Next Article in Journal
Spin-Boson Model as A Simulator of Non-Markovian Multiphoton Jaynes-Cummings Models
Next Article in Special Issue
Chemical Basis of Biological Homochirality during the Abiotic Evolution Stages on Earth
Previous Article in Journal
A Weighted Fourier and Wavelet-Like Shape Descriptor Based on IDSC for Object Recognition
Previous Article in Special Issue
A General Phenomenon of Spontaneous Amplification of Optical Purity under Achiral Chromatographic Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Asymmetric Autocatalysis in the Elucidation of Origins of Homochirality of Organic Compounds

1
Department of Applied Chemistry, Tokyo University of Science, Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
2
Department of Chemistry, Biology and Environmental Science, Nara Women’s University, Kita-Uoya Nishi-machi, Nara 630-8506, Japan
*
Author to whom correspondence should be addressed.
Symmetry 2019, 11(5), 694; https://doi.org/10.3390/sym11050694
Submission received: 29 April 2019 / Revised: 15 May 2019 / Accepted: 16 May 2019 / Published: 20 May 2019
(This article belongs to the Special Issue Possible Scenarios for Homochirality on Earth)

Abstract

:
Pyrimidyl alkanol and related compounds were found to be asymmetric autocatalysts in the enantioselective addition of diisopropylzinc to pyrimidine-5-carbaldehyde and related aldehydes. In the asymmetric autocatalysis with amplification of enantiomeric excess (ee), the very low ee (ca. 0.00005%) of 2-alkynyl-5-pyrimidyl alkanol was significantly amplified to >99.5% ee with an increase in the amount. By using asymmetric autocatalysis with amplification of ee, several origins of homochirality have been examined. Circularly polarized light, chiral quartz, and chiral crystals formed from achiral organic compounds such as glycine and carbon (13C/12C), nitrogen (15N/14N), oxygen (18O/16O), and hydrogen (D/H) chiral isotopomers were found to act as the origin of chirality in asymmetric autocatalysis. And the spontaneous absolute asymmetric synthesis was also realized without the intervention of any chiral factor.

1. Introduction

The origins of biological homochirality of l-amino acids and d-sugars have attracted considerable attention ever since Pasteur discovered molecular dissymmetry in 1848 [1]. Although several theories of the origins of homochirality of organic compounds have been proposed [2,3,4,5,6,7,8,9,10], the enantiomeric excesses induced by these have usually been very low. For organic compounds to achieve homochirality, an amplification process from low enantiomeric excess (ee) to very high ee is required [11,12,13,14,15,16,17,18,19,20,21,22,23]. Therefore, asymmetric autocatalysis with amplification of chirality has been envisaged as the efficient process. We describe the discovery of asymmetric autocatalysis with amplification of ee. We also describe the study on the elucidation of the origin of homochirality of organic compounds by using asymmetric autocatalysis [24,25,26,27,28,29,30,31,32,33,34,35,36].
Asymmetric autocatalysis involves a process where a chiral product serves as the catalyst for its own production (Scheme 1). The reaction is a catalytic self-replication, i.e., automultiplication of a chiral compound. The superiority of asymmetric autocatalysis over the conventional non-autocatalytic asymmetric catalysis is as follows: (1) Because of the process of self-replication, the efficiency is high. (2) During the reaction, the amount of catalyst increases as the product increases. The catalytic activity and amount of catalyst does not decrease. (3) Because the structure of the product and the catalyst is the same, the separation of product from catalyst is not necessary.
Frank proposed a mechanism, i.e., a mathematical equation, of asymmetric autocatalysis without showing any chemical structure in 1953 [21]. However, no real asymmetric autocatalysis had been reported until we first reported on the asymmetric autocatalysis of 3-pyridyl alkanol in 1990 [37].

2. Discovery of Asymmetric Autocatalysis with Amplification of Enantiomeric Excess

After the examination of the chiral diol system [38], we found in 1995 an efficient asymmetric autocatalysis of 5-pyrimidyl alkanol 1 with amplification of ee from 2% ee to 88% ee in the reaction between diisopropylzinc (i-Pr2Zn) and pyrimidine-5-carbaldehyde 2a (Scheme 2) [39,40]. In that reaction, pyrimidyl alkanol 1a with 2% ee serves as an asymmetric autocatalyst to produce more of itself with an amplified ee. The consecutive asymmetric autocatalysis enables the amplification from 2 to 88% ee [39]. 2-Alkynylpyrimidyl alkanol 1c with >99.5% ee was found to be an efficient asymmetric autocatalyst affording itself, 1c, with >99.5% ee and with >99% yield [41]. It was also found that the asymmetric autocatalysis of pyrimidyl alkanol 1c exhibit significant amplification of ee (Scheme 3). Indeed, starting from a very low (ca. 0.00005%) ee of (S)-pyrimidyl alkanol 1c as an asymmetric autocatalyst, three cycles of asymmetric autocatalysis enabled the amplification of ee of alkanol 1c to >99.5%. During the reaction, the amount of (S)-1c increased by a factor of ca. 630,000 times [42]. 2-Alkenylpyrimidyl alkanol 1e [43], 3-quinolyl alkanol 4 [44,45,46], and 5-carbamoylpyridyl alkanol 5 [47,48] are also highly enantioselective asymmetric autocatalysts with amplification of ee (Scheme 2). The unique aspect of amplification of ee by asymmetric autocatalysis is that it is accomplished without the intervention of any other chiral factor. The only chiral factor is the initial enantiomeric imbalance of alkanol 1 itself as an asymmetric autocatalyst. In addition, asymmetric autocatalytic self-multiplication of multi-functionalized pyrimidyl alkanol 3 [49] and ultra-remote intramolecular asymmetric autocatalysis [50] were reported.
Thus, it was proved that a chemical reaction exists in which very low enantioenrichment is amplified to almost enantiopure (>99.5% ee).

3. Study on the Mechanism of Asymmetric Autocatalysis

As described in the preceding section, asymmetric autocatalysis exhibits enormous amplification of ee during the self-replication. Thus, mechanistic insights into the asymmetric autocatalysis have attracted great attention. For the non-autocatalytic, non-linear effect in asymmetric catalysis, the dimer mechanism by Noyori [51] and MLn mechanism by Kagan [52] have been proposed.
We revealed the relationship between the reaction time and yield in the asymmetric autocatalysis using pyrimidyl alkanol 1c with >99.5% ee [53]. A sigmoidal curve of product formation was observed. We also reported the relationship between the time, yield, and ee of the product by using chiral HPLC [54], which suggested dimeric or higher order aggregated catalytic species.
Several groups also investigated the mechanism of asymmetric autocatalysis. Heat flow measurement by microcalorimeter revealed the relationship between a reaction rate and the progress of the reaction. This suggested the dimeric catalyst model [55]. The dimeric and tetrameric species were proposed by the NMR measurement of the reaction solution [56,57]. The structure of catalyst aggregates has been proposed by density functional theory (DFT) calculation [58,59,60,61]. Reaction models have also been presented based on spontaneous mirror-symmetry breakage. These works proposed the mechanistic frameworks of asymmetric autocatalysis of pyrimidyl alkanol [62,63,64,65,66,67,68,69]. We clarified the crystal structures of asymmetric autocatalyst 1c based on X-ray diffraction [70,71]. It was revealed that the structures are either tetrameric or oligomeric. The tetrameric crystal structure is formed in the presence of an excess molar amount of i-Pr2Zn, while the higher order aggregate is formed in the presence of an equimolar or slightly excess amount of i-Pr2Zn. Recently, reaction modeling was reported which suggests that the tetramer or higher order aggregates work for the asymmetric autocatalysis [72]. The clarification of the entire reaction pathway of asymmetric autocatalysis awaits further investigation.

4. Elucidation of the Origins of Homochirality by Using Asymmetric Autocatalysis

As described in the preceding section, asymmetric autocatalysis amplified ee from very low to very high. We then examined the origins of homochirality by using asymmetric autocatalysis. We envisaged that the low ee induced by the origin of chirality could be amplified by asymmetric autocatalysis. The origins of chirality so far proposed have usually induced only very low ees. To explain the very high ees observed in nature, the amplification of very low ee of organic compounds is necessary. We employed asymmetric autocatalysis of amplification of ee to examine the several proposed mechanisms of the origin of chirality.

4.1. Circularly Polarized Light

One of the representative chiral physical forces is circularly polarized light (CPL). Left (l) and right (r)-CPL have long been considered as the origin of chirality. In some of the star-forming regions, the occurrence of relatively strong CPL has been observed [73]. It is known that only ca. 2% ee is induced by irradiation of CPL to racemic organic compounds such as leucine. Asymmetric photosynthesis of hexa-helicen by CPL irradiation has been reported [5]. The induced low ee in leucine was correlated, for the first time, to the very high ee of organic compounds by using asymmetric autocatalysis [74].
The direct irradiation of l-CPL to racemic (rac) pyrimidyl alkanol 1c, and the subsequent asymmetric autocatalysis, gave (S)-alkanol 1c with >99.5% ee (Scheme 4) as a result of the amplification of ee [75]. On the other hand, r-CPL irradiation affords (R)-1c with >99.5% ee. The relationship between the handedness of l- and r-CPL and (S)-1c and (R)-1c is explained by the following consideration: The cotton effects of the circular dichroism (CD) spectra of the solid state of (R)-1 and (S)-1c are plus (+) and minus (-) at 313 nm, respectively. Thus, when l-CPL is irradiated on rac-1c, the asymmetric photodecomposition of (R)-1c is induced because l-CPL is absorbed preferentially. Then, the less reactive (S)-1c becomes the predominant enantiomer over (R)-1c. The asymmetric autocatalysis of the remaining alkanol increases the ee of (S)-1c to >99.5% ee. Thus, the direct correlation is accomplished between the handedness of CPL and that of highly enantioenriched organic compound.
The asymmetric photoequilibrium of rac-olefin 6 using CPL, and the subsequent asymmetric autocatalytic reaction, gave pyrimidyl alkanol 1c of the correlated absolute configuration to CPL [76]. Recently, under CPL irradiation, a Viedma-type racemization-crystallization of an amino acid derivative was reported [77].

4.2. Chiral Inorganic Crystals of Quartz, Sodium Chlorate, Cinnabar, and Retgersite, and the Enantiotopic Face of the Achiral Crystal of Gypsum

A chiral single crystal of silicon dioxide is known as quartz, and it exhibits enantiomorphism. Chiral minerals including quartz have been proposed as the origin of homochirality [6]. There are many reports attempting to induce chirality in organic compounds by using quartz [78]. However, no significant asymmetric induction has yet been reported by using quartz.
We thought that the asymmetric autocatalysis amplifies significantly the very low ee of the product initially induced by chiral d- and l-quartz [79]. Indeed, in the presence of d-quartz, asymmetric autocatalysis using pyrimidine-5-carbaldehyde 2c and i-Pr2Zn afforded (S)-1c with 97% ee in a yield of 95% (Scheme 5). On the other hand, l-quartz afforded (R)-1c with 97% ee. It was clearly shown by these results that d- and l-quartz act as chiral initiators of asymmetric autocatalysis. The initially formed slightly enriched (S)-(zinc alkoxide) of pyrimidyl alkanol 1c serves as an asymmetric autocatalyst and automultiplies with amplification of ee. Thus, the chirality of d- and l-quartz is correlated to the chirality of a near enantiopure organic compound.
Sodium chlorate (NaClO3) and sodium bromate (NaBrO3) are chiral inorganic ionic crystals [14,80,81]. It was also found that d-NaClO3 triggers asymmetric autocatalysis to give (S)-1c, while l-NaClO3 gives (R)-1c [82]. On the other hand, d-NaBrO3 and l-NaBrO3 trigger the formation of (R)- and (S)-1c, respectively [83]. Note that d-NaClO3 and l-NaBrO3 with the opposite signs of optical activity have the same type of enantiomorph. Enantiomorphic P- and M-crystals of cinnabar, mercury(II) sulfide (HgS), are composed of –Hg–S–Hg–S helical chains. We found that P-cinnabar acts as a chiral trigger of asymmetric autocatalysis to give (R)-1c. In contrast, M-HgS triggers the formation of (S)-1c [84]. Retgersite (NiSO4 6H2O) of [CD(+)390Nujol] triggers asymmetric autocatalysis to afford (S)-1c. In contrast, retgersite of [CD(−)390Nujol] affords (R)-1c [85].
Gypsum (calcium sulfate dihydrate) is a common mineral which has been widely used. The crystal structure is not chiral. However, gypsum exhibits two-dimensional enantiotopic cleavage (010) and (0−10) face. Pyrimidine-5-carbaldehyde 2c was put on the enantiotopic (010) face. Then, the reaction of aldehyde 2c on gypsum with the vapor of i-Pr2Zn gave (R)-pyrimidyl alkanol 1c [86]. In contrast, the reaction by exposing on the opposite (0−10) face gave (S)-alkanol 1c. Thus, it was shown that the enantiotopic face of achiral gypsum works as an origin of chirality.
In combination with asymmetric autocatalysis, chiral inorganic crystals serve as the origin of chirality to give enantioenriched organic compounds of the correlated absolute configurations.

4.3. Chiral Crystals Formed from Achiral Organic Compounds

Achiral organic compounds often form achiral crystals. However, it is known that some of the achiral organic compounds form chiral crystals [87]. In some stereospecific reactions, these chiral organic crystals have been used as reactants [10]. However, in enantioselective synthesis, chiral crystals composed of achiral organic compounds have seldom been used as inducers. We used chiral crystals formed from achiral organic compounds as chiral inducers of asymmetric autocatalysis (Scheme 6 and Scheme 7).
Natural proteinogenic amino acids, except glycine, exhibit l-form. Glycine stands as the only achiral amino acid that possesses no asymmetric carbon atoms. Although it is known that the stable crystal structure of the γ-glycine polymorph is chiral, it took years to determine the absolute crystal structure of the γ-glycine polymorph. Recently, the absolute crystal structure of the γ-glycine polymorph was correlated with optical rotatory dispersion (ORD) [88]. Guillemin reported CD spectra of γ-glycine [89].
We have correlated the absolute crystal structure of γ-glycine and have used the γ-glycine crystal as a chiral trigger of asymmetric autocatalysis [90]. It was found that the P32 crystal (left-handed) of γ-glycine triggers the formation of (S)-pyrimidyl alkanol 1c with up to >99.5% ee (Scheme 6). In contrast, the P31 crystal afforded (R)-alkanol 1c with up to >99.5% ee.
Thus, in conjunction with asymmetric autocatalysis, achiral glycine as its chiral γ-polymorph acts as the origin of homochirality.
Cytosine is a nucleobase and achiral. It may be formed under plausible prebiotic conditions [91]. When cytosine is crystallized from methanol, chiral crystals form. Chiral crystals of cytosine trigger asymmetric autocatalysis. When cytosine crystals of [CD(+)310Nujol] were used as chiral initiators of the reaction of aldehyde 2c with i-Pr2Zn, (R)-alkanol 1c was formed in combination with asymmetric autocatalysis (Scheme 7). In contrast, a [CD(−)310Nujol]-cytosine crystal afforded (S)-1c [92]. Thus, the chiral cytosine crystal serves as the origin of chirality.
Cytosine forms achiral crystals of cytosine monohydrate when it is crystallized from water. When it is heated from one of the enantiotopic faces, the crystal water is eliminated by heating and chiral dehydrated cytosine is formed [93]. Interestingly, the chirality of the dehydrated crystal is determined by the enantiotopic face of the crystal from which the heating is applied. It is worth noting that the dehydration of the crystal water of cytosine monohydrate under reduced pressure conditions [94] also gives the chiral cytosine crystal with the opposite chirality to that dehydrated by heating. Thus, by removal of crystal water from an achiral crystal of cytosine monohydrate either by heating or under reduced pressure, the formation of chiral crystals with controlled absolute chirality was achieved.
Adenine is another achiral nucleobase. Chiral crystals of adenine dinitrate act as chiral initiators of asymmetric autocatalysis (Scheme 7) [95]. Thus, achiral nucleobases, i.e., cytosine and adenine, can serve as the origin of homochirality in conjunction with asymmetric autocatalysis.
Enantiomorphous crystals formed from achiral N-benzoylglycine (hippuric acid) [96], 2-thenoylglycine [97], certain chiral cocrystals consisting of two achiral compounds [98], benzil [99], tetraphenylethylene [100], ethylenediammonium sulfate [101], aromatic triester [102], and 2,6-di-tert-butyl-p-cresol (BHT) [103] serve as chiral initiators of asymmetric autocatalysis (Scheme 7). It should be added that a chiral crystal composed of a racemic serine initiates asymmetric autocatalysis. Asymmetric autocatalysis using the M-crystals of dl-diserinium sulfate hydrate as the chiral initiator afford (R)-pyrimidyl alkanol 1c, while P-crystals afford (S)-alkanol 1c [104].

4.4. Enantiotopic Face of Achiral Organic Crystal Composed of Achiral Organic Compound

Some of the crystal faces of achiral organic crystals formed from achiral compounds become enantiotopic. Achiral 2-(tert-butyldimethylsilylethynyl) pyrimidine-5-carbaldehyde 2f forms an achiral crystal (P-1) that has enantiotopic faces. When the Re-face of the crystal was exposed to i-Pr2Zn, (R)-pyrimidyl alkanol, 1f was formed (Scheme 8) [105]. In contrast, exposure of i-Pr2Zn on the Si-face gave (S)-alkanol 1f. The ees of alkanol 1f were amplified to >99.5% ee by asymmetric autocatalysis. Thus, it was shown that the enantiotopic faces of achiral crystals act as the origin of homochirality in conjunction with asymmetric autocatalysis.

4.5. Spontaneous Absolute Asymmetric Synthesis by Asymmetric Autocatalysis

As described in the preceding section, asymmetric autocatalysis of pyrimidyl alkanol enhances extremely low ca. 0.00005% ee to near enantiopure >99.5% ee [42]. We reasoned that if i-Pr2Zn is reacted with pyrimidine-5-carbaldehyde 2 without using any chiral factor, the product with low ee based on the statistical fluctuation would form. The subsequent asymmetric autocatalysis may enhance the initial low ee to the detectable high ee (Scheme 9).
Although the term “absolute asymmetric synthesis” had been used for the asymmetric synthesis “without the use of any chiral chemical substance,” Mislow newly defined absolute asymmetric synthesis as “the formation of an enantioenriched compound from achiral compounds without the intervention of any chiral factor [3].” The spontaneous absolute asymmetric synthesis, based on the statistical fluctuation, has been thought of as one of the origins of chirality. However, it is known that the reaction between achiral reagents without any chiral factor always gives so-called racemic product. However, there are statistical fluctuations in the numbers of enantiomers [3]. Let us consider the situation of flipping a coin one hundred times: there is an 8% probability of 50 heads and 50 tails. The remaining 92% are results with either heads or tails being in excess: 49 to 51, 53 to 47, etc. Pályi et al. described the distribution of ee by statistical fluctuations of various amounts of so-called racemic molecules [106,107,108].
We found spontaneous absolute asymmetric synthesis in the reaction between pyrimidine-5-carbaldehyde 2 and i-Pr2Zn without the addition of any chiral substance. In 1996, we applied patent for this absolute asymmetric synthesis [109,110]. The reaction afforded enantioenriched (S)-pyrimidyl alkanol 1 or (R)-alkanol 1 [109]. When aldehyde 2c and i-Pr2Zn were reacted in a mixed solvent of ether-toluene, enantioenriched product was formed in situ by statistical fluctuation. The subsequent asymmetric autocatalysis gave (S) or (R)-1 with detectable enantioenrichments. The formation of (S)-alkanol 1c occurred 19 times and (R)-1c occurred 18 times in a total of 37 reactions (Figure 1a) [110]. The absolute configurations of 1c formed exhibits a stochastic distribution of S and R enantiomers. Moreover, by using achiral amorphous silica gel (Figure 1b) [111] and achiral amines (Figure 1c) [112], enantioenriched 1c was obtained and the distribution of (S)- and (R)-handedness was stochastic. The absolute asymmetric synthesis has also been reported between pyrimidine-5-carbaldehyde 2b and i-Pr2Zn (S)-1b or (R)-1b in a stochastic distribution [113]. As described, the results fulfill the conditions necessary for spontaneous absolute asymmetric synthesis [62,65,114,115,116,117].
Very recently, absolute asymmetric synthesis under heterogeneous solid-vapor phase conditions has been reported by us (Scheme 10) [118]. The powder of pyrimidine-5-carbaldehyde 2c in test tubes was exposed to the vapor of i-Pr2Zn and toluene in a desiccator. In 129 reactions, (R)-pyrimidyl alkanol 1c was formed 61 times. On the other hand, (S)-alkanol 1c was formed 58 times (10 times the formation of 1c of <0.5% ee was assigned as below the detection level). Thus, the results show that the distribution of (S) and (R)-alkanol 1c is stochastic. Although the ee values of alkanol 1c varied, these ee could be enhanced to >99.5% ee during the subsequent asymmetric autocatalysis. The present heterogeneous absolute asymmetric synthesis under solid vapor phase conditions could be possible in a more spacious platform.

4.6. Asymmetric Autocatalysis Triggered by Hydrogen, Carbon, Oxygen, and Nitrogen Chiral Isotopomers

Many apparent achiral organic compounds become chiral by substitution of carbon (12C), nitrogen (14N), and oxygen (16O) for their isotopes of 13C, 15N, and 18O, respectively. For example, dimethylphenylmethanol 8 is an achiral compound because it has the same two methyl groups. However, when one of the carbon atoms of the methyl group is labelled with 13C, the alkanol becomes a chiral (R)-alkanol 8(13C) or (S)-alkanol 8(13C) (Scheme 11). Because the difference of carbon (13C/12C) isotopomers between enantiomers is so small, no report has appeared before on the asymmetric induction by using chiral carbon (13C/12C) isotopomers.
We found that in the presence of chiral carbon (13C/12C) isotopomer and (R) or (S)-8(13C), as a chiral trigger, pyrimidine-5-carbaldehyde 2c reacts with i-Pr2Zn to give pyrimidyl alkanol 1c with a very high ee of the absolute configuration correlated to that of the carbon isotopomer (Scheme 11). (R)-Carbon isotopomer 8(13C) triggered the formation of (S)-pyrimidyl alkanol 1c with high ee. In contrast, (S)-carbon isotopomer 8(13C) gave (R)-pyrimidyl alkanol [119]. Other carbon (13C/12C) isotopomers also serve as chiral triggers on asymmetric autocatalysis. Chiral nitrogen (15N/14N) isotopomer, [15N](S) and [15N](R)-diamine 9(15N) were also found to work as chiral triggers of asymmetric autocatalysis [120]. In addition, oxygen (18O/16O) isotopomer, [18O](R), and [18O](S)-diol 10(18O), trigger asymmetric autocatalysis to give pyrimidyl alkanol 1c of high ee with the correlated absolute configuration to that of oxygen isotopomer [121,122]. As described, carbon, nitrogen, and oxygen isotopomers were found to act as the origin of homochirality in conjunction with asymmetric autocatalysis.
As to chiral hydrogen (D/H) isotopomers, there are a few examples of low asymmetric induction by hydrogen isotopomers [123,124]. It was found that chiral hydrogen isotopomers act as chiral initiators of asymmetric autocatalysis [125,126]. It should be noted that achiral glycine 7 becomes chiral by substituting one of the hydrogen atoms of the methylene group for deuterium (D). In the presence of chiral (S)-glycine-α-d 7(D), (S)-pyrimidyl alkanol 1c of high ee was formed with the correlated absolute configuration to that of chiral glycine-α-d [127].

5. Various Chiral Compounds as Triggers of Asymmetric Autocatalysis

Various chiral compounds work as chiral initiators of asymmetric autocatalysis. Amino acids even with low ee [128], such as hexa-helicene [129], tetrathia-hepta-helicene [130], and 2-aza-hexa-helicene [131], initiate asymmetric autocatalysis to give alkanol 1c of the correlated absolute configuration to those of the chiral initiators. It is known that the value of optical rotation of a chiral saturated quaternary hydrocarbon, 5-ethyl-5-propylundecane, is below detection level because the differences in the structures of the four substituents are so small. The compound is called cryptochiral. It was found that 5-ethyl-5-propylundecane triggers asymmetric autocatalysis [132]. Cryptochiral isotactic polystyrene also works as a chiral trigger [133]. Artificially designed helical [134] silica and mesoporous helical silica [135] are also chiral triggers.

6. Conclusions

Asymmetric autocatalysis of the enantioselective addition of i-Pr2Zn to pyrimidine-5-carbaldehyde was discovered by us. In this reaction, the very low ca. 0.00005% ee of (S)-2-alkynylpyrimidyl alkanol 1 was enhanced to >99.5% ee by consecutive asymmetric autocatalyses. Mislow first mentioned this reaction as the Soai reaction [3]. The asymmetric autocatalysis with amplification of ee is unique because no chiral substance other than the asymmetric autocatalyst itself is required.
To elucidate the origins of homochirality, asymmetric autocatalysis with amplification of ee was applied. By using asymmetric autocatalysis, the initially induced low ee by the proposed origin of chirality was enhanced significantly by the asymmetric autocatalysis. The racemic pyrimidyl alkanol was irradiated with l or r-circularly polarized light. The subsequent asymmetric autocatalysis correlated the chirality of CPL with that of the formed alkanol 1. Thus, for the first time, the correlation was made possible between the chirality of CPL and that of a chiral organic compound of very high ee. Chiral minerals such as quartz and cinnabar were found to act as chiral triggers of asymmetric autocatalysis. Thus, chirality of quartz was correlated to that of a highly enantioenriched organic compound. It was also found that chiral organic crystals composed of achiral compounds, i.e., glycine, cytosine, and adenine, serve as chiral triggers of asymmetric autocatalysis. Spontaneous absolute asymmetric synthesis without the intervention of any chiral factors was realized using the asymmetric autocatalysis of pyrimidyl alkanol with amplification of ee. Asymmetric autocatalysis was initiated by chiral compounds resulting from carbon (13C/12C), nitrogen (15N/14N), and oxygen (16O/18O) isotopomers. X-ray crystallographic analysis revealed the structure of asymmetric autocatalysts. It should be mentioned that bio-reactions should be studied looking for asymmetric autocatalysis.

Author Contributions

Conceptualization, K.S.; Writing-Original Draft Preparation, K.S. and T.K.; Writing—Review and Editing, K.S., T.K. and A.M.; Supervision, K.S.

Funding

KAKENHI: 19K05482 from Japan Society for the Promotion of Science.

Acknowledgments

The authors gratefully acknowledge their collaborators whose names appear in the literature cited.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pasteur, L. Recherches sur les relations qui peuvent exister entre la forme crystalline, la composition chimique et le sens de la polarisation rotatoire. Ann. Chim. Phys. 1848, 24, 442–459. [Google Scholar]
  2. Guijarro, A.; Yus, M. The Origin of Chirality in the Molecules of Life; The Royal Society of Chemistry: Cambridge, UK, 2009. [Google Scholar]
  3. Mislow, K. Absolute asymmetric synthesis: A commentary. Collect. Czech. Chem. Commun. 2003, 68, 849–864. [Google Scholar] [CrossRef]
  4. Feringa, B.L.; van Delden, R.A. Absolute asymmetric synthesis: The origin, control, and amplification of chirality. Angew. Chem. Int. Ed. 1999, 38, 3418–3438. [Google Scholar]
  5. Inoue, Y. Asymmetric photochemical reactions in solution. Chem. Rev. 1992, 92, 741–770. [Google Scholar] [CrossRef]
  6. Hazen, R.M.; Sholl, D.S. Chiral selection on inorganic crystalline surfaces. Nat. Mater. 2003, 2, 367–374. [Google Scholar] [PubMed]
  7. Bolli, M.; Micura, R.; Eschenmoser, A. Pyranosyl-RNA: Chiroselective self-assembly of base sequences by ligative oligomerization of tetranucleotide-2′, 3′-cyclophosphates (with a commentary concerning the origin of biomolecular homochirality). Chem. Biol. 1997, 4, 309–320. [Google Scholar] [CrossRef]
  8. Ribó, J.M.; Crusats, J.; Sagués, F.; Claret, J.; Rubires, R. Chiral sign induction by vortices during the formation of mesophases in stirred solutions. Science 2001, 292, 2063–2066. [Google Scholar]
  9. Ernst, K.-H. Molecular chirality at surfaces. Phys. Status Solidi 2012, 249, 2057–2088. [Google Scholar] [CrossRef]
  10. Weissbuch, I.; Lahav, M. Crystalline architectures as templates of relevance to the origins of homochirality. Chem. Rev. 2011, 111, 3236–3267. [Google Scholar] [CrossRef] [PubMed]
  11. Kitamura, M.; Okada, S.; Suga, S.; Noyori, R. Enantioselective addition of dialkylzincs to aldehydes promoted by chiral amino alcohols. Mechanism and nonlinear effect. J. Am. Chem. Soc. 1989, 111, 4028–4036. [Google Scholar]
  12. Satyanarayana, T.; Abraham, S.; Kagan, H.B. Nonlinear effects in asymmetric catalysis. Angew. Chem. Int. Ed. 2009, 48, 456–494. [Google Scholar] [CrossRef]
  13. Kondepudi, D.K.; Asakura, K. Chiral autocatalysis, spontaneous symmetry breaking, and stochastic behavior. Acc. Chem. Res. 2001, 34, 946–954. [Google Scholar] [CrossRef] [PubMed]
  14. Viedma, C. Chiral symmetry breaking during crystallization: Complete chiral purity induced by nonlinear autocatalysis and recycling. Phys. Rev. Lett. 2005, 94, 065504. [Google Scholar] [CrossRef] [PubMed]
  15. Soloshonok, V.A.; Ueki, H.; Yasumoto, M.; Mekala, S.; Hirschi, J.S.; Singleton, D.A. Phenomenon of optical self-purification of chiral non-racemic compounds. J. Am. Chem. Soc. 2007, 129, 12112–12113. [Google Scholar] [CrossRef] [PubMed]
  16. Hayashi, Y.; Matsuzawa, M.; Yamaguchi, J.; Yonehara, S.; Matsumoto, Y.; Shoji, M.; Hashizume, D.; Koshino, H. Large nonlinear effect observed in the enantiomeric excess of proline in solution and that in the solid state. Angew. Chem. Int. Ed. 2006, 45, 4593–4597. [Google Scholar] [CrossRef] [PubMed]
  17. Córdova, A.; Engqvist, M.; Ibrahem, I.; Casas, J.; Sundén, H. Plausible origins of homochirality in the amino acid catalyzed neogenesis of carbohydrates. Chem. Commun. 2005, 2047–2049. [Google Scholar] [CrossRef]
  18. Green, M.M.; Park, J.-W.; Sato, T.; Teramoto, A.; Lifson, S.; Selinger, R.L.B.; Selinger, J.V. The macromolecular route to chiral amplification. Angew. Chem. Int. Ed. 1999, 38, 3138–3154. [Google Scholar] [CrossRef]
  19. Noorduin, W.L.; Vlieg, E.; Kellogg, R.M.; Kaptein, B. From Ostwald ripening to single chirality. Angew. Chem. Int. Ed. 2009, 48, 9600–9606. [Google Scholar] [CrossRef]
  20. Saito, Y.; Hyuga, H. Colloquium: Homochirality: Symmetry breaking in systems driven far from equilibrium. Rev. Mod. Phys. 2013, 85, 603–621. [Google Scholar] [CrossRef]
  21. Frank, F.C. On spontaneous asymmetric synthesis. Biochim. Biophys. Acta 1953, 11, 459–463. [Google Scholar] [CrossRef]
  22. Han, J.; Kitagawa, O.; Wzorek, A.; Klia, K.D.; Soloshonok, V.A. The self-disproportionation of enantiomers (SDE): A menace or an opportunity? Chem. Sci. 2018, 9, 1718–1739. [Google Scholar] [CrossRef]
  23. Moberg, C. Recycling in asymmetric catalysis. Acc. Chem. Res. 2016, 49, 2736–2745. [Google Scholar] [CrossRef] [PubMed]
  24. Soai, K.; Shibata, T.; Sato, I. Enantioselective automultiplication of chiral molecules by asymmetric autocatalysis. Acc. Chem. Res. 2000, 33, 382–390. [Google Scholar] [CrossRef] [PubMed]
  25. Soai, K.; Kawasaki, T. Discovery of asymmetric autocatalysis with amplification of chirality and its implication in chiral homogeneity of biomolecules. Chirality 2006, 18, 469–478. [Google Scholar] [CrossRef] [PubMed]
  26. Soai, K.; Kawasaki, T. Asymmetric autocatalysis with amplification of chirality. Top. Curr. Chem. 2008, 284, 1–31. [Google Scholar]
  27. Kawasaki, T.; Soai, K. Amplification of chirality as a pathway to biological homochirality. J. Fluor. Chem. 2010, 131, 525–534. [Google Scholar] [CrossRef]
  28. Kawasaki, T.; Soai, K. Asymmetric induction arising from enantiomerically enriched carbon-13 isotopomers and highly sensitive chiral discrimination by asymmetric autocatalysis. Bull. Chem. Soc. Jpn. 2011, 84, 879–892. [Google Scholar] [CrossRef]
  29. Kawasaki, T.; Soai, K. Asymmetric autocatalysis triggered by chiral crystals formed from achiral compounds and chiral isotopomers. Isr. J. Chem. 2012, 52, 582–590. [Google Scholar] [CrossRef]
  30. Soai, K.; Kawasaki, T. Asymmetric Autocatalysis—Discovery and State of The Art. In The Soai Reaction and Related Topic; Palyi, G., Zicchi, C., Caglioti, C., Eds.; Academia Nationale di Scienze Lettere e Arti Modena, Edizioni Artestampa: Modena, Italy, 2012; pp. 9–34. [Google Scholar]
  31. Soai, K.; Kawasaki, T.; Matsumoto, A. The origins of homochirality examined by using asymmetric autocatalysis. Chem. Rec. 2014, 14, 70–83. [Google Scholar] [CrossRef]
  32. Soai, K.; Kawasaki, T.; Matsumoto, A. Asymmetric autocatalysis of pyrimidyl alkanol and its application to the study on the origin of homochirality. Acc. Chem. Res. 2014, 47, 3643–3654. [Google Scholar] [CrossRef]
  33. Soai, K.; Kawasaki, T.; Matsumoto, A. Asymmetric autocatalysis of pyrimidyl alkanol and related compounds. Self-replication, amplification of chirality and implication for the origin of biological enantioenriched chirality. Tetrahedron 2018, 74, 1973–1990. [Google Scholar] [CrossRef]
  34. Soai, K. Asymmetric autocatalysis. Chiral symmetry breaking and the origins of homochirality of organic molecules. Proc. Jpn. Acad. Ser. B 2019, 95, 89–110. [Google Scholar] [CrossRef]
  35. Podlech, J.; Gehring, T. New aspects of Soai’s asymmetric autocatalysis. Angew. Chem. Int. Ed. 2005, 44, 5776–5777. [Google Scholar] [CrossRef]
  36. Gehring, T.; Busch, M.; Schlageter, M.; Weingand, D. A concise summary of experimental facts about the Soai reaction. Chirality 2010, 22, E173–E182. [Google Scholar] [CrossRef]
  37. Soai, K.; Niwa, S.; Hori, H. Asymmetric self-catalytic reaction. Self-production of chiral 1-(3-pyridyl) alkanols as chiral self-catalysts in the enantioselective addition of dialkylzinc reagents to pyridine-3-carbaldehyde. J. Chem. Soc. Chem. Commun. 1990, 982–983. [Google Scholar] [CrossRef]
  38. Soai, K.; Hayase, T.; Shimada, C.; Isobe, K. Catalytic asymmetric synthesis of chiral diol, bis[2-(l-hydroxyalkyl)phenylether, an asymmetric autocatalytic reaction. Tetrahedron Asymm. 1994, 5, 789–792. [Google Scholar] [CrossRef]
  39. Soai, K.; Shibata, T.; Morioka, H.; Choji, K. Asymmetric autocatalysis and amplification of enantiomeric excess of a chiral molecule. Nature 1995, 378, 767–768. [Google Scholar] [CrossRef]
  40. Shibata, T.; Morioka, H.; Hayase, T.; Choji, K.; Soai, K. Highly enantioselective catalytic asymmetric automultiplication of chiral pyrimidylalcohol. J. Am. Chem. Soc. 1996, 118, 471–472. [Google Scholar] [CrossRef]
  41. Shibata, T.; Yonekubo, S.; Soai, K. Practically perfect asymmetric autocatalysis using 2-alkynyl-5-pyrimidylalkanol. Angew. Chem. Int. Ed. 1999, 38, 659–661. [Google Scholar] [CrossRef]
  42. Sato, I.; Urabe, H.; Ishiguro, S.; Shibata, T.; Soai, K. Amplification of chirality from extremely low to greater than 99.5% ee by asymmetric autocatalysis. Angew. Chem. Int. Ed. 2003, 42, 315–317. [Google Scholar] [CrossRef]
  43. Sato, I.; Yanagi, T.; Soai, K. Highly enantioselective asymmetric autocatalysis of 2-alkenyl- and 2-vinyl-5-pyrimidyl alkanols with significant amplification of anantiomeric axcess. Chirality 2002, 14, 166–168. [Google Scholar] [CrossRef]
  44. Shibata, T.; Choji, K.; Morioka, H.; Hayase, T.; Soai, K. Highly enantioselective synthesis of a chiral 3-quinolylalkanol by an asymmetric autocatalytic reaction. Chem. Commun. 1996, 751–752. [Google Scholar] [CrossRef]
  45. Shibata, T.; Choji, K.; Hayase, T.; Aizu, Y.; Soai, K. Asymmetric autocatalytic reaction of 3-quinolylalkanol with amplification of enantiomeric excess. Chem. Commun. 1996, 1235–1236. [Google Scholar] [CrossRef]
  46. Sato, I.; Nakao, T.; Sugie, R.; Kawasaki, T.; Soai, K. Enantioselective synthesis of substituted 3-quinolyl alkanols and their application to asymmetric autocatalysis. Synthesis 2004, 1419–1428. [Google Scholar] [CrossRef]
  47. Shibata, T.; Morioka, H.; Tanji, S.; Hayase, T.; Kodaka, Y.; Soai, K. Enantioselective synthesis of chiral 5-carbamoyl-3-pyridyl alcohols by asymmetric autocatalytic reaction. Tetrahedron Lett. 1996, 37, 8783–8786. [Google Scholar] [CrossRef]
  48. Tanji, S.; Kodaka, Y.; Ohno, A.; Shibata, T.; Sato, I.; Soai, K. Asymmetric autocatalysis of 5-carbamoyl-3-pyridyl alkanols with amplification of enantiomeric excess. Tetrahedron Asymm. 2000, 11, 4249–4253. [Google Scholar] [CrossRef]
  49. Kawasaki, T.; Nakaoda, M.; Takahashi, Y.; Kanto, Y.; Kuruhara, N.; Hosoi, K.; Sato, I.; Matsumoto, A.; Soai, K. Self-replication and amplification of enantiomeric excess of chiral multi-functionalized large molecule by asymmetric autocatalysis. Angew. Chem. Int. Ed. 2014, 53, 11199–11202. [Google Scholar] [CrossRef]
  50. Kawasaki, T.; Ishikawa, Y.; Minato, Y.; Otsuka, T.; Yonekubo, S.; Sato, I.; Shibata, T.; Matsumoto, A.; Soai, K. Point-to-point ultra-remote asymmetric control with flexible linker. Chem. A Eur. J. 2017, 23, 282–285. [Google Scholar] [CrossRef]
  51. Kitamura, M.; Suga, S.; Oka, H.; Noyori, R. Quantitative analysis of the chiral amplification in the amino alcohol-promoted asymmetric alkylation of aldehydes with dialkylzincs. J. Am. Chem. Soc. 1998, 120, 9800–9809. [Google Scholar] [CrossRef]
  52. Guillaneux, D.; Zhao, S.-H.; Samuel, O.; Rainford, D.; Henri, B.; Kagan, H.B. Nonlinear effects in asymmetric catalysis. J. Am. Chem. Soc. 1994, 116, 9430–9439. [Google Scholar] [CrossRef]
  53. Sato, I.; Omiya, D.; Tsukiyama, K.; Ogi, Y.; Soai, K. Evidence of asymmetric autocatalysis in the enantioselective addition of diisopropylzinc to pyrimidine-5-carbaldehyde using chiral pyrimidyl alkanol. Tetrahedron Asymm. 2001, 12, 1965–1969. [Google Scholar] [CrossRef]
  54. Sato, I.; Omiya, D.; Igarashi, H.; Kato, K.; Ogi, Y.; Tsukiyama, K.; Soai, K. Relationship between the time, yield, and enantiomeric excess of asymmetric autocatalysis of chiral 2-alkynyl-5-pyrimidyl alkanol with amplification of enantiomeric excess. Tetrahedron Asymm. 2003, 14, 975–979. [Google Scholar] [CrossRef]
  55. Blackmond, D.G.; McMillan, C.R.; Ramdeehul, S.; Schorm, A.; Brown, J.M. Origins of asymmetric amplification in autocatalytic alkylzinc additions. J. Am. Chem. Soc. 2001, 123, 10103–10104. [Google Scholar] [CrossRef]
  56. Quaranta, M.; Gehring, T.; Odell, B.; Brown, J.M.; Blackmond, D.G. Unusual inverse temperature dependence on reaction rate in the asymmetric autocatalytic alkylation of pyrimidyl aldehydes. J. Am. Chem. Soc. 2010, 132, 15104–15107. [Google Scholar] [CrossRef]
  57. Gehring, T.; Quaranta, M.; Odell, B.; Blackmond, D.G.; Brown, J.M. Observation of a transient intermediate in Soai’s asymmetric autocatalysis: Insights from 1H NMR turnover in real time. Angew. Chem. Int. Ed. 2012, 51, 9539–9542. [Google Scholar] [CrossRef]
  58. Schiaffino, L.; Ercolani, G. Unraveling the mechanism of the Soai asymmetric autocatalytic reaction by first-principles calculations: Induction and amplification of chirality by self-assembly of hexamolecular complexes. Angew. Chem. Int. Ed. 2008, 47, 6832–6835. [Google Scholar] [CrossRef]
  59. Ercolani, G.; Schiaffino, L. Putting the mechanism of the Soai reaction to the test: DFT study of the role of aldehyde and dialkylzinc structure. J. Org. Chem. 2011, 76, 2619–2626. [Google Scholar] [CrossRef]
  60. Gridnev, I.D.; Vorobiev, A.K. Quantification of sophisticated equilibria in the reaction pool and amplifying catalytic cycle of the Soai reaction. ACS Catal. 2012, 2, 2137–2149. [Google Scholar] [CrossRef]
  61. Gridnev, I.D.; Vorobiev, A.K. On the origin and structure of the recently observed acetal in the Soai reaction. Bull. Chem. Soc. Jpn. 2015, 88, 333–340. [Google Scholar] [CrossRef]
  62. Barabás, B.; Caglioti, L.; Micskei, K.; Pályi, G. Data-based stochastic approach to absolute asymmetric synthesis by autocatalysis. Bull. Chem. Soc. Jpn. 2009, 82, 1372–1376. [Google Scholar] [CrossRef]
  63. Micheau, J.C.; Cruz, J.M.; Coudret, C.; Buhse, T. An autocatalytic cycle model of asymmetric amplification and mirror-symmetry breaking in the Soai reaction. ChemPhysChem 2010, 11, 3417–3419. [Google Scholar] [CrossRef]
  64. Micheau, J.C.; Coudret, C.; Cruz, J.M.; Buhse, T. Amplification of enantiomeric excess, mirror-image symmetry breaking and kinetic proofreading in Soai reaction models with different oligomeric orders. Phys. Chem. Chem. Phys. 2012, 14, 13239–13248. [Google Scholar] [CrossRef]
  65. Micskei, K.; Rábai, G.; Gál, E.; Caglioti, L.; Pályi, G. Oscillatory symmetry breaking in the Soai reaction. J. Phys. Chem. B 2008, 112, 9196–9200. [Google Scholar] [CrossRef]
  66. Maioli, M.; Micskei, K.; Caglioti, L.; Zucchi, C.; Pályi, G. Evolution of chirality in consecutive asymmetric autocatalytic reaction cycles. J. Math. Chem. 2008, 43, 1505–1515. [Google Scholar] [CrossRef]
  67. Crusats, J.; Hochberg, D.; Moyano, A.; Ribó, J.M. Frank model and spontaneous emergence of chirality in closed systems. Chem. Phys. Chem. 2009, 10, 2123–2131. [Google Scholar] [CrossRef]
  68. Dóka, É.; Lente, G. Mechanism-based chemical understanding of chiral symmetry breaking in the Soai reaction. A combined probabilistic and deterministic description of chemical reactions. J. Am. Chem. Soc. 2011, 133, 17878–17881. [Google Scholar] [CrossRef]
  69. Lavabre, D.; Micheau, J.-C.; Islas, J.R.; Buhse, T. Enantioselectivity Reversal by achiral additives in the Soai reaction:  A kinetic understanding. J. Phys. Chem. A 2007, 111, 281–286. [Google Scholar] [CrossRef]
  70. Matsumoto, A.; Abe, T.; Hara, A.; Tobita, T.; Sasagawa, T.; Kawasaki, T.; Soai, K. Crystal structure of isopropylzinc alkoxide of pyrimidyl alkanol: Mechanistic insights for asymmetric autocatalysis with amplification of enantiomeric excess. Angew. Chem. Int. Ed. 2015, 54, 15218–15221. [Google Scholar] [CrossRef]
  71. Matsumoto, A.; Fujiwara, S.; Abe, T.; Hara, A.; Tobita, T.; Sasagawa, T.; Kawasaki, T.; Soai, K. Elucidation of the structures of asymmetric autocatalyst based on X-ray crystallography. Bull. Chem. Soc. Jpn. 2016, 89, 1170–1177. [Google Scholar] [CrossRef]
  72. Noble-Teran, M.E.; Cruz, J.-M.; Micheau, J.-C.; Buhse, T.W. A quantification of the Soai reaction. ChemCatChem 2018, 10, 642–648. [Google Scholar] [CrossRef]
  73. Bailey, J.; Chrysostomou, A.; Hough, J.H.; Gledhill, T.M.; McCall, A.; Clark, S.; Ménard, F.; Tamura, M. Circular polarization in star-formation regions: Implications for biomolecular homochirality. Science 1998, 281, 672–674. [Google Scholar] [CrossRef] [PubMed]
  74. Shibata, T.; Yamamoto, J.; Matsumoto, N.; Yonekubo, S.; Osanai, S.; Soai, K. Amplification of a slight enantiomeric imbalance in molecules based on asymmetric autocatalysis—The first correlation between high enantiomeric enrichment in a chiral molecule and circularly polarized light. J. Am. Chem. Soc. 1998, 120, 12157–12158. [Google Scholar] [CrossRef]
  75. Kawasaki, T.; Sato, M.; Ishiguro, S.; Saito, T.; Morishita, Y.; Sato, I.; Nishino, H.; Inoue, Y.; Soai, K. Enantioselective synthesis of near enantiopure compound by asymmetric autocatalysis triggered by asymmetric photolysis with circularly polarized light. J. Am. Chem. Soc. 2005, 127, 3274–3275. [Google Scholar] [CrossRef] [PubMed]
  76. Sato, I.; Sugie, R.; Matsueda, Y.; Furumura, Y.; Soai, K. Asymmetric synthesis utilizing circularly polarized light mediated by the photoequilibrium of chiral olefins in conjunction with asymmetric autocatalysis. Angew. Chem. Int. Ed. 2004, 43, 4490–4492. [Google Scholar] [CrossRef] [PubMed]
  77. Noorduin, W.L.; Bode, A.C.; van der Meijden, M.; Meekes, H.; van Etteger, A.F.; van Enckevort, W.J.P.; Christianen, P.C.M.; Kaptein, B.; Kellogg, R.M.; Rasing, T.; Vlieg, E. Complete chiral symmetry breaking of an amino acid derivative directed by circularly polarized light. Nat. Chem. 2009, 1, 729–732. [Google Scholar] [CrossRef] [Green Version]
  78. Bonner, W.A.; Kavasmaneck, P.R.; Martin, F.S.; Flores, J.J. Asymmetric adsorption of alanine by quartz. Science 1974, 186, 143–144. [Google Scholar] [CrossRef]
  79. Soai, K.; Osanai, S.; Kadowaki, K.; Yonekubo, S.; Shibata, T.; Sato, I. d- and l-Quartz-promoted highly enantioselective synthesis of a chiral organic compound. J. Am. Chem. Soc. 1999, 121, 11235–11236. [Google Scholar] [CrossRef]
  80. Kondepudi, D.K.; Kaufman, R.J.; Singh, N. Chiral symmetry breaking in sodium chlorate crystallizaton. Science 1990, 250, 975–976. [Google Scholar] [CrossRef]
  81. McBride, J.M.; Carter, R.L. Spontaneous resolution by stirred crystallization. Angew. Chem. Int. Ed. 1991, 30, 293–295. [Google Scholar] [CrossRef]
  82. Sato, I.; Kadowaki, K.; Soai, K. Asymmetric synthesis of an organic compound with high enantiomeric excess induced by inorganic ionic sodium chlorate. Angew. Chem. Int. Ed. 2000, 39, 1510–1512. [Google Scholar] [CrossRef]
  83. Sato, I.; Kadowaki, K.; Ohgo, Y.; Soai, K. Highly enantioselective asymmetric autocatalysis induced by chiral ionic crystals of sodium chlorate and sodium bromate. J. Mol. Cat. A Chem. 2004, 216, 209–214. [Google Scholar] [CrossRef]
  84. Shindo, H.; Shirota, Y.; Niki, K.; Kawasaki, T.; Suzuki, K.; Araki, Y.; Matsumoto, A.; Soai, K. Asymmetric autocatalysis induced by cinnabar: Observation of the enantioselective adsorption of a 5-pyrimidyl alkanol on the crystal surface. Angew. Chem. Int. Ed. 2013, 52, 9135–9138. [Google Scholar] [CrossRef]
  85. Matsumoto, A.; Ozawa, H.; Inumaru, A.; Soai, K. Asymmetric induction by retgersite, nickel sulfate hexahydrate, in conjunction with asymmetric autocatalysis. New. J. Chem. 2015, 39, 6742–6745. [Google Scholar] [CrossRef] [Green Version]
  86. Matsumoto, A.; Kaimori, Y.; Uchida, M.; Omori, H.; Kawasaki, T.; Soai, K. Achiral inorganic gypsum acts as an origin of chirality through its enaniotopic surface in conjunction with asymmetric autocatalysis. Angew. Chem. Int. Ed. 2017, 56, 545–548. [Google Scholar] [CrossRef]
  87. Matsuura, T.; Koshima, H. Introduction to chiral crystallization of achiral organic compounds. Spontaneous generation of chirality. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 7–24. [Google Scholar] [CrossRef]
  88. Ishikawa, K.; Tanaka, M.; Suzuki, T.; Sekine, A.; Kawasaki, T.; Soai, K.; Shiro, M.; Lahav, M.; Asahi, T. Absolute chirality of the gamma-polymorph of glycine: Correlation of the absolute structure with the optical rotation. Chem. Commun. 2012, 48, 6031–6033. [Google Scholar] [CrossRef] [PubMed]
  89. Tarasevych, A.V.; Sorochinsky, A.E.; Kukhar, V.P.; Toupet, L.; Crassous, J.; Guillemin, J.-C. Attrition-induced spontaneous chiral amplification of the γ polymorphic modification of glycine. Cryst. Eng. Comm. 2015, 17, 1513–1517. [Google Scholar] [CrossRef] [Green Version]
  90. Matsumoto, A.; Ozaki, H.; Tsuchiya, S.; Asahi, T.; Lahav, M.; Kawasaki, T.; Soai, K. Achiral amino acid glycine acts as an origin of homochirality in asymmetric autocatalysis. Org. Biomol. Chem. 2019, 17, 4200–4203. [Google Scholar] [CrossRef]
  91. Robertson, M.P.; Miller, S.L. An efficient prebiotic synthesis of cytosine and uracil. Nature 1995, 375, 772–774. [Google Scholar] [CrossRef]
  92. Kawasaki, T.; Suzuki, K.; Hakoda, Y.; Soai, K. Achiral nucleobase cytosine acts as an origin of homochirality of biomolecules in conjunction with asymmetric autocatalysis. Angew. Chem. Int. Ed. 2008, 47, 496–499. [Google Scholar] [CrossRef]
  93. Kawasaki, T.; Hakoda, Y.; Mineki, H.; Suzuki, K.; Soai, K. Generation of absolute controlled crystal chirality by the removal of crystal water from achiral crystal of nucleobase cytosine. J. Am. Chem. Soc. 2010, 132, 2874–2875. [Google Scholar] [CrossRef]
  94. Mineki, H.; Kaimori, Y.; Kawasaki, T.; Matsumoto, A.; Soai, K. Enantiodivergent formation of a chiral cytosine crystal by removal of crystal water from an achiral monohydrate crystal under reduced pressure. Tetrahedron Asymm. 2013, 24, 1365–1367. [Google Scholar] [CrossRef]
  95. Mineki, H.; Hanasaki, T.; Matsumoto, A.; Kawasaki, T.; Soai, K. Asymmetric autocatalysis initiated by achiral nucleic acid base adenine: Implications on the origin of homochirality of biomolecul. Chem. Commun. 2012, 48, 10538–10540. [Google Scholar] [CrossRef]
  96. Kawasaki, T.; Suzuki, K.; Hatase, K.; Otsuka, M.; Koshima, H.; Soai, K. Enantioselective synthesis mediated by chiral crystal of achiral hippuric acid in conjunction with asymmetric autocatalysis. Chem. Commun. 2006, 1869–1871. [Google Scholar] [CrossRef]
  97. Carter, D.J.; Rohl, A.L.; Shtukenberg, A.; Bian, S.D.; Hu, C.-H.; Baylon, L.; Kahr, B.; Mineki, H.; Abe, K.; Kawasaki, T. Prediction of Soai reaction enantioselectivity induced by crystals of N-(2-thienylcarbonyl) glycine. Cryst. Growth Des. 2012, 12, 2138–2145. [Google Scholar] [CrossRef]
  98. Kawasaki, T.; Jo, K.; Igarashi, H.; Sato, I.; Nagano, M.; Koshima, H.; Soai, K. Asymmetric amplification using chiral co-crystal f formed from achiral organic molecules by asymmetric autocatalysis. Angew. Chem. Int. Ed. 2005, 44, 2774–2777. [Google Scholar] [CrossRef]
  99. Kawasaki, T.; Harada, Y.; Suzuki, K.; Tobita, T.; Florini, N.; Palyi, G.; Soai, K. Enantioselective synthesis utilizing enantiomorphous organic crystal of achiral benzils as a source of chirality in asymmetric autocatalysis. Org. Lett. 2008, 10, 4085–4088. [Google Scholar] [CrossRef]
  100. Kawasaki, T.; Nakaoda, M.; Kaito, N.; Sasagawa, T.; Soai, K. Asymmetric autocatalysis induced by chiral crystals of achiral tetraphenylethylenes. Orig. Life Evol. Biosph. 2010, 40, 65–78. [Google Scholar] [CrossRef]
  101. Matsumoto, A.; Ide, T.; Kaimori, Y.; Fujiwara, S.; Soai, K. Asymmetric autocatalysis triggered by chiral crystal of achiral ethylenediamine sulfate. Chem. Lett. 2015, 44, 688–690. [Google Scholar] [CrossRef]
  102. Kawasaki, T.; Uchida, M.; Kaimori, Y.; Sasagawa, T.; Matsumoto, A.; Soai, K. Enantioselective synthesis induced by the helical molecular arrangement in the chiral crystal of achiral tris(2-hydroxyethyl)-1,3,5-benzenetricarboxylate in conjunction with asymmetric autocatalysis. Chem. Lett. 2013, 42, 711–713. [Google Scholar] [CrossRef]
  103. Matsumoto, A.; Takeda, S.; Harada, S.; Soai, K. Determination of the absolute structure of the chiral crystal consisting of achiral dibutylhydroxytoluene and asymmetric autocatalysis triggered by this chiral crystal. Tetrahedron Asymm. 2016, 27, 943–946. [Google Scholar] [CrossRef] [Green Version]
  104. Kawasaki, T.; Sasagawa, T.; Shiozawa, K.; Uchida, M.; Suzuki, K.; Soai, K. Enantioselective synthesis induced by chiral crystal composed of dl-serine in conjunction with asymmetric autocatalysis. Org. Lett. 2011, 13, 2361–2363. [Google Scholar] [CrossRef]
  105. Kawasaki, T.; Kamimura, S.; Amihara, A.; Suzuki, K.; Soai, K. Enantioselective C-C bond formation as a result of the oriented prochirality of an achiral aldehyde at the single-crystal face upon treatment with a dialkyl zinc vapor. Angew. Chem. Int. Ed. 2011, 50, 6796–6798. [Google Scholar] [CrossRef]
  106. Caglioti, L.; Hajdu, C.; Holczknecht, O.; Zékány, L.; Zucchi, C.; Micskei, K.; Pályi, G. The concept of racemates and the Soai-reaction. Viva Origino 2006, 34, 62–80. [Google Scholar]
  107. Maioli, M.; Varadi, G.; Kurdi, R.; Caglioti, L.; Palyi, G. Limits of the classical concept of concentration. J. Phys. Chem. B 2016, 120, 7438–7445. [Google Scholar] [CrossRef]
  108. Barabas, B.; Caglioti, L.; Zucchi, C.; Maioli, M.; Gál, E.; Micskei, K.; Pályi, G. Violation of distribution symmetry in statistical evaluation of absolute enantioselective synthesis. J. Phys. Chem. B 2007, 111, 11506–11510. [Google Scholar] [CrossRef]
  109. Soai, K.; Shibata, T.; Kowata, Y. Japan Kokai Tokkyo Koho. Patent No. JP1997–268179, 2 January 1997. [Google Scholar]
  110. Soai, K.; Sato, I.; Shibata, T.; Komiya, S.; Hayashi, M.; Matsueda, Y.; Imamura, H.; Hayase, T.; Morioka, H.; Tabira, H.; Yamamoto, J.; Kowata, Y. Asymmetric synthesis of pyrimidyl alkanol without adding chiral substances by the addition of diisopropylzinc to pyrimidine-5-carbaldehyde in conjunction with asymmetric autocatalysis. Tetrahedron Asymm. 2003, 14, 185–188. [Google Scholar] [CrossRef]
  111. Kawasaki, T.; Suzuki, K.; Shimizu, M.; Ishikawa, K.; Soai, K. Spontaneous absolute asymmetric synthesis in the presence of achiral silica gel in conjunction with asymmetric autocatalysis. Chirality 2006, 18, 479–482. [Google Scholar] [CrossRef]
  112. Suzuki, K.; Hatase, K.; Nishiyama, D.; Kawasaki, T.; Soai, K. Spontaneous absolute asymmetric synthesis promoted by achiral amines in conjunction with asymmetric autocatalysis. J. Syst. Chem. 2010, 1, 5. [Google Scholar] [CrossRef] [Green Version]
  113. Singleton, D.A.; Vo, L.K. A few molecules can control the enantiomeric outcome. Evidence supporting absolute asymmetric synthesis using the Soai asymmetric autocatalysis. Org. Lett. 2003, 5, 4337–4339. [Google Scholar] [CrossRef]
  114. Lente, G. Stochastic kinetic models of chiral autocatalysis: A general tool for the quantitative interpretation of total asymmetric synthesis. J. Phys. Chem. A 2005, 109, 11058–11063. [Google Scholar] [CrossRef]
  115. Islas, J.R.; Lavabre, D.; Grevy, J.-M.; Lamoneda, R.H.; Cabrera, H.R.; Micheau, J.-C.; Buhse, T. Mirror-symmetry breaking in the Soai reaction: A kinetic understanding. Proc. Natl. Acad. Sci. USA 2005, 102, 13743–13748. [Google Scholar] [CrossRef] [Green Version]
  116. Lavabre, D.; Micheau, J.-C.; Rivera Islas, J.; Buhse, T. Kinetic insight into specific features of the autocatalytic Soai reaction. Top. Curr. Chem. 2008, 284, 67–96. [Google Scholar]
  117. Saito, Y.; Hyuga, H. Rate equation approaches to amplification of enantiomeric excess and chiral symmetry breaking. Top. Curr. Chem. 2008, 284, 97–118. [Google Scholar]
  118. Kaimori, Y.; Hiyoshi, Y.; Kawasaki, T.; Matsumoto, A.; Soai, K. Formation of enantioenriched alkanol with stochastic distribution of enantiomers in the absolute asymmetric synthesis under heterogeneous solid–vapor phase conditions. Chem. Commun. 2019, 55, 5223–5226. [Google Scholar] [CrossRef]
  119. Kawasaki, T.; Matsumura, Y.; Tsutsumi, T.; Suzuki, K.; Ito, M.; Soai, K. Asymmetric autocatalysis triggered by carbon isotope (13C/12C) chirality. Science 2009, 324, 492–495. [Google Scholar] [CrossRef] [PubMed]
  120. Matsumoto, A.; Ozaki, H.; Harada, S.; Tada, K.; Ayugase, T.; Ozawa, H.; Kawasaki, T.; Soai, K. Asymmetric induction by nitrogen 14N/15N isotopomer in conjunction with asymmetric autocatalysis. Angew. Chem. Int. Ed. 2016, 55, 15246–15249. [Google Scholar] [CrossRef]
  121. Kawasaki, T.; Okano, Y.; Suzuki, E.; Takano, S.; Oji, S.; Soai, K. Asymmetric autocatalysis: Triggered by chiral isotopomer arising from oxygen isotope substitution. Angew. Chem. Int. Ed. 2011, 50, 8131–8133. [Google Scholar] [CrossRef]
  122. Matsumoto, A.; Oji, S.; Takano, S.; Tada, K.; Kawasaki, T.; Soai, K. Asymmetric autocatalysis triggered by oxygen isotopically chiral glycerin. Org. Biomol. Chem. 2013, 11, 2928–2931. [Google Scholar] [CrossRef]
  123. Horeau, A.; Nouaille, A.; Mislow, K. Secondary deuterium isotope effects in asymmetric syntheses and kinetic resolutions. J. Am. Chem. Soc. 1965, 87, 4957–4958. [Google Scholar] [CrossRef]
  124. Pracejus, H. Ein sterischer isotopeneffekt als ursache einer katalytisch-asymmetrischen synthese. Tetrahedron Lett. 1966, 7, 3809–3813. [Google Scholar] [CrossRef]
  125. Sato, I.; Omiya, D.; Saito, T.; Soai, K. Highly enantioselective synthesis induced by chiral primary alcohols due to deuterium substitution. J. Am. Chem. Soc. 2000, 122, 11739–11740. [Google Scholar] [CrossRef]
  126. Kawasaki, T.; Ozawa, H.; Ito, M.; Soai, K. Enantioselective synthesis induced by compounds with chirality arising from partially deuterated methyl groups in conjunction with asymmetric autocatalysis. Chem. Lett. 2011, 40, 320–321. [Google Scholar] [CrossRef]
  127. Kawasaki, T.; Shimizu, M.; Nishiyama, D.; Ito, M.; Ozawa, H.; Soai, K. Asymmetric autocatalysis induced by meteoritic amino acids with hydrogen isotope chirality. Chem. Commun. 2009, 4396–4398. [Google Scholar] [CrossRef]
  128. Sato, I.; Ohgo, Y.; Igarashi, H.; Nishiyama, D.; Kawasaki, T.; Soai, K. Determination of absolute configurations of amino acids by asymmetric autocatalysis of 2-alkynylpyrimidyl alkanol as a chiral sensor. J. Organomet. Chem. 2007, 692, 1783–1787. [Google Scholar] [CrossRef]
  129. Sato, I.; Yamashima, R.; Kadowaki, K.; Yamamoto, J.; Shibata, T.; Soai, K. Asymmetric induction by helical hydrocarbons: [6]- and [5]helicenes. Angew. Chem. Int. Ed. 2001, 40, 1096–1098. [Google Scholar] [CrossRef]
  130. Kawasaki, T.; Suzuki, K.; Licandro, E.; Bossi, A.; Maiorana, S.; Soai, K. Enantioselective synthesis induced by tetrathia-[7]-helicenes in conjunction with asymmetric autocatalysis. Tetrahedron Asymm. 2006, 17, 2050–2053. [Google Scholar] [CrossRef]
  131. Matsumoto, A.; Yonemitsu, K.; Ozaki, H.; Míšek, J.; Starý, I.; Stará, I.G.; Soai, K. Reversal of the sense of enantioselectivity between 1- and 2-aza[6]helicenes used as chiral inducers of asymmetric autocatalysis. Org. Biomol. Chem. 2017, 15, 1321–1324. [Google Scholar] [CrossRef] [Green Version]
  132. Kawasaki, T.; Tanaka, H.; Tsutsumi, T.; Kasahara, T.; Sato, I.; Soai, K. Chiral discrimination of cryptochiral saturated quaternary and tertiary hydrocarbons by asymmetric autocatalysis. J. Am. Chem. Soc. 2006, 128, 6032–6033. [Google Scholar] [CrossRef]
  133. Kawasaki, T.; Hohberger, C.; Araki, Y.; Hatase, K.; Beckerle, K.; Okuda, J.; Soai, K. Discrimination of cryptochirality in chiral isotactic polystyrene by asymmetric autocatalysis. Chem. Commun. 2009, 5621–5623. [Google Scholar] [CrossRef]
  134. Sato, I.; Kadowaki, K.; Urabe, H.; Hwa Jung, J.; Ono, Y.; Shinkai, S.; Soai, K. Highly enantioselective synthesis of organic compound using right- and left-handed helical silica. Tetrahedron Lett. 2003, 44, 721–724. [Google Scholar] [CrossRef]
  135. Kawasaki, T.; Araki, Y.; Hatase, K.; Suzuki, K.; Matsumoto, A.; Yokoi, T.; Kubota, Y.; Tatsumi, T.; Soai, K. Helical mesoporous silica as an inorganic heterogeneous chiral trigger for asymmetric autocatalysis with amplification of enantiomeric excess. Chem. Commun. 2015, 51, 8742–8744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Principle of asymmetric autocatalysis.
Scheme 1. Principle of asymmetric autocatalysis.
Symmetry 11 00694 sch001
Scheme 2. Asymmetric autocatalysis. Structures of the autocatalysts of pyrimidyl alkanols 1af, multi-functionalized pyrimidyl alkanol, 3; 3-quinolyl alkanol, 4; and 5-carbamoyl-3-pyridyl alkanol, 5.
Scheme 2. Asymmetric autocatalysis. Structures of the autocatalysts of pyrimidyl alkanols 1af, multi-functionalized pyrimidyl alkanol, 3; 3-quinolyl alkanol, 4; and 5-carbamoyl-3-pyridyl alkanol, 5.
Symmetry 11 00694 sch002
Scheme 3. Asymmetric autocatalysis of 5-pyrimidyl alkanol, 1c, with amplification of enantiomeric excess from ca. 0.00005% to >99.5% ee.
Scheme 3. Asymmetric autocatalysis of 5-pyrimidyl alkanol, 1c, with amplification of enantiomeric excess from ca. 0.00005% to >99.5% ee.
Symmetry 11 00694 sch003
Scheme 4. Circularly polarized light (CPL) triggers asymmetric autocatalysis.
Scheme 4. Circularly polarized light (CPL) triggers asymmetric autocatalysis.
Symmetry 11 00694 sch004
Scheme 5. Asymmetric autocatalysis triggered by chiral quartz, sodium chlorate, cinnabar and enantiotopic face of achiral crystal of gypsum.
Scheme 5. Asymmetric autocatalysis triggered by chiral quartz, sodium chlorate, cinnabar and enantiotopic face of achiral crystal of gypsum.
Symmetry 11 00694 sch005
Scheme 6. Asymmetric autocatalysis triggered by chiral γ-polymorph of achiral glycine.
Scheme 6. Asymmetric autocatalysis triggered by chiral γ-polymorph of achiral glycine.
Symmetry 11 00694 sch006
Scheme 7. Asymmetric autocatalysis initiated by chiral crystals composed of achiral organic compounds.
Scheme 7. Asymmetric autocatalysis initiated by chiral crystals composed of achiral organic compounds.
Symmetry 11 00694 sch007
Scheme 8. Asymmetric autocatalysis initiated on the enantiotopic face of an achiral 2-(tert-butyldimethylsilylethynyl) pyrimidine-5-carbaldehyde 2f.
Scheme 8. Asymmetric autocatalysis initiated on the enantiotopic face of an achiral 2-(tert-butyldimethylsilylethynyl) pyrimidine-5-carbaldehyde 2f.
Symmetry 11 00694 sch008
Scheme 9. Spontaneous absolute asymmetric synthesis by asymmetric autocatalysis without the intervention of any chiral factor.
Scheme 9. Spontaneous absolute asymmetric synthesis by asymmetric autocatalysis without the intervention of any chiral factor.
Symmetry 11 00694 sch009
Figure 1. Spontaneous absolute asymmetric synthesis of pyrimidyl alkanol 1. Histograms of the absolute configuration and ee of products.
Figure 1. Spontaneous absolute asymmetric synthesis of pyrimidyl alkanol 1. Histograms of the absolute configuration and ee of products.
Symmetry 11 00694 g001
Scheme 10. Absolute asymmetric synthesis of pyrimidyl alkanol 1c under solid-vapor phase conditions.
Scheme 10. Absolute asymmetric synthesis of pyrimidyl alkanol 1c under solid-vapor phase conditions.
Symmetry 11 00694 sch010
Scheme 11. Asymmetric autocatalysis triggered by carbon (13C/12C), nitrogen (15N/14N), oxygen (18O/16O), and hydrogen (D/H) isotope chirality.
Scheme 11. Asymmetric autocatalysis triggered by carbon (13C/12C), nitrogen (15N/14N), oxygen (18O/16O), and hydrogen (D/H) isotope chirality.
Symmetry 11 00694 sch011

Share and Cite

MDPI and ACS Style

Soai, K.; Kawasaki, T.; Matsumoto, A. Role of Asymmetric Autocatalysis in the Elucidation of Origins of Homochirality of Organic Compounds. Symmetry 2019, 11, 694. https://doi.org/10.3390/sym11050694

AMA Style

Soai K, Kawasaki T, Matsumoto A. Role of Asymmetric Autocatalysis in the Elucidation of Origins of Homochirality of Organic Compounds. Symmetry. 2019; 11(5):694. https://doi.org/10.3390/sym11050694

Chicago/Turabian Style

Soai, Kenso, Tsuneomi Kawasaki, and Arimasa Matsumoto. 2019. "Role of Asymmetric Autocatalysis in the Elucidation of Origins of Homochirality of Organic Compounds" Symmetry 11, no. 5: 694. https://doi.org/10.3390/sym11050694

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop