Homochirality

Jump to navigation Jump to search

Homochirality is a term used to refer to a group of molecules that possess the same sense of chirality. Molecules involved are not necessarily the same compound, but similar groups are arranged in the same way around a central atom. In biology homochirality is found inside living organisms. Active forms of amino acids are all of the L-form and most biologically relevant sugars are of the D-form. Typically, the alternative form is inactive and sometimes even toxic to living things. The origin of this phenomenon is not clearly understood. It is even unclear if homochirality has a purpose. One suggestion is that it reduces entropy barriers in the formation of large organized molecules. It has been experimentally verified that amino acids form large aggregates in larger abundance from enantiopure substrates than from racemic ones. [1].

Homochirality is said to evolve in three distinct steps: mirror-symmetry breaking creates a minute enantiomeric imbalance and is key to homochirality, chiral amplification is a process of enantiomeric enrichment and chiral transmission allows the transfer of chirality of one set of molecules to another.

Mirror-symmetry breaking

Explaining how an enantiomeric imbalance is created is the first place is the most difficult question to answer. Supporters exist for an extraterrestrial origin based on findings relating to the Murchison meteorite. There is evidence for the existence of circularly polarized light in space (originating from white dwarfs) which may trigger the formation of optical isomers.

One classic study involves an experiment that takes place in the laboratory.[2] When sodium chlorate is allowed to crystallize from water and the collected crystals examined in a polarimeter, each crystal turns out to be chiral and either the L form or the D form. In an ordinary experiment the amount of L crystals collected equals the amount of D crystals (corrected for statistical effects). However when the sodium chlorate solution is stirred during the crystallization process the crystals are either exclusively L or exclusively D. In 32 consecutive crystallization experiments 14 experiments deliver D-crystals and 18 others L-crystals. The explanation for this symmetry breaking is unclear but is related to autocatalysis taking place in the nucleation process.

Chiral amplification

Laboratory experiments exist demonstrating how in certain autocatalytic reaction systems the presence of a small amount of reaction product with enantiomeric excess at the start of the reaction can result in a much larger enantiomeric excess at the end of the reaction. In one pioneering study,[3] pyrimidine-5-carbaldehyde (Scheme 1) is alkylated by diisopropylzinc to the corresponding pyrimidyl alcohol. Because the initial reaction product is also an effective catalyst the reaction is autocatalytic. The presence of just 0.2 equivalent of the alcohol S-enantiomer at the start of the reaction is sufficient to amplify the enantiomeric excess to 93%.

Scheme 1. Soai autocatalysis
Scheme 1. Soai autocatalysis

Another study [4] concerns the proline catalyzed aminoxylation of propionaldehyde by nitrosobenzene (scheme 2). In this system too the presence of enantioenriched catalyst drives the reaction towards one of the two possible optical isomers.

Scheme 2. Blackmond autocatalysis
Scheme 2. Blackmond autocatalysis

Serine octamer clusters [5][6] are also contenders. These clusters of 8 serine molecules appear in mass spectroscopy with an unusual homochiral preference, however there is no evidence that such clusters exist under non-ionizing conditions and amino acid phase behavior is far more prebiotically relevant [7]. The recent observation that partial sublimation of a 10% enantioenriched sample of leucine results in up to 82% enrichment in the sublimate shows that enantioenrichment of amino acids could occur in space [8]. Partial sublimation processes can take place on the surface of meteors where large variations in temperature exist. This finding may have consequences for the development of the Mars Organic Detector scheduled for lauch in 2013 which aims to recover trace amounts of amino acids from the Mars surface exactly by a sublimation technique.

A high asymmetric amplification of the enantiomeric excess of sugars are also present in the amino acid catalyzed asymmetric formation of carbohydrates[9]

Chiral transmission

Many strategies in asymmetric synthesis are built on chiral transmission. Especially important is the so-called organocatalysis of organic reactions by proline for example in Mannich reactions.

Optical resolution in racemic amino acids

There exists no theory elucidating correlations among L-amino acids. If one takes, for example, alanine, which has a small methyl group, and phenylalanine, which has a big benzyl group, a simple question is in what aspect, L-alanine resembles L-phenylalanine more than D-phenylalanine, and what kind of mechanism causes the selection of all L-amino acids. Because it might be possible that alanine was L and phenylalanine was D.

It was reported[10] in 2004 that excess racemic D,L-asparagine (Asn), which spontaneously forms crystals of either isomer during recrystallization, induces asymmetric resolution of a co-existing racemic amino acid such as arginine (Arg), aspartic acid (Asp), glutamine (Gln), histidine (His), leucine (Leu), methionine (Met), phenylalanine (Phe), serine (Ser), valine (Val), tyrosine (Tyr), and tryptophan (Trp). The enantiomeric excess {ee=100x(L-D)/(L+D)} of these amino acids was correlated almost linearly with that of the inducer, i.e., Asn. When recrystallizations from a mixture of 12 D,L-amino acids (Ala, Asp, Arg, Glu, Gln, His, Leu, Met, Ser, Val, Phe, and Tyr) and excess D,L-Asn were made, all amino acids with the same configuration with Asn were preferentially co-crystallized.[10] It was incidental whether the enrichment took place in L- or D-Asn, however, once the selection was made, the co-existing amino acid with the same configuration at the α-carbon was preferentially involved because of thermodynamic stability in the crystal formation. The maximal ee was reported to be 100%. Based on these results, it is proposed that a mixture of racemic amino acids causes spontaneous and effective optical resolution, even if asymmetric synthesis of a single amino acid does not occur without an aid of an optically active molecule.

This is the first study elucidating reasonably the formation of chirality from racemic amino acids with experimental evidences.

History

This term was introduced by Kelvin in 1904, the year that published his Baltimore Lecture of 1884.[11][9] Recently, however, homochiral has been used in the same sense as enantiomerically pure. This is permitted in some journals (but not encouraged), its meaning changing into the preference of a process or system for a single optical isomer in a pair of isomers in these journals.

See also

References

  1. Do Homochiral Aggregates Have an Entropic Advantage? Julian, R. R.; Myung, S.; Clemmer, D. E. J. Phys. Chem. B.; (Article); 2005; 109(1); 440-444. doi:10.1021/jp046478x
  2. Kondepudi, D. K., Kaufman, R. J. & Singh, N. (1990). "Chiral Symmetry Breaking in Sodium Chlorate Crystallization". Science. 250: 975–976.
  3. Takanori Shibata, Hiroshi Morioka, Tadakatsu Hayase, Kaori Choji, and Kenso Soai (1996). "Highly Enantioselective Catalytic Asymmetric Automultiplication of Chiral Pyrimidyl Alcohol". J. Am. Chem. Soc. 118 (2): 471–472. doi:10.1021/ja953066g.
  4. Suju P. Mathew, Hiroshi Iwamura and Donna G. Blackmond (21 Jun 2004). "Amplification of Enantiomeric Excess in a Proline-Mediated Reaction". Angewandte Chemie International Edition. 43 (25): 3317–3321.
  5. Cooks, R. G., Zhang, D., Koch, K. J. "Chiroselective Self-Directed Octamerization of Serine: Implications for Homochirogenesis". Anal. Chem. 73 (15)): 3646–3655. doi:10.1021/ac010284l. Text " volume2001 " ignored (help)
  6. Nanita, S., Cooks, R. G. (2006). "Serine Octamers: Cluster Formation, Reactions, and Implications for Biomolecule Homochirality". Angew. Chem. Int. Ed. 45 (4): 554–569. doi:10.1002/anie.200501328.
  7. Donna G. Blackmond and Martin Klussmann (2007). "Spoilt for choice: assessing phase behaviour models for the evolution of homochirality". Chem. Commun.: xxxx–xxxx. doi:10.1039/b709314b.
  8. Stephen P. Fletcher, Richard B. C. Jagt and Ben L. Feringa (2007). "An astrophysically relevant mechanism for amino acid enantiomer enrichment". Chem. Commun. 2007: 2578–2580. doi:10.1039/b702882b.
  9. 9.0 9.1 Armando Córdova, Magnus Engqvist, Ismail Ibrahem, Jesús Casas, Henrik Sundén (2005). "Plausible origins of homochirality in the amino acid catalyzed neogenesis of carbohydrates". Chem. Commun. 15: 2047–2049.
  10. 10.0 10.1 S. Kojo, H. Uchino, M. Yoshimura, and K. Tanaka (2004). "Racemic D,L-asparagine causes enantiomeric excess of other coexisting racemic D,L-amino acids during recrystallization: a hypothesis accounting for the origin of L-amino acids in the biosphere". Chem. Comm.: 2146–2147. doi:10.1039/b409941a.
  11. Stereochemistry David G. Morris, Cambridge : Royal Society of Chemistry, 2001, p30.

External links

  • On the Genesis of Homochirality A. Maureen Rouhi Chemical & Engineering News June 17, 2004 Link
  • Observations Support Homochirality Theory Photonics TechnologyWorld November 1998 Link
  • Scienceweek digest 1998 Link
  • How left-handed amino acids got ahead: a demonstration of the evolution of biological homochirality in the lab Press release Imperial College London 2004 Link

Template:WikiDoc Sources