino


Delta-ruthenium-tris(bipyridine) cation
acid complexes. If the metal exhibits catalytic properties, its combination with a
chiral ligand is the basis of asymmetric catalysis.[18]

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
Any non-racemic chiral substance is called scalemic. Scalemic materials can be enantiopure or
enantioenriched.[19]
A chiral substance is enantiopure when only one of two possible enantiomers is present so that all molecules
within a sample have the same chirality sense. Use of homochiral as a synonym is strongly discouraged.[20]
A chiral substance is enantioenriched or heterochiral when its enantiomeric ratio is greater than 50:50 but less
than 100:0.[21]
Enantiomeric excess or e.e. is the difference between how much of one enantiomer is present compared to the
other. For example, a sample with 40% e.e. of R contains 70% R and 30% S (70% − 30% = 40%).[22]

History
The rotation of plane polarized light by chiral substances was first observed by Jean-Baptiste Biot in 1815,[23] and gained
considerable importance in the sugar industry, analytical chemistry, and pharmaceuticals. Louis Pasteur deduced in 1848 that this
phenomenon has a molecular basis.[24][25] The term chirality itself was coined by Lord Kelvin in 1894.[26] Different enantiomers or
diastereomers of a compound were formerly called optical isomers due to their different optical properties.[27] 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.

See also
Chemical chirality in popular fiction
Chirality (mathematics)
Chirality (physics)
Enantiopure drug
Enantioselective synthesis
Handedness
Orientation (vector space)
Pfeiffer effect
Stereochemistry for overview of stereochemistry in general
Stereoisomerism
Supramolecular chirality

References
1. Organic Chemistry (4th Edition) Paula Y. Bruice. Pearson Educational Books. ISBN 9780131407480
2. Organic Chemistry (3rd Edition) Marye Anne Fox, James K. Whitesell Jones & Bartlett Publishers (2004)
ISBN 0763721972
3. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "Chirality (http://goldbook.iupac.org/C01058.html)". doi:10.1351/goldbook.C01058 (https://doi.org/10.135
1%2Fgoldbook.C01058)
4. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "Superposability (http://goldbook.iupac.org/S06144.html)". doi:10.1351/goldbook.S06144 (https://doi.org/1
0.1351%2Fgoldbook.S06144)
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 (https://doi.org/10.1016%2F0040-4020%2859%2980014-4).
7. Gal, Joseph (2012). "The Discovery of Stereoselectivity at Biological Receptors: Arnaldo Piutti and the Taste of the
Asparagine Enantiomers-History and Analysis on the 125th Anniversary". Chirality. 24 (12): 959–976.
doi:10.1002/chir.22071 (https://doi.org/10.1002%2Fchir.22071). PMID 23034823 (https://www.ncbi.nlm.nih.gov/pub
med/23034823).
8. Theodore J. Leitereg; Dante G. Guadagni; Jean Harris; Thomas R. Mon; 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 (https://doi.org/10.1021%2Fjf60176a035).
9. Lepola U, Wade A, Andersen HF (May 2004). "Do equivalent doses of escitalopram and citalopram have similar
efficacy? A pooled analysis of two positive placebo-controlled studies in major depressive disorder". Int Clin
Psychopharmacol. 19 (3): 149–55. doi:10.1097/00004850-200405000-00005 (https://doi.org/10.1097%2F00004850
-200405000-00005). PMID 15107657 (https://www.ncbi.nlm.nih.gov/pubmed/15107657).
10. Hyttel, J.; Bøgesø, K. P.; Perregaard, J.; Sánchez, C. (1992). "The pharmacological effect of citalopram resides in
the (S)-(+)-enantiomer". Journal of Neural Transmission. 88 (2): 157–160. doi:10.1007/BF01244820 (https://doi.org/
10.1007%2FBF01244820). PMID 1632943 (https://www.ncbi.nlm.nih.gov/pubmed/1632943).
11. JAFFE, IA; ALTMAN, K; MERRYMAN, P (Oct 1964). "The Antipyridoxine Effect of Penicillamine in Man" (https://ww
w.ncbi.nlm.nih.gov/pmc/articles/PMC289631). The Journal of Clinical Investigation. 43 (10): 1869–73.
doi:10.1172/JCI105060 (https://doi.org/10.1172%2FJCI105060). PMC 289631 (https://www.ncbi.nlm.nih.gov/pmc/ar
ticles/PMC289631). PMID 14236210 (https://www.ncbi.nlm.nih.gov/pubmed/14236210).
12. Meierhenrich, Uwe J. (2008). Amino acids and the Asymmetry of Life. Berlin, GER: Springer. ISBN 978-
3540768852.
13. McKee, Maggie (2005-08-24). "Space radiation may select amino acids for life" (https://www.newscientist.com/articl
e/dn7895-space-radiation-may-select-amino-acids-for-life.html). New Scientist. Retrieved 2016-02-05.
14. Meierhenrich Uwe J., Nahon Laurent, Alcaraz Christian, Hendrik Bredehöft Jan, Hoffmann Søren V., Barbier
Bernard, Brack André (2005). "Asymmetric Vacuum UV photolysis of the Amino Acid Leucine in the Solid State".
Angew. Chem. Int. Ed. 44 (35): 5630–5634. doi:10.1002/anie.200501311 (https://doi.org/10.1002%2Fanie.2005013
11). PMID 16035020 (https://www.ncbi.nlm.nih.gov/pubmed/16035020).
15. Theodore J. Leitereg; Dante G. Guadagni; Jean Harris; Thomas R. Mon; 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 (https://doi.org/10.1021%2Fjf60176a035).
16. Srinivasarao, M. (1999). "Chirality and Polymers". Current Opinion in Colloid & Interface Science. 4 (5): 369–376.
doi:10.1016/S1359-0294(99)00024-2 (https://doi.org/10.1016%2FS1359-0294%2899%2900024-2).
17. von Zelewsky, A. (1995). Stereochemistry of Coordination Compounds. Chichester: John Wiley..
ISBN 047195599X.
18. Hartwig, J. F. Organotransition Metal Chemistry, from Bonding to Catalysis; University Science Books: New York,
2010. ISBN 189138953X
19. Eliel, E.L. (1997). "Infelicitous Stereochemical Nomenclatures" (https://web.archive.org/web/20160303230750/http://
www.uottawa.ca/publications/interscientia/inter.4/eliel/eliel.html). Chirality. 9 (56): 428–430. doi:10.1002/(sici)1520-
636x(1997)9:5/6<428::aid-chir5>3.3.co;2-e (https://doi.org/10.1002%2F%28sici%291520-636x%281997%299%3A
5%2F6%3C428%3A%3Aaid-chir5%3E3.3.co%3B2-e). Archived from the original (https://www.uottawa.ca/publicatio
ns/interscientia/inter.4/eliel/eliel.html) on 3 March 2016. Retrieved 5 February 2016.
20. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "asymmetric synthesis (http://goldbook.iupac.org/E02072.html)". doi:10.1351/goldbook.E02072 (https://doi.
org/10.1351%2Fgoldbook.E02072)
21. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "enantiomerically enriched (enantioenriched) (http://goldbook.iupac.org/E02071.html)".
doi:10.1351/goldbook.E02071 (https://doi.org/10.1351%2Fgoldbook.E02071)
22. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "enantiomer excess (enantiomeric excess) (http://goldbook.iupac.org/E02070.html)".
doi:10.1351/goldbook.E02070 (https://doi.org/10.1351%2Fgoldbook.E02070)
23. Lakhtakia, A. (ed.) (1990). Selected Papers on Natural Optical Activity (SPIE Milestone Volume 15). SPIE.
24. 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".
25. Eliel, Ernest Ludwig; Wilen, Samuel H.; Mander, Lewis N. (1994). "Chirality in Molecules Devoid of Chiral Centers
(Chapter 14)" (https://books.google.com/books?id=IyfwAAAAMAAJ). Stereochemistry of Organic Compounds (1st
ed.). New York, NY, USA: Wiley & Sons. ISBN 978-0471016700. Retrieved 2 February 2016.
26. 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 (https://doi.org/10.1353%2Fpbm.199
5.0069). PMID 7899056 (https://www.ncbi.nlm.nih.gov/pubmed/7899056).
27. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "Optical isomers (http://goldbook.iupac.org/O04308.html)". doi:10.1351/goldbook.O04308 (https://doi.org/1
0.1351%2Fgoldbook.O04308)

Further reading
Clayden, Jonathan; Greeves, Nick; Warren, Stuart (2012). Organic Chemistry (https://books.google.com/books?isb
n=0199270295) (2nd ed.). Oxford, UK: Oxford University Press. pp. 319f, 432, 604np, 653, 746int, 803ketals, 839,
846f. ISBN 978-0199270293. Retrieved 2 February 2016.
Eliel, Ernest Ludwig; Wilen, Samuel H.; Mander, Lewis N. (1994). "Chirality in Molecules Devoid of Chiral Centers
(Chapter 14)" (https://books.google.com/books?id=IyfwAAAAMAAJ). Stereochemistry of Organic Compounds.
Chirality. 9 (1st ed.). New York, NY, USA: Wiley & Sons. pp. 428–430. doi:10.1002/(SICI)1520-
636X(1997)9:5/6<428::AID-CHIR5>3.0.CO;2-1 (https://doi.org/10.1002%2F%28SICI%291520-636X%281997%29
9%3A5%2F6%3C428%3A%3AAID-CHIR5%3E3.0.CO%3B2-1). ISBN 978-0471016700. Retrieved 2 February
2016.
Eliel, E.L. (1997). "Infelicitous Stereochemical Nomenclatures" (https://web.archive.org/web/20160303230750/http://
www.uottawa.ca/publications/interscientia/inter.4/eliel/eliel.html). Chirality. 9 (5–6): 428–430.
doi:10.1002/(SICI)1520-636X(1997)9:5/6<428::AID-CHIR5>3.0.CO;2-1 (https://doi.org/10.1002%2F%28SICI%2915
20-636X%281997%299%3A5%2F6%3C428%3A%3AAID-CHIR5%3E3.0.CO%3B2-1). Archived from the original (h
ttps://www.uottawa.ca/publications/interscientia/inter.4/eliel/eliel.html) on 3 March 2016. Retrieved 5 February 2016.
Gal, Joseph (2013). "Molecular Chirality: Language, History, and Significance". Differentiation of Enantiomers I.
Chirality. Topics in Current Chemistry. 340. pp. 1–20. doi:10.1007/128_2013_435 (https://doi.org/10.1007%2F128_2
013_435). ISBN 978-3-319-03238-2. PMID 23666078 (https://www.ncbi.nlm.nih.gov/pubmed/23666078).

External links
21st International Symposium on Chirality (http://www.chirality2009.org/)
STEREOISOMERISM - OPTICAL ISOMERISM (http://www.chemguide.co.uk/basicorg/isomerism/optical.html#top)
Symposium highlights-Session 5: New technologies for small molecule synthesis (http://www.nature.com/horizon/ch
emicalspace/highlights/s5_nonspec1.html)
IUPAC nomenclature for amino acid configurations. (http://www.chem.qmul.ac.uk/iupac/AminoAcid/AA3t5.html)
Michigan State University's explanation of R/S nomenclature (http://arquivo.pt/wayback/20160521132055/http://ww
w.cem.msu.edu/~reusch/VirtualText/sterism3.htm)
Chirality & Odour Perception at leffingwell.com (http://www.leffingwell.com/chirality/chirality.htm)
Chirality & Bioactivity I.: Pharmacology (http://www.leffingwell.com/download/chirality-phamacology.pdf)
Chirality and the Search for Extraterrestrial Life (http://www.eurekalert.org/pub_releases/2009-04/nios-sga042309.p
hp)
The Handedness of the Universe by Roger A Hegstrom and Dilip K Kondepudi
http://quantummechanics.ucsd.edu/ph87/ScientificAmerican/Sciam/Hegstrom_The_Handedness_of_the_universe.pdf

Retrieved from "https://en.wikipedia.org/w/index.php?title=Chirality_(chemistry)&oldid=922155951"

This page was last edited on 20 October 2019, at 09:19 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like