Next Article in Journal
One-Pot Synthesis of Tetrazole–Triazole Bis-Heterocycles via Ugi–Azide Reaction
Previous Article in Journal
Biological Docking and BSA Binding Studies of 1,4-Disubstituted Piperdine Containing 1,2,4-Triazoles: Comparative Synthesis Leveraging Microwave-Assisted and Conventional Protocols
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Organocatalytic Properties of 3,4-Dihydroxyprolines †

Centro Singular de Investigación en Química Biolóxica e Materiais Moleculares, Departamento de Química Orgánica, Universidade de Santiago de Compostela, 15782 Santiago de Compostela, Spain
*
Author to whom correspondence should be addressed.
Presented at the 27th International Electronic Conference on Synthetic Organic Chemistry (ECSOC-27), 15–30 November 2023; Available online: https://ecsoc-27.sciforum.net/.
Chem. Proc. 2023, 14(1), 107; https://doi.org/10.3390/ecsoc-27-16117
Published: 15 November 2023

Abstract

:
The synthesis and organocatalytic properties of (2S,3R,4R)-3,4-bis((tert-butyldimethylsilyl)oxy)pyrrolidine-2-carboxylic acid are reported. Using the aldol condensation of cyclohexanone with p-nitrobenzaldehyde as a model, a yield of 86%, an enantiomeric excess of 99% and a diastereomeric excess of 25:1 were achieved.

1. Introduction

Asymmetric synthesis is the most powerful tool available to organic chemists for synthesizing molecules of high functional and stereochemical complexity as it allows stereoselective introduction of stereogenic centers.
Among the available strategies, catalytic methods are particularly attractive as they avoid having to use stoichiometric amounts of expensive chiral reagents. In addition to enzymes and transition metals, the use of organocatalysts has shown enormous potential, having allowed access to natural products by efficient, economical and environmentally benign procedures. Their tolerance to moisture and oxygen, as well as their compatibility with mild reaction conditions and low toxicity, is particularly attractive.
The use of small organic molecules as organocatalysts was first described independently by Eder [1] and by Hajos [2]. However, it was only recently, following the contributions of List and Barbas III [3] and the seminal work of McMillan [4], that the high potential of organocatalysis was rediscovered, leading to an intensive study of its synthetic possibilities which continues today.
The asymmetric aldol condensation reaction is an attractive method of forming carbon–carbon bonds by the enantioselective production of aldols. A wide range of organocatalysts have been developed that have allowed excellent results to be achieved in the asymmetric version of this and other organic reactions [5], and two main mechanisms of organocatalytic processes have been proposed: enamine catalysis [4] and iminium catalysis [5]. While iminium catalysis makes use of chiral imidazolium salts to activate aldehydes by reversible formation of an iminium ion, enamine catalysis uses amino acids (or derivatives), of particular interest being L-proline, whose conformational rigidity favors selectivity. L-proline and similar catalysts act via an enamine intermediate, with the catalyst performing two specific functions: it first activates the nucleophile through the formation of an enamine, and then activates the electrophile to which it coordinates via its carboxyl group. All this leads to a transition state that explains the high selectivity of the reaction [6].
Although the existing organocatalysts have achieved excellent results, there are problems that have not yet been satisfactorily solved: the use of organic solvents is required and, in addition, a fairly high catalyst load is usually necessary.
The most recent efforts in this field are directed towards the use of water as a solvent as it is a safe and environmentally friendly medium, avoiding the contamination problems inherent to organic solvents.
The first case of the use of proline as organocatalysts in an aqueous medium for the aldol reaction is shown in Scheme 1, corresponding to the reaction of benzaldehyde with cyclohexanone, catalyzed by (4R)-4-((tert-butyldiphenylsilyl)oxy)-L-proline (3), which gave the adduct 4 with a diastereoisomeric ratio of 13:1 and an enantiomeric excess of 99% [7].
An ulterior similar contribution involved the aldol condensation of p-nitrobenzaldehyde with cyclohexanone, catalyzed by (2S,4R)-4-((tert-butyldiphenylsilyl)oxy)-L-proline (3), providing adduct 9 with a 20:1 diastereoisomeric ratio and an enantiomeric excess of 99% (Scheme 2) [8]. This reaction was used as a model for a similar organocatalytic with 3,4-dihydroxyproline, as reported by studies.

2. Result and Discussion

Proline 9 was prepared from the known proline 7, according to the protocol shown in Scheme 3.
Next, proceeding as depicted in Scheme 4, to a solution of p-nitrobenzaldehyde (1.0 eq) and cyclohexanone (5.0 eq) in water the catalyst 8 (0.1 eq) was added and the mixture was stirred at room temperature for 5 hours, stopping the reaction by neutralizing with a phosphate-buffered solution pH 7. The elaboration of the reaction mixture was followed by purification by column chromatography [AcOEt/Hex 1:4]. The mixture of aldols 10, 11, 12 and 13 was isolated with a yield of 86%.
From its 1H NMR spectrum (Figure 1), the diastereomeric relationship between the anti and syn enantiomer pairs could be easily established. For the anti enantiomers, the signal due to the proton at position 4 (at α to the hydroxyl group) appears as a doublet of doublets located at 4.89 ppm, whereas in the case of the syn enantiomers this signal appears as a triplet at 5.48 ppm. Relative integration of the two signals gave an approximate diastereomeric ratio of 1 (syn):25 (anti) (Figure 2).
The enantiomeric excess was determined by HPLC-UV. To establish the optimal separation conditions for the anti enantiomers, a sample of the racemic mixtures without and anti, obtained when the standard reaction was carried out using pyrrolidine as catalyst, was first prepared. Optimum separation conditions were obtained when an OD-H column was used in a hexane/isopropanol mixture (8:2) and a flow rate of 0.5 mL/min, as indicated in the chromatogram shown in Figure 2, for 3 min, corresponding to the mixture of enantiomers of the aldols, a peak at 34.6 min due to aldol (2S-4R)-10 and a peak at 38.3 min due to aldol (2R-4S)-12. Peak assignments were carried out by comparison with the literature [9].
Using these separation conditions, the chromatogram obtained for the reaction mixture resulting from carrying out the standard reaction with organocatalyst 9 showed that the major product of the reaction was aldol (2S,4R)-10, obtained with an enantiomeric excess greater than 99% (Figure 3). In addition, the optical rotation value obtained [+10.4° (c 4.95, CHCl3)] for the mixture also showed that the major compound corresponded to the dextrorotatory aldol (2S,4R)-10.

3. Conclusions

Therefore, if we compare the results obtained with our catalyst 9 with those obtained with proline 3 reported in the literature, it can be observed that the yield (86%) and the enantiomeric excess (>99%) are identical, with a diastereomeric ratio of 25:1 versus 20:1 (Table 1). This allows us to establish that the presence of an additional substituent at the C-3 position of proline or the opposite configuration at C-4 does not have a major influence on this reaction, beyond slightly increasing the diastereoselectivity.

Author Contributions

Conceptualization, R.J.E. and J.C.E.; methodology R.J.E. and J.C.E.; investigation, R.B. and A.F.; resources, J.C.E.; writing—original draft preparation, R.J.E.; writing—review and editing, J.C.E.; supervision, J.C.E.; project administration, J.C.E.; funding acquisition, J.C.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union (European Regional Development Fund-ERDF), the Xunta de Galicia (Centro Singular de Investigación de Galicia acreditation 2019–2022, ED431G 2019/03, and grants ED431C 2018/30 and ED431C 2018/04).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data not available, due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Eder, U.; Sauer, G.; Wiechert, R. New type of asymmetric cyclization to optically active steroid CD partial structures. Angew. Chem. Int. Ed. 1971, 10, 496–497. [Google Scholar] [CrossRef]
  2. Hajos, Z.G.; Parrish, D.R. Asymmetric synthesis of bicyclic intermediates of natural product chemistry. J. Org. Chem. 1974, 39, 1615–1621. [Google Scholar] [CrossRef]
  3. List, B.; Lerner, R.A.; Barbas, C.F., III. Proline-catalyzed direct asymmetric aldol reactions. J. Am. Chem. Soc. 2000, 122, 2395–2396. [Google Scholar] [CrossRef]
  4. Ahrendt, K.A.; Borths, C.J.; MacMillan, D.W.C. New strategies for organic catalysis: The first highly enantioselective organocatalytic Diels−Alder reaction. J. Am. Chem. Soc. 2000, 122, 4243–4244. [Google Scholar] [CrossRef]
  5. Trost, B.M.; Brindle, C.S. The direct catalytic asymmetric aldol reaction. Chem. Soc. Rev. 2010, 39, 1600–1632. [Google Scholar] [CrossRef] [PubMed]
  6. Vishumaya, M.R.; Singh, V.K. Highly efficient small organic molecules for enantioselective direct aldol reaction in organic and aqueous media. J. Org. Chem. 2009, 74, 4289–4297. [Google Scholar] [CrossRef] [PubMed]
  7. Hayashi, Y.; Sumiya, T.; Takahashi, J.; Gotoh, H.; Urushima, T.; Shoji, M. Highly diastereo-and enantioselective direct aldol reactions in water. Angew. Chem. Int. Ed. 2006, 45, 958–961. [Google Scholar] [CrossRef] [PubMed]
  8. Aratake, S.; Itoh, T.; Okano, T.; Nagae, N.; Sumiya, T.; Shoji, M.; Hayashi, Y. Highly Diastereo- and Enantioselective Direct Aldol Reactions of Aldehydes and Ketones Catalyzed by Siloxyproline in the Presence of Water. Chemistry 2007, 13, 10246–10256. [Google Scholar] [CrossRef] [PubMed]
  9. Córdova, A.; Zou, W.; Dziedzic, P.; Ibrahem, I.; Reyes, E.; Xu, Y. Direct asymmetric intermolecular aldol reactions catalyzed by amino acids and small peptides. Chem. Eur. J. 2006, 12, 5383–5397. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Aldol reaction catalyzed by proline 3, in an aqueous medium.
Scheme 1. Aldol reaction catalyzed by proline 3, in an aqueous medium.
Chemproc 14 00107 sch001
Scheme 2. Aldol reaction catalyzed by proline 3 in an aqueous medium.
Scheme 2. Aldol reaction catalyzed by proline 3 in an aqueous medium.
Chemproc 14 00107 sch002
Scheme 3. Synthesis of proline 9.
Scheme 3. Synthesis of proline 9.
Chemproc 14 00107 sch003
Scheme 4. Aldol reaction catalyzed by proline 9, in an aqueous medium.
Scheme 4. Aldol reaction catalyzed by proline 9, in an aqueous medium.
Chemproc 14 00107 sch004
Figure 1. 1H NMR of the aldolic mixture 10-13.
Figure 1. 1H NMR of the aldolic mixture 10-13.
Chemproc 14 00107 g001
Figure 2. Chromatogram of racemic mixtures of the aldols 1013.
Figure 2. Chromatogram of racemic mixtures of the aldols 1013.
Chemproc 14 00107 g002
Figure 3. Chromatogram of the aldol mixture of the standard reaction with catalyst 9.
Figure 3. Chromatogram of the aldol mixture of the standard reaction with catalyst 9.
Chemproc 14 00107 g003
Table 1. Results obtained for catalyst 3 and 9.
Table 1. Results obtained for catalyst 3 and 9.
CatalystOverall Yielde.e.Anti/syn RatioMajor Reaction Product
386%>99%20:110
986%>99%25:110
Comparison of the results for compounds 3 and 9.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Estévez, R.J.; Balo, R.; Fernández, A.; Estévez, J.C. Organocatalytic Properties of 3,4-Dihydroxyprolines. Chem. Proc. 2023, 14, 107. https://doi.org/10.3390/ecsoc-27-16117

AMA Style

Estévez RJ, Balo R, Fernández A, Estévez JC. Organocatalytic Properties of 3,4-Dihydroxyprolines. Chemistry Proceedings. 2023; 14(1):107. https://doi.org/10.3390/ecsoc-27-16117

Chicago/Turabian Style

Estévez, Ramón J., Rosalino Balo, Andrés Fernández, and Juan C. Estévez. 2023. "Organocatalytic Properties of 3,4-Dihydroxyprolines" Chemistry Proceedings 14, no. 1: 107. https://doi.org/10.3390/ecsoc-27-16117

Article Metrics

Back to TopTop