Homochirality

Last updated

Homochirality is a uniformity of chirality, or handedness. Objects are chiral when they cannot be superposed on their mirror images. For example, the left and right hands of a human are approximately mirror images of each other but are not their own mirror images, so they are chiral. In biology, 19 of the 20 natural amino acids are homochiral, being L-chiral (left-handed), while sugars are D-chiral (right-handed). [1] Homochirality can also refer to enantiopure substances in which all the constituents are the same enantiomer (a right-handed or left-handed version of an atom or molecule), but some sources discourage this use of the term. In fact Professor Eliel, Wilen, and Gal expressed their worry over the misuse of the term "homochiral" in scientific publications to designate enantiomerically pure drugs, which is inappropriate. [2] [3] [4]

Contents

It is unclear whether homochirality has a purpose; however, it appears to be a form of information storage. [5] One suggestion is that it reduces entropy barriers in the formation of large organized molecules. [6] It has been experimentally verified that amino acids form large aggregates in larger abundance from an enantiopure samples of the amino acid than from racemic (enantiomerically mixed) ones. [6]

It is not clear whether homochirality emerged before or after life, and many mechanisms for its origin have been proposed. [7] Some of these models propose three distinct steps: mirror-symmetry breaking creates a minute enantiomeric imbalance, chiral amplification builds on this imbalance, and chiral transmission is the transfer of chirality from one set of molecules to another.

In biology

Amino acids are the building blocks of peptides and enzymes while sugar-peptide chains are the backbone of RNA and DNA. [8] [9] In biological organisms, amino acids appear almost exclusively in the left-handed form (L-amino acids) and sugars in the right-handed form (R-sugars). [10] [ verification needed ] Since the enzymes catalyze reactions, they enforce homochirality on a great variety of other chemicals, including hormones, toxins, fragrances and food flavors. [11] :493–494 Glycine is achiral, as are some other non-proteinogenic amino acids that are either achiral (such as dimethylglycine) or of the D enantiomeric form.

Biological organisms easily discriminate between molecules with different chiralities. This can affect physiological reactions such as smell and taste. Carvone, a terpenoid found in essential oils, smells like mint in its L-form and caraway in its R-form. [11] :494[ verification needed ] Limonene tastes like citrus when right-handed and pine when left-handed. [12] :168

Homochirality also affects the response to drugs. Thalidomide, in its left-handed form, cures morning sickness; in its right-handed form, it causes birth defects. [12] :168 Unfortunately, even if a pure left-handed version is administered, some of it can convert to the right-handed form in the patient. [13] Many drugs are available as both a racemic mixture (equal amounts of both chiralities) and an enantiopure drug (only one chirality). Depending on the manufacturing process, enantiopure forms can be more expensive to produce than stereochemical mixtures. [12] :168

Chiral preferences can also be found at a macroscopic level. Snail shells can be right-turning or left-turning helices, but one form or the other is strongly preferred in a given species. In the edible snail Helix pomatia , only one out of 20,000 is left-helical. [14] :61–62 The coiling of plants can have a preferred chirality and even the chewing motion of cows has a 10% excess in one direction. [15]

Origins

Unsolved problem in chemistry:

What is the origin of homochirality in living organisms?

Symmetry breaking

Theories for the origin of homochirality in the molecules of life can be classified as deterministic or based on chance depending on their proposed mechanism. If there is a relationship between cause and effect — that is, a specific chiral field or influence causing the mirror symmetry breaking — the theory is classified as deterministic; otherwise it is classified as a theory based on chance (in the sense of randomness) mechanisms. [16]

Another classification for the different theories of the origin of biological homochirality could be made depending on whether life emerged before the enantiodiscrimination step (biotic theories) or afterwards (abiotic theories). Biotic theories claim that homochirality is simply a result of the natural autoamplification process of life—that either the formation of life as preferring one chirality or the other was a chance rare event which happened to occur with the chiralities we observe, or that all chiralities of life emerged rapidly but due to catastrophic events and strong competition, the other unobserved chiral preferences were wiped out by the preponderance and metabolic, enantiomeric enrichment from the 'winning' chirality choices.[ citation needed ] If this was the case, remains of the extinct chirality sign should be found. Since this is not the case, nowadays biotic theories are no longer supported.

The emergence of chirality consensus as a natural autoamplification process has also been associated with the 2nd law of thermodynamics. [17]

Deterministic theories

Deterministic theories can be divided into two subgroups: if the initial chiral influence took place in a specific space or time location (averaging zero over large enough areas of observation or periods of time), the theory is classified as local deterministic; if the chiral influence is permanent at the time the chiral selection occurred, then it is classified as universal deterministic. The classification groups for local determinist theories and theories based on chance mechanisms can overlap. Even if an external chiral influence produced the initial chiral imbalance in a deterministic way, the outcome sign could be random since the external chiral influence has its enantiomeric counterpart elsewhere.

In deterministic theories, the enantiomeric imbalance is created due to an external chiral field or influence, and the ultimate sign imprinted in biomolecules will be due to it. Deterministic mechanisms for the production of non-racemic mixtures from racemic starting materials include: asymmetric physical laws, such as the electroweak interaction (via cosmic rays [18] ) or asymmetric environments, such as those caused by circularly polarized light, quartz crystals, or the Earth's rotation, β-Radiolysis or the magnetochiral effect. [19] [20] The most accepted universal deterministic theory is the electroweak interaction. Once established, chirality would be selected for. [21]

One supposition is that the discovery of an enantiomeric imbalance in molecules in the Murchison meteorite supports an extraterrestrial origin of homochirality: there is evidence for the existence of circularly polarized light originating from Mie scattering on aligned interstellar dust particles which may trigger the formation of an enantiomeric excess within chiral material in space. [14] :123–124 Interstellar and near-stellar magnetic fields can align dust particles in this fashion. [22] Another speculation (the Vester-Ulbricht hypothesis) suggests that fundamental chirality of physical processes such as that of the beta decay (see Parity violation) leads to slightly different half-lives of biologically relevant molecules.

Chance theories

Chance theories are based on the assumption that "Absolute asymmetric synthesis, i.e., the formation of enantiomerically enriched products from achiral precursors without the intervention of chiral chemical reagents or catalysts, is in practice unavoidable on statistical grounds alone". [23]

Consider the racemic state as a macroscopic property described by a binomial distribution; the experiment of tossing a coin, where the two possible outcomes are the two enantiomers is a good analogy. The discrete probability distribution of obtaining n successes out of Bernoulli trials, where the result of each Bernoulli trial occurs with probability and the opposite occurs with probability is given by:

.

The discrete probability distribution of having exactly molecules of one chirality and of the other, is given by:

.

As in the experiment of tossing a coin, in this case, we assume both events ( or ) to be equiprobable, . The probability of having exactly the same amount of both enantiomers is inversely proportional to the square root of the total number of molecules . For one mol of a racemic compound, molecules, this probability becomes . The probability of finding the racemic state is so small that we can consider it negligible.

In this scenario, there is a need to amplify the initial stochastic enantiomeric excess through any efficient mechanism of amplification. [7] The most likely path for this amplification step is by asymmetric autocatalysis. An autocatalytic chemical reaction is that in which the reaction product is itself a reactive, in other words, a chemical reaction is autocatalytic if the reaction product is itself the catalyst of the reaction. In asymmetric autocatalysis, the catalyst is a chiral molecule, which means that a chiral molecule is catalysing its own production. An initial enantiomeric excess, such as can be produced by polarized light, then allows the more abundant enantiomer to outcompete the other.

Amplification

Theory

Phase portrait of Frank's model: starting from almost everywhere in L-D plane (except L = D line), the system approaches to one of the homochiral states (L=0 or D=0). Frank's model phase portrait.jpg
Phase portrait of Frank's model: starting from almost everywhere in L-D plane (except L = D line), the system approaches to one of the homochiral states (L=0 or D=0).

In 1953, Charles Frank proposed a model to demonstrate that homochirality is a consequence of autocatalysis. [24] [25] In his model the L and D enantiomers of a chiral molecule are autocatalytically produced from an achiral molecule A

while suppressing each other through a reaction that he called mutual antagonism

In this model the racemic state is unstable in the sense that the slightest enantiomeric excess will be amplified to a completely homochiral state. This can be shown by computing the reaction rates from the law of mass action:

where is the rate constant for the autocatalytic reactions, is the rate constant for mutual antagonism reaction, and the concentration of A is kept constant for simplicity.

The analytical solutions for are found to be . The ratio increases at a more than exponential rate if is positive (and vice versa). Every starting conditions different to

lead to one of the asymptotes or . Thus the equality of and and so of and represents a condition of unstable equilibrium, this result depending on the presence of the term representing mutual antagonism.

By defining the enantiomeric excess as

the rate of change of enantiomeric excess can be calculated using chain rule from the rate of change of the concentrations of enantiomers L and D.

Linear stability analysis of this equation shows that the racemic state is unstable. Starting from almost everywhere in the concentration space, the system evolves to a homochiral state.

It is generally understood that autocatalysis alone does not yield to homochirality, and the presence of the mutually antagonistic relationship between the two enantiomers is necessary for the instability of the racemic mixture. However, recent studies show that homochirality could be achieved from autocatalysis in the absence of the mutually antagonistic relationship, but the underlying mechanism for symmetry-breaking is different. [7] [26]

Experiments

There are several laboratory experiments that demonstrate how a small amount of one enantiomer at the start of a reaction can lead to a large excess of a single enantiomer as the product. For example, the Soai reaction is autocatalytic. [27] [28] If the reaction is started with some of one of the product enantiomers already present, the product acts as an enantioselective catalyst for production of more of that same enantiomer. [29] The initial presence of just 0.2 equivalent one enantiomer can lead to up to 93% enantiomeric excess of the product.

Soai autocatalysis.png

Another study [30] concerns the proline catalyzed aminoxylation of propionaldehyde by nitrosobenzene. In this system, a small enantiomeric excess of catalyst leads to a large enantiomeric excess of product.

Homochiralproline.png

Serine octamer clusters [31] [32] are also contenders. These clusters of 8 serine molecules appear in mass spectrometry 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. [33] 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. [34] 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 launch 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 [35]

One classic study involves an experiment that takes place in the laboratory. [36] 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.

In a related experiment, a crystal suspension of a racemic amino acid derivative continuously stirred, results in a 100% crystal phase of one of the enantiomers because the enantiomeric pair is able to equilibrate in solution (compare with dynamic kinetic resolution). [37]

Transmission

Once a significant enantiomeric enrichment has been produced in a system, the transference of chirality through the entire system is customary. This last step is known as the chiral transmission step. 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.

Some proposed models for the transmission of chiral asymmetry are polymerization, [38] [39] [40] [41] [42] [43] epimerization [44] [45] or copolymerization. [46] [47]

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 larger 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 [48] 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 = 100 ×(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. [48] 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 of term

This term was introduced by Kelvin in 1904, the year that he published his Baltimore Lecture of 1884. Kelvin used the term homochirality as a relationship between two molecules, i.e. two molecules are homochiral if they have the same chirality. [35] [49] Recently, however, homochiral has been used in the same sense as enantiomerically pure. This is permitted in some journals (but not encouraged), [50] :342 [51] 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

Related Research Articles

In chemistry, a racemic mixture or racemate, is one that has equal amounts of left- and right-handed enantiomers of a chiral molecule or salt. Racemic mixtures are rare in nature, but many compounds are produced industrially as racemates.

<span class="mw-page-title-main">Enantiomer</span> Stereoisomers which are non-superposable mirror images of each other

In chemistry, an enantiomer – also called optical isomer, antipode, or optical antipode – is one of two stereoisomers that are non-superposable onto their own mirror image. Enantiomers are much like one's right and left hands; without mirroring one of them, hands cannot be superposed onto each other. No amount of reorientation in three spatial dimensions will allow the four unique groups on the chiral carbon to line up exactly. The number of stereoisomers a molecule has can be determined by the number of chiral carbons it has. Stereoisomers include both enantiomers and diastereomers.

A chemical reaction is said to be autocatalytic if one of the reaction products is also a catalyst for the same reaction. Many forms of autocatalysis are recognized.

<span class="mw-page-title-main">Chirality (chemistry)</span> Geometric property of some molecules and ions

In chemistry, a molecule or ion is called chiral if it cannot be superposed on its mirror image by any combination of rotations, translations, and some conformational changes. This geometric property is called chirality. The terms are derived from Ancient Greek χείρ (cheir) 'hand'; which is the canonical example of an object with this property.

<span class="mw-page-title-main">Enantioselective synthesis</span> Chemical reaction(s) which favor one chiral isomer over another

Enantioselective synthesis, also called asymmetric synthesis, is a form of chemical synthesis. It is defined by IUPAC as "a chemical reaction in which one or more new elements of chirality are formed in a substrate molecule and which produces the stereoisomeric products in unequal amounts."

<span class="mw-page-title-main">Carbonaceous chondrite</span> Class of chondritic meteorites

Carbonaceous chondrites or C chondrites are a class of chondritic meteorites comprising at least 8 known groups and many ungrouped meteorites. They include some of the most primitive known meteorites. The C chondrites represent only a small proportion (4.6%) of meteorite falls.

In stereochemistry, enantiomeric excess (ee) is a measurement of purity used for chiral substances. It reflects the degree to which a sample contains one enantiomer in greater amounts than the other. A racemic mixture has an ee of 0%, while a single completely pure enantiomer has an ee of 100%. A sample with 70% of one enantiomer and 30% of the other has an ee of 40%.

<span class="mw-page-title-main">Biocatalysis</span> Use of natural catalysts to perform chemical transformations

Biocatalysis refers to the use of living (biological) systems or their parts to speed up (catalyze) chemical reactions. In biocatalytic processes, natural catalysts, such as enzymes, perform chemical transformations on organic compounds. Both enzymes that have been more or less isolated and enzymes still residing inside living cells are employed for this task. Modern biotechnology, specifically directed evolution, has made the production of modified or non-natural enzymes possible. This has enabled the development of enzymes that can catalyze novel small molecule transformations that may be difficult or impossible using classical synthetic organic chemistry. Utilizing natural or modified enzymes to perform organic synthesis is termed chemoenzymatic synthesis; the reactions performed by the enzyme are classified as chemoenzymatic reactions.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

<span class="mw-page-title-main">Ronald Breslow</span> American chemist

Ronald Charles David Breslow was an American chemist from Rahway, New Jersey. He was University Professor at Columbia University, where he was based in the Department of Chemistry and affiliated with the Departments of Biological Sciences and Pharmacology; he had also been on the faculty of its Department of Chemical Engineering. He had taught at Columbia since 1956 and was a former chair of the university's chemistry department.

<span class="mw-page-title-main">Cryptochirality</span>

In stereochemistry, cryptochirality is a special case of chirality in which a molecule is chiral but its specific rotation is non-measurable. The underlying reason for the lack of rotation is the specific electronic properties of the molecule. The term was introduced by Kurt Mislow in 1977.

In organic chemistry, kinetic resolution is a means of differentiating two enantiomers in a racemic mixture. In kinetic resolution, two enantiomers react with different reaction rates in a chemical reaction with a chiral catalyst or reagent, resulting in an enantioenriched sample of the less reactive enantiomer. As opposed to chiral resolution, kinetic resolution does not rely on different physical properties of diastereomeric products, but rather on the different chemical properties of the racemic starting materials. The enantiomeric excess (ee) of the unreacted starting material continually rises as more product is formed, reaching 100% just before full completion of the reaction. Kinetic resolution relies upon differences in reactivity between enantiomers or enantiomeric complexes.

<span class="mw-page-title-main">Chiral derivatizing agent</span> Reagent for converting a chemical compound to a chiral derivative

In analytical chemistry, A chiral derivatizing agent (CDA), also known as a chiral resolving reagent, is a derivatization reagent that is a chiral auxiliary used to convert a mixture of enantiomers into diastereomers in order to analyze the quantities of each enantiomer present and determine the optical purity of a sample. Analysis can be conducted by spectroscopy or by chromatography. Some analytical techniques such as HPLC and NMR, in their most commons forms, cannot distinguish enantiomers within a sample, but can distinguish diastereomers. Therefore, converting a mixture of enantiomers to a corresponding mixture of diastereomers can allow analysis. The use of chiral derivatizing agents has declined with the popularization of chiral HPLC. Besides analysis, chiral derivatization is also used for chiral resolution, the actual physical separation of the enantiomers.

<span class="mw-page-title-main">Soai reaction</span>

In organic chemistry, the Soai reaction is the alkylation of pyrimidine-5-carbaldehyde with diisopropylzinc. The reaction is autocatalytic and leads to rapidly increasing amounts of the same enantiomer of the product. The product pyrimidyl alcohol is chiral and induces that same chirality in further catalytic cycles. Starting with a low enantiomeric excess ("ee") produces a product with very high enantiomeric excess. The reaction has been studied for clues about the origin of homochirality among certain classes of biomolecules.

Spontaneous absolute asymmetric synthesis is a chemical phenomenon that stochastically generates chirality based on autocatalysis and small fluctuations in the ratio of enantiomers present in a racemic mixture. In certain reactions which initially do not contain chiral information, stochastically distributed enantiomeric excess can be observed. The phenomenon is different from chiral amplification, where enantiomeric excess is present from the beginning and not stochastically distributed. Hence, when the experiment is repeated many times, the average enantiomeric excess approaches 0%. The phenomenon has important implications concerning the origin of homochirality in nature.

In enantioselective synthesis, a non-linear effect refers to a process in which the enantiopurity of the catalyst or chiral auxiliary does not correspond with the enantiopurity of the product produced. For example: a racemic catalyst would be expected to convert a prochiral substrate into a racemic product, but this is not always the case and a chirally enriched product can be produced instead.

Donna Blackmond is an American chemical engineer and the John C. Martin Endowed Chair in Chemistry at Scripps Research in La Jolla, CA. Her research focuses on prebiotic chemistry, the origin of biological homochirality, and kinetics and mechanisms of asymmetric catalytic reactions. Notable works include the development of Reaction Progress Kinetic Analysis (RPKA), analysis of non-linear effects of catalyst enantiopurity, biological homochirality and amino acid behavior.

In homogeneous catalysis, C2-symmetric ligands refer to ligands that lack mirror symmetry but have C2 symmetry. Such ligands are usually bidentate and are valuable in catalysis. The C2 symmetry of ligands limits the number of possible reaction pathways and thereby increases enantioselectivity, relative to asymmetrical analogues. C2-symmetric ligands are a subset of chiral ligands. Chiral ligands, including C2-symmetric ligands, combine with metals or other groups to form chiral catalysts. These catalysts engage in enantioselective chemical synthesis, in which chirality in the catalyst yields chirality in the reaction product.

Viedma ripening or attrition-enhanced deracemization is a chiral symmetry breaking phenomenon observed in solid/liquid mixtures of enantiomorphous crystals that are subjected to comminution. It can be classified in the wider area of spontaneous symmetry breaking phenomena observed in chemistry and physics.

Chemical compounds that come as mirror-image pairs are referred to by chemists as chiral or handed molecules. Each twin is called an enantiomer. Drugs that exhibit handedness are referred to as chiral drugs. Chiral drugs that are equimolar (1:1) mixture of enantiomers are called racemic drugs and these are obviously devoid of optical rotation. The most commonly encountered stereogenic unit, that confers chirality to drug molecules are stereogenic center. Stereogenic center can be due to the presence of tetrahedral tetra coordinate atoms (C,N,P) and pyramidal tricoordinate atoms (N,S). The word chiral describes the three-dimensional architecture of the molecule and does not reveal the stereochemical composition. Hence "chiral drug" does not say whether the drug is racemic, single enantiomer or some other combination of stereoisomers. To resolve this issue Joseph Gal introduced a new term called unichiral. Unichiral indicates that the stereochemical composition of a chiral drug is homogenous consisting of a single enantiomer.

References

  1. Nelson, Lehninger; et al. (2008). Lehninger Principles of Biochemistry. Macmillan. p. 474.
  2. Ernest, L, Eliel; Samuel H, Wilen (1990). "Misuse of homochiral". Chemical & Engineering News (10): 2.
  3. Eliel, Ernest L. (1997). "Infelicitous stereochemical nomenclature". Chirality. 9 (5–6): 428–430. doi:10.1002/(SICI)1520-636X(1997)9:5/6<428::AID-CHIR5>3.0.CO;2-1. ISSN   0899-0042.
  4. Gal, Joseph (2013), Schurig, Volker (ed.), "Molecular Chirality: Language, History, and Significance", Differentiation of Enantiomers I, Cham: Springer International Publishing, vol. 340, pp. 1–20, doi:10.1007/128_2013_435, ISBN   978-3-319-03238-2, PMID   23666078 , retrieved 2023-11-30
  5. Carroll, James D. (March 2009). "A new definition of life". Chirality. 21 (3): 354–358. doi:10.1002/chir.20590. PMID   18571800.
  6. 1 2 Julian, Ryan R.; Myung, Sunnie; Clemmer, David E. (January 2005). "Do Homochiral Aggregates Have an Entropic Advantage?". The Journal of Physical Chemistry B. 109 (1): 440–444. doi:10.1021/jp046478x. PMID   16851034. S2CID   10599051.
  7. 1 2 3 Jafarpour, Farshid; Biancalani, Tommaso; Goldenfeld, Nigel (2017). "Noise-induced symmetry breaking far from equilibrium and the emergence of biological homochirality". Physical Review E. 95 (3): 032407. Bibcode:2017PhRvE..95c2407J. doi: 10.1103/PhysRevE.95.032407 . PMID   28415353.
  8. Reusch, William. "Peptides & Proteins". Natural Products. Michigan State University. Retrieved 8 May 2018.
  9. Lam, Eric (1997). "Nucleic acids and proteins". In Dey, P.M.; Harborne, J.B. (eds.). Plant Biochemistry. Burlington: Elsevier. p. 315. ISBN   9780080525723.
  10. Zubay, Geoffrey (2000). Origins of Life: On Earth and in the Cosmos. Elsevier. p. 96. ISBN   9780080497617.
  11. 1 2 Seckbach, Joseph, ed. (2012). Genesis - in the beginning : precursors of life, chemical models and early biological evolution. Dordrecht: Springer. ISBN   9789400729407.
  12. 1 2 3 Hazen, Robert M. (2007). Genesis : the scientific quest for life's origins. Washington, D.C.: Joseph Henry. ISBN   9780309103107.
  13. Smith, Silas (July 2009). "Chiral Toxicology: It's the Same Thing... Only Different". Toxicological Sciences. 110 (1): 4–30. doi: 10.1093/toxsci/kfp097 . PMID   19414517.
  14. 1 2 Meierhenrich, Uwe (2008). Amino acids and the asymmetry of life caught in the act of formation. Berlin: Springer. ISBN   9783540768869.
  15. Shaw, Andrew M. (2007). Astrochemistry From Astronomy to Astrobiology. Chichester: John Wiley & Sons. p. 247. ISBN   9780470091388.
  16. Guijarro, A. and Yus, M. The Origin of Chirality in the Molecules of Life (RSC Publishing, Cambridge, 2009), 1st ed.
  17. Jaakkola, S., Sharma, V. and Annila, A. (2008). "Cause of chirality consensus". Curr. Chem. Biol. 2 (2): 53–58. arXiv: 0906.0254 . doi:10.2174/187231308784220536. S2CID   8294807.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  18. Globus, Noemie; Blandford, Roger D. (20 May 2020). "The Chiral Puzzle of Life". The Astrophysical Journal Letters. 895 (1): L11. arXiv: 2002.12138 . Bibcode:2020ApJ...895L..11G. doi: 10.3847/2041-8213/ab8dc6 . S2CID   211532577.
  19. Barron, L. D. (1986-09-01). "True and false chirality and absolute asymmetric synthesis". Journal of the American Chemical Society. 108 (18): 5539–5542. doi:10.1021/ja00278a029. ISSN   0002-7863.
  20. Barron, L. D. (1981-08-20). "Optical activity and time reversal". Molecular Physics. 43 (6): 1395–1406. Bibcode:1981MolPh..43.1395B. doi:10.1080/00268978100102151. ISSN   0026-8976.
  21. Clark, Stuart (July–August 1999). "Polarized Starlight and the Handedness of Life". American Scientist . 87 (4): 336. Bibcode:1999AmSci..87..336C. doi:10.1511/1999.4.336. ISSN   0003-0996. S2CID   221585816.
  22. Helman, Daniel S (6 July 2018). "Galactic Distribution of Chirality Sources of Organic Molecules". Acta Astronautica. 151: 595–602. arXiv: 1612.06720 . Bibcode:2018AcAau.151..595H. doi:10.1016/j.actaastro.2018.07.008. ISSN   0094-5765. S2CID   10024470.
  23. Mislow, Kurt (2003). "Absolute Asymmetric Synthesis: A Commentary". Collection of Czechoslovak Chemical Communications. 68 (5): 849–864. doi:10.1135/cccc20030849. ISSN   1212-6950.
  24. Frank, F.C. (1953). "On spontaneous asymmetric synthesis". Biochimica et Biophysica Acta. 11 (4): 459–463. doi:10.1016/0006-3002(53)90082-1. PMID   13105666.
  25. Note that on his original paper, Frank did not propose any set of chemical reactions but a set of dynamical equations, where the concentrations of both enantiomers were denoted as [n1] and [n2] respectively.
  26. Jafarpour, Farshid; Biancalani, Tommaso; Goldenfeld, Nigel (2015). "Noise-induced mechanism for biological homochirality of early life self-replicators". Physical Review Letters. 115 (15): 158101. arXiv: 1507.00044 . Bibcode:2015PhRvL.115o8101J. doi:10.1103/PhysRevLett.115.158101. PMID   26550754. S2CID   9775791.
  27. Shibata, Takanori; Morioka, Hiroshi; Hayase, Tadakatsu; et al. (17 January 1996). "Highly Enantioselective Catalytic Asymmetric Automultiplication of Chiral Pyrimidyl Alcohol". Journal of the American Chemical Society. 118 (2): 471–472. doi:10.1021/ja953066g. ISSN   0002-7863.
  28. Soai, Kenso; Sato, Itaru; Shibata, Takanori (2001). "Asymmetric autocatalysis and the origin of chiral homogeneity in organic compounds". The Chemical Record. 1 (4): 321–332. doi:10.1002/tcr.1017. ISSN   1528-0691. PMID   11893072.
  29. Takanori Shibata; Hiroshi Morioka; Tadakatsu Hayase; Kaori Choji; Kenso Soai (1996). "Highly Enantioselective Catalytic Asymmetric Automultiplication of Chiral Pyrimidyl Alcohol". J. Am. Chem. Soc. 118 (2): 471–472. doi:10.1021/ja953066g.
  30. Suju P. Mathew, Hiroshi Iwamura and Donna G. Blackmond (21 June 2004). "Amplification of Enantiomeric Excess in a Proline-Mediated Reaction". Angewandte Chemie International Edition . 43 (25): 3317–3321. doi: 10.1002/anie.200453997 . PMID   15213963.
  31. Cooks, R. G., Zhang, D., Koch, K. J. (2001). "Chiroselective Self-Directed Octamerization of Serine: Implications for Homochirogenesis". Anal. Chem. 73 (15): 3646–3655. doi:10.1021/ac010284l. PMID   11510829. S2CID   27891319.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  32. 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. PMID   16404754.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  33. Donna G. Blackmond; Martin Klussmann (2007). "Spoilt for choice: assessing phase behaviour models for the evolution of homochirality". Chem. Commun. (39): 3990–3996. doi:10.1039/b709314b. PMID   17912393.
  34. Stephen P. Fletcher; Richard B. C. Jagt; Ben L. Feringa (2007). "An astrophysically relevant mechanism for amino acid enantiomer enrichment". Chem. Commun. 2007 (25): 2578–2580. doi:10.1039/b702882b. PMID   17579743.
  35. 1 2 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 (15): 2047–2049. doi:10.1039/b500589b. PMID   15834501.
  36. Kondepudi, D. K., Kaufman, R. J. & Singh, N. (1990). "Chiral Symmetry Breaking in Sodium Chlorate Crystallization". Science . 250 (4983): 975–976. Bibcode:1990Sci...250..975K. doi:10.1126/science.250.4983.975. PMID   17746924. S2CID   41866132.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  37. Noorduin, Wim L.; Izumi, Toshiko; Millemaggi, Alessia; Leeman, Michel; Meekes, Hugo; Van Enckevort, Willem J. P.; Kellogg, Richard M.; Kaptein, Bernard; Vlieg, Elias; Blackmond, Donna G. (January 2008). "Emergence of a Single Solid Chiral State from a Nearly Racemic Amino Acid Derivative" (PDF). Journal of the American Chemical Society. 130 (4): 1158–1159. doi:10.1021/ja7106349. PMID   18173274.
  38. Sandars, P. G. H. (2003). "A toy model for the generation of homochirality during polymerization". Origins of Life and Evolution of the Biosphere. 33 (6): 575–587. Bibcode:2003OLEB...33..575S. doi:10.1023/a:1025705401769. ISSN   0169-6149. PMID   14601927. S2CID   25241450.
  39. Brandenburg, Axel; Multamäki, Tuomas (July 2004). "How long can left and right handed life forms coexist?". International Journal of Astrobiology. 3 (3): 209–219. arXiv: q-bio/0407008 . Bibcode:2004IJAsB...3..209B. doi:10.1017/s1473550404001983. ISSN   1473-5504. S2CID   16991953.
  40. Brandenburg, A.; Andersen, A. C.; Höfner, S.; Nilsson, M. (June 2005). "Homochiral Growth Through Enantiomeric Cross-Inhibition". Origins of Life and Evolution of Biospheres. 35 (3): 225–241. arXiv: q-bio/0401036 . Bibcode:2005OLEB...35..225B. doi:10.1007/s11084-005-0656-9. ISSN   0169-6149. PMID   16228640. S2CID   16833396.
  41. Wattis, Jonathan A. D.; Coveney, Peter V. (June 2005). "Symmetry-breaking in Chiral Polymerisation". Origins of Life and Evolution of Biospheres. 35 (3): 243–273. arXiv: physics/0402091 . Bibcode:2005OLEB...35..243W. doi:10.1007/s11084-005-0658-7. ISSN   0169-6149. PMID   16228641. S2CID   12451904.
  42. Saito, Yukio; Hyuga, Hiroyuki (2005-05-15). "Chirality Selection in Open Flow Systems and in Polymerization". Journal of the Physical Society of Japan. 74 (5): 1629–1635. arXiv: physics/0503057 . Bibcode:2005JPSJ...74.1629S. doi:10.1143/jpsj.74.1629. ISSN   0031-9015. S2CID   18419335.
  43. Blanco, Celia; Hochberg, David (2011). "Chiral polymerization: symmetry breaking and entropy production in closed systems". Phys. Chem. Chem. Phys. 13 (3): 839–849. arXiv: 1104.2225 . Bibcode:2011PCCP...13..839B. doi:10.1039/c0cp00992j. ISSN   1463-9076. PMID   21057681. S2CID   516456.
  44. Plasson, R.; Bersini, H.; Commeyras, A. (2004-11-17). "Recycling Frank: Spontaneous emergence of homochirality in noncatalytic systems". Proceedings of the National Academy of Sciences. 101 (48): 16733–16738. Bibcode:2004PNAS..10116733P. doi: 10.1073/pnas.0405293101 . ISSN   0027-8424. PMC   534711 . PMID   15548617.
  45. Stich, Michael; Blanco, Celia; Hochberg, David (2013). "Chiral and chemical oscillations in a simple dimerization model". Phys. Chem. Chem. Phys. 15 (1): 255–261. arXiv: 1210.1872 . Bibcode:2013PCCP...15..255S. doi:10.1039/c2cp42620j. ISSN   1463-9076. PMID   23064600. S2CID   2655068.
  46. Wattis, Jonathan A. D.; Coveney, Peter V. (August 2007). "Sequence Selection during Copolymerization". The Journal of Physical Chemistry B. 111 (32): 9546–9562. doi:10.1021/jp071767h. ISSN   1520-6106. PMID   17658787.
  47. Blanco, Celia; Hochberg, David (2012). "Homochiral oligopeptides by chiral amplification: interpretation of experimental data with a copolymerization model". Physical Chemistry Chemical Physics. 14 (7): 2301–11. arXiv: 1202.2268 . Bibcode:2012PCCP...14.2301B. doi:10.1039/c2cp22813k. ISSN   1463-9076. PMID   22237639. S2CID   16960638.
  48. 1 2 S. Kojo; H. Uchino; M. Yoshimura; 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. (19): 2146–2147. doi:10.1039/b409941a. PMID   15467844.
  49. Morris, David G. (2001). Stereochemistry . Cambridge: Royal Society of Chemistry. p.  30. ISBN   978-1-84755-194-8.
  50. Anslyn, Eric V.; Dougherty, Dennis A. (2006). Modern physical organic chemistry. Sausalito, Calif.: University Science Books. ISBN   9781891389313.
  51. However, the message can be confusing. In Moss, G. P. (1 January 1996). "Basic terminology of stereochemistry (IUPAC Recommendations 1996)" (PDF). Pure and Applied Chemistry. 68 (12): 2193–2222. doi:10.1351/pac199668122193. S2CID   98272391 . Retrieved 7 May 2018., the entry for Enantiomerically Pure/Enantiopure says "Use of homochiral as a synonym is strongly discouraged"; but the entry for Homochiral says "See enantiomerically pure/enantiopure."

Further reading