Next Article in Journal
Phytochemical Compounds and Nanoparticles as Phytochemical Delivery Systems for Alzheimer’s Disease Management
Next Article in Special Issue
Synthesis and Application Dichalcogenides as Radical Reagents with Photochemical Technology
Previous Article in Journal
The Impact of Software Used and the Type of Target Protein on Molecular Docking Accuracy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Facile Synthesis of 2-Oxazolines via Dehydrative Cyclization Promoted by Triflic Acid

1
Institute of Drug Discovery Technology, Ningbo University, Ningbo 315211, China
2
Qian Xuesen Collaborative Research Center of Astrochemistry and Space Life Sciences, Ningbo University, Ningbo 315211, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(24), 9042; https://doi.org/10.3390/molecules27249042
Submission received: 29 October 2022 / Revised: 10 December 2022 / Accepted: 15 December 2022 / Published: 19 December 2022
(This article belongs to the Special Issue Synthesis and Modification of Nitrogen Heterocyclic Compounds)

Abstract

:
2-oxazolines are common moieties in numerous natural products, pharmaceuticals, and functional copolymers. Current methods for synthesizing 2-oxazolines mainly rely on stoichiometric dehydration agents or catalytic dehydration promoted by specific catalysts. These conditions either generate stoichiometric amounts of waste or require forcing azeotropic reflux conditions. As such, a practical and robust method that promotes dehydrative cyclization while generating no byproducts would be attractive to oxazoline production. Herein, we report a triflic acid (TfOH)-promoted dehydrative cyclization of N-(2-hydroxyethyl)amides for synthesizing 2-oxazolines. This reaction tolerates various functional groups and generates water as the only byproduct. This method affords oxazoline with inversion of α-hydroxyl stereochemistry, suggesting that alcohol is activated as a leaving group under these conditions. Furthermore, the one-pot synthesis protocol of 2-oxazolines directly from carboxylic acids and amino alcohols is also provided.

1. Introduction

2-oxazoline is a privileged structural motif in numerous bioactive molecules and pharmaceuticals [1,2,3,4,5,6,7] (Figure 1a) as well as functional copolymers [8,9,10,11,12,13,14,15] (Figure 1c). Various natural products and synthetic molecules that contain this structural unit possess biological activities, such as antibiotics [16,17], antineoplastics [18,19,20], anti-fungals [21], and anti-inflammatories [22], among others. Furthermore, 2-oxazolines have a wide range of synthetic applications, including protective groups for carboxylic acid and aldehyde, directing groups in C-H functionalization, and valuable chiral Box and Pybox ligands [23,24,25,26] (Figure 1b). These important applications have fueled the development of various approaches to the efficient construction of 2-oxazolines over the last few decades. The typical approaches involve coupling amino alcohols with carboxylic acid derivatives [27,28,29,30,31], nitriles [32,33,34], and aldehydes [35,36,37] in the presence of activation reagents, catalysts, or oxidants. Recently, the functionalization of alkenes with amides provided a valuable alternative approach to the 2-oxazoline synthesis [38,39,40,41,42]. Although these advances expanded the chemist’s toolbox for 2-oxazoline synthesis, developing practical and cost-effective new methods for constructing 2-oxazolines would complement current methods.
Despite significant advances in 2-oxazoline synthesis, dehydrative cyclization of N-(β-hydroxyethyl)amides remains the most widely used method for producing 2-oxazolines. Numerous stoichiometric reagents, including DAST, XtalFluor-E, PPE, Ph3P/DEAD, and Burgess reagent, have proven to be efficient at forging the oxazoline moiety [43,44,45,46,47,48,49,50,51,52,53] (Figure 2a). These conditions generally require either harsh conditions or corrosive reagents, which may cause additional operating costs and stoichiometric byproduct generation. To address this issue, several groups have developed catalytic dehydrative approaches [54,55,56,57]. The Ishihara group reported a molybdenum complex-catalyzed dehydrative cyclization of N-(2-hydroxyethyl)amides [54,55]. Saito and co-workers demonstrated a phosphorus-based organocatalytic dehydrative cyclization approach [57] (Figure 2a). In addition, one example of cyclization catalyzed by sulfuric acid was also reported but under harsh high-temperature conditions [58]. While these methods avoid using stoichiometric dehydration agents and thus have higher atom economy, the requirement for specific catalysts and forcing azeotropic reflux conditions might limit their industrial application. As a result, a practical and robust method that promotes dehydrative cyclization while generating no byproducts would be attractive to oxazoline production. Herein, we report our effort in the TfOH-promoted synthesis of 2-oxazolines by dehydrative cyclization of N-(2-hydroxyethyl)amides (Figure 2b).

2. Results

2.1. Optimization of the Reaction Conditions

We began our reaction optimization by examining the cyclization reaction of β-hydroxyamide 1 in the presence of several organic acids in 1,2-dichloroethane (DCE) (Table 1, entries 1–3). It was found that TfOH in DCE at 80 °C effectively promoted the formation of the desired 2-oxazoline. The acidity of the acid seemed to be important, as weaker acids such as MsOH and TFA only afforded product in low yields (Table 1, entries 1–2). Stoichiometry optimization on acid (Table 1, entries 4–8) revealed that a 1.5 equivalent of TfOH was optimal (Table 1, entry 7). Several other solvents (Table 1, entries 9–11) gave similar results albeit in a slightly lower yield than DCE, suggesting no significant solvent effect for this transformation. In addition, running the reaction at lower temperatures afforded product 2 in lower yields (Table 1, entries 12–14).

2.2. Substrate Scope Studies

With the optimized reaction conditions in hand, we then investigated the generality of this protocol. We initially tested a range of substrates derived from monosubstituted benzoic acid and ethanolamine. Functional groups, such as halides, ether, ester, CF3, and nitro, were well tolerated in standard reaction conditions and afforded the desired products in good to excellent yields (Figure 3, products 311). Although generally unstable under acidic conditions in the presence of water, the substrate with the cyano group also gave product albeit in a lower yield. It appears that the steric hindrance had a minimal impact on the reactivity as evident by the similar yield observed in the reaction of the sterically hindered substrates (Figure 3, products 12 and 13). N-(2-hydroxyethyl)amides derived from 2-thiophenecarboxylic acid and 2-furoic acid were also viable substrates, delivering the desired products 14 and 15 in 96% and 73% yield, respectively. N-(2-hydroxyethyl)amides derived from secondary and tertiary aliphatic acids proceeded smoothly under standard conditions affording the desired 2-oxazolines with moderate to good yields (Figure 3, products 1618). We then turned our attention to exploring the substrates derived from β-substituted 1,2-amino alcohols. The substrates derived from L-valinol, L-tert-Leucinol, L-Leucinol, D-Phenylglycinol, 2-amino-2-methyl-1-propanol, and D-serine methyl ester were all viable substrates and delivered the desired products in good to excellent yields (Figure 3, products 1924). Moreover, the substrates derived from (S)-(+)-1-Amino-2-propanol, L-Threonine methyl ester and (1S, 2R)-(−)-cis-1-amino-2-indanol that bear α-substitution, and α, β-disubstitution were also well tolerated in this protocol (Figure 3, products 2527). Notably, products 26 and 27 were isolated as a single diastereomer, and no other diastereomers were detected from crude NMR. Mechanistic studies suggested that products 25 and 26 were formed with an inversion of the stereochemistry at carbon β. Depending on the starting material, a product with a rigid backbone such as 27 can be generated with either inversion or retention of the stereochemistry at position β. In addition, 1,3-amino alcohol derivative afforded 5,6-dihydro-4H-1,3-oxazine in moderate yields (Figure 3, product 28).
Given the robustness of this practical protocol, we envisioned the possibility of a one-pot synthesis of 2-oxazolines directly from the carboxylic acid and 1,2-amino alcohols. To construct a TfOH-friendly system, we tested the base-free ynamide invented by Zhao [59] as a coupling reagent. A variety of oxazolines were successfully synthesized in a one-pot fashion via in situ coupling of carboxylic acids with amino alcohols followed by cyclization under standard conditions. (Figure 4, products 2932, 18).

2.3. Control Experiments and Mechanistic Studies

According to previous reports, this reaction has two possible pathways that result in products with opposite stereochemical outcomes. One pathway involves acid activation of the amide carbonyl group followed by nucleophilic attack of the hydroxyl group resulting in 2-oxazoline with retention of stereochemistry (Figure 5a, pathway A). Alcohol activation followed by intramolecular SN2-like substitution, on the other hand, would produce cyclized products with reversed a-hydroxyl stereochemistry (Figure 5a, pathway B). We then conducted several control experiments to study the reaction mechanism. Our studies started from treating sterically rigid cis-β-hydroxyl amide 34 and trans-β-hydroxyl amide 35 with standard conditions to probe the possible reaction pathway (Figure 5a). Surprisingly, the formation of product 27 was observed in both cases, suggesting that both pathways are operatable under standard conditions. While the higher yield obtained from 35 suggested that pathway B might be more favored, more information is required to gain a better understanding of the mechanism. We then subjected enantiopure β-hydroxyl amide 36 to the reaction conditions and analyzed the stereoselectivity using chiral HPLC (Figure 5b). 2-oxazoline 25 was obtained with stereochemical inversion as the major isomer (94:6 e.r.), which indicates that the pathway involving alcohol activation is more favored. We think that the erosion of optical purities observed in product 25 might result from a hybrid reaction pathway. To validate this hypothesis, we conducted the 18O labeling experiment. N-(2-hydroxyethyl)amides 37 with 95% 18O enrichment was smoothly converted to product, and the ratio of 18O-19 and 19 was 83:17 (Figure 5c). These data are consistent with the hypothesis of a hybrid mechanism, in which activation of the hydroxyl group is the dominant pathway under our reaction condition.

3. Conclusions

In conclusion, a practical and effective strategy for synthesizing 2-oxazolines via dehydrative cyclization of N-(2-hydroxyethyl)amides has been developed. This efficient cyclization process was promoted by TfOH and had good functional group tolerance. Stereoselectivity and 18O labeling data suggested that the reaction might proceed through a hybrid mechanism, in which activation of the hydroxyl group is the dominant pathway. Notably, this robust reaction condition can be adapted to a one-pot reaction by directly utilizing readily available carboxylic acid and amino alcohols.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27249042/s1, Table S1: Reaction optimization; Figure S1: Control experiments on stereochemical outcome of C(4) position; Figure S2: The HPLC analysis of the (S)-oxazoline 19 and (R)-oxazoline 19; Figure S3: Cis- and trans- 1-amino-2-indanol derived mechanistic probe; Figure S4: Stereochemical outcome of this protocol; Figure S5: HPLC analysis of oxazoline product 25; Figure S6: 18O-labeling study of product 19; Figure S7: HRMS data of the 18O-labeled N-(2-hydroxyethyl)amides and oxazoline 19.

Author Contributions

Conceptualization, F.W. and F.N.; methodology, F.W.; formal analysis, T.Y., C.H. and J.J.; investigation, T.Y., C.H. and J.J.; writing—original draft preparation, F.W. and T.Y.; writing—review and editing, F.W. and F.N.; funding acquisition, F.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (91856126, 21778042), Scientific Research Grant of Ningbo University (215-432000282), and Ningbo Top Talent Project (215-432094250).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Inahashi, Y.; Iwatsuki, M.; Ishiyama, A.; Namatame, M.; Nishihara-Tsukashima, A.; Matsumoto, A.; Hirose, T.; Sunazuka, T.; Yamada, H.; Otoguro, K.; et al. Spoxazomicins A–C, Novel Antitrypanosomal Alkaloids Produced by an Endophytic Actinomycete, Streptosporangium oxazolinicum K07-0460T. J. Antibiot. 2011, 64, 303–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nelson, K.M.; Salomon, C.E.; Aldrich, C.C. Total Synthesis and Biological Evaluation of Transvalencin Z. J. Nat. Prod. 2012, 75, 1037–1043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Marson, C.M.; Matthews, C.J.; Atkinson, S.J.; Lamadema, N.; Thomas, N.S.B. Potent and Selective Inhibitors of Histone Deacetylase-3 Containing Chiral Oxazoline Capping Groups and a N-(2-Aminophenyl)-benzamide Binding Unit. J. Med. Chem. 2015, 58, 6803–6818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tyler, A.R.; Mosaei, H.; Morton, S.; Waddell, P.G.; Wills, C.; McFarlane, W.; Gray, J.; Goodfellow, M.; Errington, J.; Allenby, N.; et al. Structural Reassignment and Absolute Stereochemistry of Madurastatin C1 (MBJ-0034) and the Related Aziridine Siderophores: Madurastatins A1, B1, and MBJ-0035. J. Nat. Prod. 2017, 80, 1558–1562. [Google Scholar] [CrossRef] [PubMed]
  5. Shaaban, K.A.; Saunders, M.A.; Zhang, Y.; Tran, T.; Elshahawi, S.I.; Ponomareva, L.V.; Wang, X.; Zhang, J.; Copley, G.C.; Sunkara, M.; et al. Spoxazomicin D and Oxachelin C, Potent Neuroprotective Carboxamides from the Appalachian Coal Fire-Associated Isolate Streptomyces sp. RM-14-6. J. Nat. Prod. 2017, 80, 2–11. [Google Scholar] [CrossRef] [Green Version]
  6. Mohr, J.F.; Baldeweg, F.; Deicke, M.; Morales-Reyes, C.F.; Hoffmeister, D.; Wichard, T. Frankobactin Metallophores Produced by Nitrogen-Fixing Frankia Actinobacteria Function in Toxic Metal Sequestration. J. Nat. Prod. 2021, 84, 1216–1225. [Google Scholar] [CrossRef]
  7. Chen, S.; Zhang, Y.; Liu, Y.; Wang, Q. Highly Efficient Synthesis and Acaricidal and Insecticidal Activities of Novel Oxazolines with N-Heterocyclic Substituents. J. Agric. Food Chem. 2021, 69, 3601–3606. [Google Scholar] [CrossRef]
  8. Luxenhofer, R.; Han, Y.; Schulz, A.; Tong, J.; He, Z.; Kabanov, A.V.; Jordan, R. Poly(2-oxazoline)s as Polymer Therapeutics. Macromol. Rapid Commun. 2012, 33, 1613–1631. [Google Scholar] [CrossRef] [Green Version]
  9. Salgarella, A.R.; Zahoranová, A.; Srámková, P.; Majerčíková, M.; Pavlova, E.; Luxenhofer, R.; Kronek, J.; Lacík, I.; Ricotti, L. Investigation of Drug Release Modulation from Poly(2-oxazoline) Micelles through Ultrasound. Sci. Rep. 2018, 8, 9893–9906. [Google Scholar] [CrossRef] [Green Version]
  10. Sedlacek, O.; Lava, K.; Verbraeken, B.; Kasmi, S.; De Geest, B.G.; Hoogenboom, R. Unexpected Reactivity Switch in the Statistical Copolymerization of 2-Oxazolines and 2-Oxazines Enabling the One-Step Synthesis of Amphiphilic Gradient Copolymers. J. Am. Chem. Soc. 2019, 141, 9617–9622. [Google Scholar] [CrossRef]
  11. Wu, Y.-C.M.; Swager, T.M. Living Polymerization of 2-Ethylthio-2-oxazoline and Postpolymerization Diversification. J. Am. Chem. Soc. 2019, 141, 12498–12501. [Google Scholar] [CrossRef] [Green Version]
  12. Sahn, M.; Weber, C.; Schubert, U.S. Poly(2-oxazoline)-Containing Triblock Copolymers: Synthesis and Applications. Polym. Rev. 2019, 59, 240–279. [Google Scholar] [CrossRef]
  13. Park, J.-R.; Sarwat, M.; Bolle, E.C.L.; de Laat, M.A.; Van Guyse, J.F.R.; Podevyn, A.; Hoogenboom, R.; Dargaville, T.R. Drug–polymer Conjugates with Dynamic Cloud Point Temperatures Based on Poly(2-oxazoline) Copolymers. Polym. Chem. 2020, 11, 5191–5199. [Google Scholar] [CrossRef]
  14. Sedlacek, O.; Hoogenboom, R. Drug Delivery Systems Based on Poly(2-Oxazoline)s and Poly(2-Oxazine)s. Adv. Ther. 2020, 3, 1900168. [Google Scholar] [CrossRef] [Green Version]
  15. Zhou, M.; Qian, Y.; Xie, J.; Zhang, W.; Jiang, W.; Xiao, X.; Chen, S.; Dai, C.; Cong, Z.; Ji, Z.; et al. Poly(2-Oxazoline)-Based Functional Peptide Mimics: Eradicating MRSA Infections and Persisters while Alleviating Antimicrobial Resistance. Angew. Chem. Int. Ed. 2020, 59, 6412–6419. [Google Scholar] [CrossRef] [PubMed]
  16. Kline, T.; Andersen, N.H.; Harwood, E.A.; Bowman, J.; Malanda, A.; Endsley, S.; Erwin, A.L.; Doyle, M.; Fong, S.; Harris, A.L.; et al. Potent, Novel in Vitro Inhibitors of the Pseudomonas aeruginosa Deacetylase LpxC. J. Med. Chem. 2002, 45, 3112–3129. [Google Scholar] [CrossRef]
  17. Pirrung, M.C.; Tumey, L.N.; McClerren, A.L.; Raetz, C.R.H. High-Throughput Catch-and-Release Synthesis of Oxazoline Hydroxamates. Structure–Activity Relationships in Novel Inhibitors of Escherichia coli LpxC: In Vitro Enzyme Inhibition and Antibacterial Properties. J. Am. Chem. Soc. 2003, 125, 1575–1586. [Google Scholar] [CrossRef]
  18. Carmeli, S.; Moore, R.E.; Patterson, G.M.L.; Corbett, T.H.; Valeriote, F.A. Tantazoles, Unusual Cytotoxic Alkaloids from the Blue-green Alga Scytonema Mirabile. J. Am. Chem. Soc. 1990, 112, 8195–8197. [Google Scholar] [CrossRef]
  19. Prinsep, M.R.; Moore, R.E.; Levine, I.A.; Patterson, G.M.L. Westiellamide, a Bistratamide-Related Cyclic Peptide from the Blue-Green Alga Westiellopsis prolifica. J. Nat. Prod. 1992, 55, 140–142. [Google Scholar] [CrossRef]
  20. Kim, M.Y.; Vankayalapati, H.; Shin-ya, K.; Wierzba, K.; Hurley, L.H.J. Telomestatin, a Potent Telomerase Inhibitor That Interacts Quite Specifically with the Human Telomeric Intramolecular G-Quadruplex. J. Am. Chem. Soc. 2002, 124, 2098–2099. [Google Scholar] [CrossRef]
  21. Bode, B.H.; Irschik, H.; Wenzel, S.C.; Reichenbach, H.; Müller, R.; Höfle, G. The Leupyrrins: A Structurally Unique Family of Secondary Metabolites from the Myxobacterium Sorangium cellulosum. J. Nat. Prod. 2003, 66, 1203–1206. [Google Scholar] [CrossRef] [PubMed]
  22. Nicolaou, K.C.; Lizos, D.E.; Kim, D.W.; Schlawe, D.; de Noronha, R.G.; Longbottom, D.A.; Rodriquez, M.; Bucci, M.; Cirino, G. Total Synthesis and Biological Evaluation of Halipeptins A and D and Analogues. J. Am. Chem. Soc. 2006, 128, 4460–4470. [Google Scholar] [CrossRef] [PubMed]
  23. Gant, T.G.; Meyers, A.I. The Chemistry of 2-oxazolines (1985–present). Tetrahedron 1994, 50, 2297–2360. [Google Scholar] [CrossRef]
  24. Byrne, C.M.; Church, T.L.; Kramer, J.W.; Coates, G.W. Catalytic Synthesis of β3-Amino Acid Derivatives from α-Amino Acids. Angew. Chem. Int. Ed. 2008, 47, 3979–3983. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, K.; Li, Z.-W.; Shen, P.-X.; Zhao, H.-W.; Shi, Z.-J. Development of Modifiable Bidentate Amino Oxazoline Directing Group for Pd-Catalyzed Arylation of Secondary C-H Bonds. Chem. Eur. J. 2015, 21, 7389–7393. [Google Scholar] [CrossRef]
  26. Shang, M.; Wang, M.-M.; Saint-Denis, T.G.; Li, M.-H.; Dai, H.-X.; Yu, J.-Q. Copper-Mediated Late-Stage Functionalization of Heterocycle-Containing Molecules. Angew. Chem. Int. Ed. 2017, 56, 5317–5321. [Google Scholar] [CrossRef] [PubMed]
  27. Vorbrüggen, H.; Krolikiewicz, K. A Simple Synthesis of Δ2-oxazines, Δ2-oxazines, Δ2-thiazolines and 2-Substituted Benzoxazoles. Tetrahedron 1993, 49, 9353–9372. [Google Scholar] [CrossRef]
  28. Marrero-Terrero, A.L.; Loupy, A. Synthesis of 2-Oxazolines from Carboxylic Acids and α, α, α-Tris(hydroxymethyl)methylamine under Microwaves in Solvent-Free Conditions. Synlett 1996, 1996, 245–246. [Google Scholar] [CrossRef]
  29. Kangani, C.O.; Kelley, D.E. One Pot Direct Synthesis of Amides or Oxazolines from Carboxylic Acids Using Deoxo-Fluor Reagent. Tetrahedron Lett. 2005, 46, 8917–8920. [Google Scholar] [CrossRef] [Green Version]
  30. Zhou, P.; Blubaum, J.E.; Burns, C.T.; Natale, N.R. The Direct Synthesis of 2-Oxazolines from Carboxylic Esters using Lanthanide Chloride as Catalyst. Tetrahedron Lett. 1997, 38, 7019–7020. [Google Scholar] [CrossRef]
  31. Zhou, H.; Zeng, X.; Ding, L.; Xie, Y.; Zhong, G. Triflic Acid Catalyzed Formal [3 + 2] Cycloaddition of Donor–Acceptor Oxiranes and Nitriles: A Facile Access to 3-Oxazolines. Org. Lett. 2015, 17, 2385–2387. [Google Scholar] [CrossRef] [PubMed]
  32. Bolm, C.; Weickhardt, K.; Zehnder, M.; Ranff, T. Synthesis of Optically Active Bis(2-oxazolines): Crystal Structure of a 1,2-Bis(2-oxazolinyl)benzene ZnCl2 Complex. Chem. Ber. 1991, 124, 1173–1180. [Google Scholar] [CrossRef]
  33. Mohammadpoor-Baltork, I.; Khosropour, A.R.; Hojati, S.F. A Novel and Chemoselective Synthesis of 2-Aryloxazolines and Bis-oxazolines Catalyzed by Bi(III) Salts. Synlett 2005, 18, 2747–2750. [Google Scholar] [CrossRef]
  34. Cai, A.-J.; Zheng, Y.; Ma, J.-A. Copper-Triggered Three-Component Reaction of CF3CHN2, Nitriles, and Aldehydes: Highly Diastereoselective Synthesis of CF3-substituted Oxazolines and Vicinal Amino Alcohols. Chem. Commun. 2015, 51, 8946–8949. [Google Scholar] [CrossRef] [PubMed]
  35. Hayashi, T.; Sawamura, M.; Ito, Y. Asymmetric Synthesis Catalyzed by Chiral Ferrocenyl Phosphine Transition Metal Complexes. 10 gold(i)-Catalyzed Asymmetric Aldol Reaction of Isocyanoacetate. Tetrahedron 1992, 48, 1999–2012. [Google Scholar] [CrossRef]
  36. Badiang, J.G.; Aubé, J. One-Step Conversion of Aldehydes to Oxazolines and 5,6-Dihydro-4H-1,3-oxazines Using 1,2- and 1,3-Azido Alcohols. J. Org. Chem. 1996, 61, 2484–2487. [Google Scholar] [CrossRef]
  37. Ishihara, M.; Togo, H. Direct Oxidative Conversion of Aldehydes and Alcohols to 2-imidazolines and 2-oxazolines Using Molecular Iodine. Tetrahedron 2007, 63, 1474–1480. [Google Scholar] [CrossRef]
  38. Minakata, S.; Nishimura, M.; Takahashi, T.; Oderaotoshi, Y.; Komatsu, M. Direct Asymmetric Synthesis of Oxazolines from Olefins Using a Chiral Nitridomanganese Complex: A Novel Three-component Coupling Leading to Chiral Oxazolines. Tetrahedron Lett. 2001, 42, 9019–9022. [Google Scholar] [CrossRef]
  39. Minakata, S.; Morino, Y.; Ide, T.; Oderaotoshi, Y.; Komatsu, M. Direct Synthesis of Oxazolines from Olefins and Amides Using t-BuOI. Chem. Commun. 2007, 31, 3279–3281. [Google Scholar] [CrossRef]
  40. Wu, F.; Alom, N.-E.; Ariyarathna, J.P.; Naβ, J.; Li, W. Regioselective Formal [3+2] Cycloadditions of Urea Substrates with Activated and Unactivated Olefins for Intermolecular Olefin Aminooxygenation. Angew. Chem. Int. Ed. 2019, 58, 11676–11680. [Google Scholar] [CrossRef]
  41. Wu, F.; Kaur, N.; Alom, N.-E.; Li, W. Chiral Hypervalent Iodine Catalysis Enables an Unusual Regiodivergent Intermolecular Olefin Aminooxygenation. JACS Au 2021, 1, 734–741. [Google Scholar] [CrossRef] [PubMed]
  42. Mumford, E.M.; Hemric, B.N.; Denmark, S.E. Catalytic, Enantioselective Syn-Oxyamination of Alkenes. J. Am. Chem. Soc. 2021, 143, 13408–13417. [Google Scholar] [CrossRef]
  43. Roush, D.M.; Patel, M.M. A Mild Procedure for the Preparation of 2-oxazolines. Synth. Commun. 1985, 15, 675–679. [Google Scholar] [CrossRef]
  44. Galéotti, N.; Montagne, C.; Poncet, J.; Jouin, P. Formation of Oxazolines and Thiazolines in Peptides by the Mitsunobu Reaction. Tetrahedron Lett. 1992, 33, 2807–2810. [Google Scholar] [CrossRef]
  45. Wipf, P.; Fritch, P.C. Total Synthesis and Assignment of Configuration of Lissoclinamide 7. J. Am. Chem. Soc. 1996, 118, 12358–12367. [Google Scholar] [CrossRef]
  46. Phillips, A.J.; Uto, Y.; Wipf, P.; Reno, M.J.; Williams, D.R. Synthesis of Functionalized Oxazolines and Oxazoles with DAST and Deoxo-Fluor. Org. Lett. 2000, 2, 1165–1168. [Google Scholar] [CrossRef]
  47. Plant, A.; Stieber, F.; Scherkenbeck, J.; Losel, P.; Dyker, H. Syntheses of Analogues of the Insect Neuropeptide Proctolin Containing an Oxazole Ring as an Amide Bond Replacement. Org. Lett. 2001, 3, 3427–3430. [Google Scholar] [CrossRef]
  48. Doi, T.; Yoshida, M.; Shin-ya, K.; Takahashi, T. Total Synthesis of (R)-Telomestatin. Org. Lett. 2006, 8, 4165–4167. [Google Scholar] [CrossRef]
  49. Ying, Y.; Hong, J. Synthesis of Brasilibactin A and Confirmation of Absolute Configuration of β-hydroxy Acid Fragment. Tetrahedron Lett. 2007, 48, 8104–8107. [Google Scholar] [CrossRef]
  50. Liyanage, W.; Weerasinghe, L.; Strong, R.K.; Del Valle, J.R. Synthesis of Carbapyochelins via Diastereoselective Azidation of 5-(ethoxycarbonyl) Methylproline Derivatives. J. Org. Chem. 2008, 73, 7420–7423. [Google Scholar] [CrossRef] [PubMed]
  51. Castellano, S.; Kuck, D.; Sala, M.; Novellino, E.; Lyko, F.; Sbardella, G. Constrained Analogues of Procaine as Novel Small Molecule Inhibitors of DNA Methyltransferase-1. J. Med. Chem. 2008, 51, 2321–2325. [Google Scholar] [CrossRef]
  52. Pouliot, M.; Angers, L.; Hamel, J.; Paquin, J. Synthesis of 2-oxazolines and Related N-containing Heterocycles Using [Et2NSF2] BF4 as a Cyclodehydration Agent. Tetrahedron Lett. 2012, 53, 4121–4123. [Google Scholar] [CrossRef]
  53. Mollo, M.C.; Orelli, L.R. Microwave-Assisted Synthesis of 2-Aryl-2-oxazolines, 5,6-Dihydro-4H-1,3-oxazines, and 4,5,6,7-Tetrahydro-1,3-oxazepines. Org. Lett. 2016, 18, 6116–6119. [Google Scholar] [CrossRef] [PubMed]
  54. Sakakura, A.; Kondo, R.; Ishihara, K. Molybdenum Oxides as Highly Effective Dehydrative Cyclization Catalysts for the Synthesis of Oxazolines and Thiazolines. Org. Lett. 2005, 7, 1971–1974. [Google Scholar] [CrossRef]
  55. Sakakura, A.; Umemura, S.; Kondo, R.; Ishihara, K. Dehydrative Cyclization Catalyzed by the Combination of Molybdenum (VI) Oxides and Benzoic Acids: First Synthesis of the Antitumour Substance BE-70016. Adv. Synth. Catal. 2007, 349, 551–555. [Google Scholar] [CrossRef]
  56. Rodriguez del Rey, F.O.; Floreancig, P.E. Synthesis of Nitrogen-Containing Heterocycles through Catalytic Dehydrative Cyclization Reactions. Org. Lett. 2021, 23, 150–154. [Google Scholar] [CrossRef]
  57. Movahed, F.S.; Foo, S.W.; Mori, S.; Ogawa, S.; Saito, S. Phosphorus-Based Organocatalysis for the Dehydrative Cyclization of N-(2-hydroxyethyl)amides into 2-Oxazolines. J. Org. Chem. 2022, 87, 243–257. [Google Scholar] [CrossRef] [PubMed]
  58. de Benneville, P.L.; Luskin, L.S.; Sims, H.J. Transesterification of Methyl Methacrylate with Amino Alcohols. Preparation of a Primary Aminoalkyl Methacrylate and 2-Isopropenyl-4,4-dimethyloxazoline. J. Org. Chem. 1958, 17, 1355–1357. [Google Scholar] [CrossRef]
  59. Hu, L.; Xu, S.; Zhao, Z.; Yang, Y.; Peng, Z.; Yang, M.; Wang, C.; Zhao, J. Ynamides as Racemization-free Coupling Reagents for Amide and Peptide Synthesis. J. Am. Chem. Soc. 2016, 138, 13135–13138. [Google Scholar] [CrossRef]
Figure 1. Functional molecules containing 2-oxazoline moiety.
Figure 1. Functional molecules containing 2-oxazoline moiety.
Molecules 27 09042 g001
Figure 2. Dehydrative cyclization of N-(β-hydroxyethyl)amides.
Figure 2. Dehydrative cyclization of N-(β-hydroxyethyl)amides.
Molecules 27 09042 g002
Figure 3. Scope of dehydrative cyclization reaction. Standard reaction conditions, 12 h. The yields shown here are isolated yield. b Running reaction at 60 °C instead of 80 °C.
Figure 3. Scope of dehydrative cyclization reaction. Standard reaction conditions, 12 h. The yields shown here are isolated yield. b Running reaction at 60 °C instead of 80 °C.
Molecules 27 09042 g003
Figure 4. Scope of one-pot synthesis of 2-oxazolines. See the Supplementary Materials for details.
Figure 4. Scope of one-pot synthesis of 2-oxazolines. See the Supplementary Materials for details.
Molecules 27 09042 g004
Figure 5. Mechanistic studies.
Figure 5. Mechanistic studies.
Molecules 27 09042 g005
Table 1. Optimization of the reaction conditions a.
Table 1. Optimization of the reaction conditions a.
Molecules 27 09042 i001
EntryAcid (Equiv.)SolventTemperatureYield b
1MsOH (1.0)DCE80 °C16
2TFA (1.0)DCE80 °C9
3TfOH (1.0)DCE80 °C89
4TfOH (0.2)DCE80 °C14
5TfOH (0.5)DCE80 °C29
6TfOH (1.2)DCE80 °C94
7TfOH (1.5)DCE80 °C96 (88) c
8TfOH (2.0)DCE80 °C86
9TfOH (1.5)Toluene80 °C92
10TfOH (1.5)PhCF380 °C86
11TfOH (1.5)CH3CN80 °C95
12TfOH (1.5)DCE70 °C91
13TfOH (1.5)DCE60 °C85
14TfOH (1.5)DCE25 °C<5
a Reaction conditions: N-(2-hydroxyethyl)amide 1 (0.2 mmol), acid (0.2–2.0 equiv) and solvent (1 mL) at 25–80 °C, t = 12h. b NMR yield using 1,3-benzodioxole as the internal standard; the NMR yield was calculated based on the ratio of CH2 signal (5.8 ppm) of 1,3-benzodioxole and CH signal of product 2 (4.5 ppm). c Isolated yield.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, T.; Huang, C.; Jia, J.; Wu, F.; Ni, F. A Facile Synthesis of 2-Oxazolines via Dehydrative Cyclization Promoted by Triflic Acid. Molecules 2022, 27, 9042. https://doi.org/10.3390/molecules27249042

AMA Style

Yang T, Huang C, Jia J, Wu F, Ni F. A Facile Synthesis of 2-Oxazolines via Dehydrative Cyclization Promoted by Triflic Acid. Molecules. 2022; 27(24):9042. https://doi.org/10.3390/molecules27249042

Chicago/Turabian Style

Yang, Tao, Chengjie Huang, Jingyang Jia, Fan Wu, and Feng Ni. 2022. "A Facile Synthesis of 2-Oxazolines via Dehydrative Cyclization Promoted by Triflic Acid" Molecules 27, no. 24: 9042. https://doi.org/10.3390/molecules27249042

Article Metrics

Back to TopTop