Next Article in Journal
Multiple Technology Approach Based on Stable Isotope Ratio Analysis, Fourier Transform Infrared Spectrometry and Thermogravimetric Analysis to Ensure the Fungal Origin of the Chitosan
Next Article in Special Issue
The Corey-Seebach Reagent in the 21st Century: A Review
Previous Article in Journal
Discovery of New 2-Phenylamino-3-acyl-1,4-naphthoquinones as Inhibitors of Cancer Cells Proliferation: Searching for Intra-Cellular Targets Playing a Role in Cancer Cells Survival
Previous Article in Special Issue
Fungal Secondary Metabolites/Dicationic Pyridinium Iodide Combinations in Combat against Multi-Drug Resistant Microorganisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organocatalytic Asymmetric Peroxidation of γ,δ-Unsaturated β-Keto Esters—A Novel Route to Chiral Cycloperoxides

by
Mary C. Hennessy
1,2,
Hirenkumar Gandhi
1,2 and
Timothy P. O’Sullivan
1,2,3,*
1
School of Chemistry, University College Cork, T12 YN60 Cork, Ireland
2
Analytical and Biological Chemistry Research Facility, University College Cork, T12 YN60 Cork, Ireland
3
School of Pharmacy, University College Cork, T12 YN60 Cork, Ireland
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(11), 4317; https://doi.org/10.3390/molecules28114317
Submission received: 19 April 2023 / Revised: 18 May 2023 / Accepted: 19 May 2023 / Published: 24 May 2023
(This article belongs to the Special Issue Recent Advances in Organic Synthesis Related to Natural Compounds)

Abstract

:
A methodology for the asymmetric peroxidation of γ,δ-unsaturated β-keto esters is presented. Using a cinchona-derived organocatalyst, the target δ-peroxy-β-keto esters were obtained in high enantiomeric ratios of up to 95:5. Additionally, these δ-peroxy esters can be readily reduced to chiral δ-hydroxy-β-keto esters without impacting the β-keto ester functionality. Importantly, this chemistry opens up a concise route to chiral 1,2-dioxolanes, a common motif in many bioactive natural products, via a novel P2O5-mediated cyclisation of the corresponding δ-peroxy-β-hydroxy esters.

1. Introduction

Over the past several decades, hundreds of peroxide-containing natural products have been isolated and characterized [1,2,3,4,5]. A large number of these peroxy compounds were subsequently found to exhibit potent activity against a wide range of diseases that impact both human and animal health [6,7]. For example, artemisinin (1) and its semi-synthetic derivatives are highly effective against Plasmodium falciparum, the causative agent of malaria (Figure 1) [8,9]. Many simpler chiral peroxides are also biologically interesting. Examples include 1,2-dioxolanes such as epiplakinidioic acid (2), obtained from a marine sponge, which inhibits both prostate and melanoma cancer cell growth [10,11]. Similarly, plakortide P (3), a 1,2-dioxane derivative, exhibits potent anti-neuroinflammatory effects and disrupts TXB2 production at submicromolar levels [5]. Cycloperoxides such as 2 and 3 often incorporate stereogenic centers adjacent to a carboxyl group, and this motif represents an attractive target in organic synthesis.
Despite their abundance in nature, methods for the enantioselective synthesis of chiral peroxides are still relatively scarce [12,13]. One approach with proven potential is the organocatalysed stereoselective peroxidation of unsaturated carbonyls. Asymmetric organocatalysis is an area of synthetic chemistry which is the focus of intense research, as evidenced by the many reviews on the topic since the early 2000s [14,15,16,17,18,19,20,21,22,23,24,25]. In particular, cinchona-derived organocatalysts have been successfully exploited by several groups for enantioselective peroxidations (Scheme 1) [26,27]. For example, List and Deng utilised a quinine-based catalyst for the asymmetric peroxidation of α,β-unsaturated ketones by the modification of a well-established epoxidation pathway utilising quinine-derived catalyst 4 [28,29,30]. Deng synthesised acyclic peroxides while List generated 1,2-dioxolanes and epoxides. More recently, Deng described the stereoselective peroxidation of α,β-unsaturated aldehydes [31]. In parallel work, our group investigated their asymmetric peroxidation with tert-butyl hydroperoxide (TBHP) followed by in situ Pinnick oxidation to afford stable β-peroxycarboxylic acids [32,33]. Outside of unsaturated ketone and aldehyde substrates, Russo and Lattanzi exploited a diaryl-2-pyrrolidinemethanol-derived catalyst to produce enantiomerically-enriched β-peroxynitroalkanes from nitrostyrenes [34]. The stereoselective peroxidation of unsaturated aromatic and aliphatic nitroalkenes has likewise been reported [35].
β-Keto esters constitute highly versatile building blocks in organic synthesis which can be readily modified through a wide variety of transformations [36,37,38,39,40,41,42]. To date, β-keto esters have been mostly employed as nucleophiles in organocatalysed reactions [43]. We are aware of only one report of a quinine-mediated addition of nitroalkanes to an α,β-unsaturated β-keto ester [44]. The main aim of this work is the development of a methodology for the asymmetric peroxidation of γ,δ-unsaturated β-keto esters (Scheme 2). This chemistry could open up a concise route to chiral 1,2-dioxolane-3-acetic acids, a recurring motif in several bioactive natural products.

2. Results and Discussion

Based on our prior work with α,β-unsaturated aldehydes, 9-amino cinchona derivative 4 was chosen as our catalyst along with pentafluorobenzoic acid as the co-catalyst. The choice of solvent is often a crucial aspect for achieving high enantioselectivities and yields in organocatalysed transformations. Accordingly, a solvent screen was conducted to identify the ideal reaction medium using 5 as our test substrate (Table 1). As the resulting chiral peroxides were inseparable by chiral chromatography, they were instead reduced to the corresponding δ-hydroxy-β-keto esters on reaction with zinc powder and acetic acid. This reduction strategy has been previously utilised for the analysis of β-peroxy ketones [28]. Reduction to more polar δ-hydroxy-β-keto esters affords products which are more amenable to resolution by chiral HPLC. A number of amylose and cellulose chiral columns were tested with the best results achieved with a Phenomenex AMY-2 column. An eluent system consisting of 5–15% isopropyl alcohol in hexane provided good separation. The stereochemical configuration of the major isomer was determined by comparison of the retention times with chiral δ-hydroxy-β-keto esters prepared by way of an asymmetric Mukaiyama aldol addition [45]. The (R)-enantiomer was found to predominate as previously observed with the peroxidation of α,β-unsaturated aldehydes and ketones [31,32]. Additionally, a negative optical rotation was recorded in each case, again supporting the formation of the (R)-isomer as the preferred product [45].
The least suitable solvent was acetonitrile, which resulted in a low yield accompanied by a poor e.r. of 69:31 (Table 1, entry 1). Much-improved enantiomeric ratios were recorded in tetrahydrofuran (entry 2), ethyl acetate (entry 3) and toluene (entry 4), with the latter affording a high e.r. of 95:5. This result is unsurprising, as Deng has previously demonstrated that toluene is the optimal medium for the peroxidation of unsaturated ketones and aldehydes [28,31]. However, for our substrates, these solvents were associated with moderate conversions and, consequently, yields in the range of 31%–42%. Switching to more polar solvents, such as ethanol, was no more successful (entry 5). Attempted peroxidation of 5 in ethanol instead saw the formation of a δ-ethoxy-β-keto ester as the major product in 19% yield, along with 6 in 13% yield. When this reaction was repeated in the absence of TBHP, the δ-ethoxy-β-keto ester was recovered as the sole product in 12% yield, suggesting that the solvent can add directly to the substrate, rather than displacing the δ-peroxy group. The use of water as a potential green solvent was equally disappointing (entry 6). By contrast, chlorinated solvents (entries 7–10) consistently afforded higher yields and enantiomeric ratios compared to non-chlorinated solvents (entries 1–6). 1,2-Dichloroethane afforded the best overall result with a 69% yield and a 93:7 e.r. (entry 8). Higher e.r. values were recorded with chloroform (entry 9) and carbon tetrachloride (entry 10), but at the cost of lower yields. The link between solvent choice and observed stereoselectivity is likely related to the ability of the solvent to influence the catalyst conformation [25]. Conformational investigations of the cinchona alkaloids, based on computational and spectroscopic techniques, have provided insight into how the reaction medium influences chiral induction and discrimination processes [46]. Cinchona catalysts can adopt multiple conformations and the catalytic capabilities of the cinchona scaffold are linked with its spatial arrangement [47]. Molecular mechanics calculations have demonstrated that the parent alkaloids preferentially adopt an anti-open conformation in non-polar solvents [48,49]. In polar solvents, the syn-closed and anti-closed conformations are strongly stabilised compared to the anti-open conformer [50]. The dielectric constant and the extent of hydrogen bonding can also influence the effectiveness of a catalyst in a given solvent [50].
The addition of an acid co-catalyst often has a beneficial effect on 9-amino cinchona-catalysed reactions [51]. In the specific case of asymmetric peroxidations, it has been postulated that the acid has a dual role, i.e., to firstly activate the carbonyl for condensation with the amine catalyst and to subsequently block intramolecular cyclisation and epoxide formation [51,52]. Consequently, a number of potential co-catalysts of varying acidity were examined for their impact on yields and stereoselectivity (Table 2). Although no strong relationship was apparent between the pKa of the co-catalyst and reaction outcome, less acidic additives were generally associated with both poor conversions and reduced selectivity (entries 9–15). The highest e.r. of 96:4 was recorded in the presence of either heptafluorobutyric acid (entry 4) or trifluoroacetic acid (entry 5) with yields of 63% and 58%, respectively. Pentafluorobenzoic acid (entry 6) afforded a comparable e.r. of 93:7 but a higher yield of 69%, and this was considered a better compromise between stereoselectivity and yields. Changing to more acidic co-catalysts proved detrimental. For example, reactions in the presence of either p-toluenesulfonic acid (entry 2) or methanesulfonic acid (entry 3) returned high e.r. values along with poor conversions and yields. The use of highly acidic triflic acid negatively impacted both enantiomeric ratios and yields (entry 1). Finally, incorporation of a chiral center into the co-catalyst, which often imparts improved discrimination, was not successful on this occasion (entries 13–14) [53]. Attempted optimisation of the reaction by reducing the temperature afforded lower enantiomeric ratios and yields (entry 7). Surprisingly, raising the temperature to 50 °C did not affect enantioselectivity but did give rise to a poor yield of 21% (entry 8). Subsequent efforts to improve conversions (e.g., increasing the concentration or changing the catalyst/co-catalyst loadings) were ineffectual. Extending the reaction time beyond 96 h resulted in no significant improvement in conversions. In all cases, the peroxidation reactions proceeded cleanly, with no unidentified side products apparent. It is possible that keto:enol tautomerisation of the β-keto ester starting material is partly responsible for the less than complete conversions. On the other hand, the reactions furnished the target δ-peroxy esters exclusively, with no evidence for the formation of any epoxide side products as often observed with other α,β-unsaturated carbonyl compounds [28,30,54].
In order to investigate reaction scope, a library of linear, aliphatic γ,δ-unsaturated β-keto esters was initially prepared via the niobium chloride-catalysed C-H insertion of ethyl diazoacetate into the corresponding α,β-unsaturated aldehydes [55]. Additional substrates, including aromatic and branched aliphatic substrates, were synthesised using Wittig chemistry, several of which are novel (see page 5 of Supporting Information) [56]. In each case, the E-isomer was isolated exclusively in good to excellent yields, with no trace of the Z-isomer having formed. The resulting γ,δ-unsaturated β-keto esters were found to exist as a pair of keto-enol tautomers whose ratios could be determined by the integrations of the α-protons in the 1H-NMR spectra. Each of the substrates was subjected to our optimised peroxidation conditions and subsequently reduced to the corresponding δ-hydroxy ester (Table 3). High e.r.s over 90:10 were obtained from aliphatic esters across both non-branched (entries 1–5) and branched (entries 6–8) examples. In general, yields tended to decrease as chain length increased (entries 1–5) or with increased branching (entries 6–8). Simple aromatic substrates, incorporating either a phenyl (entry 9) or naphthyl ring (entry 10), were associated with a marked reduction in enantioselectivity. By contrast, peroxidation of a homobenzylic β-keto ester (entry 11) proceeded with an e.r of 92:8. We and others have previously noted how highly conjugated, unsaturated aldehydes and ketones are resistant to cinchona-catalysed peroxidation [28,32]. In that light, the differences in outcome observed between conjugated substrates 14–15 (entries 9–10) and non-conjugated 16 are not unexpected (entry 11). However, this trend did not hold up across all conjugated γ,δ-unsaturated β-keto esters, as the incorporation of a substituent onto the aromatic ring was found to greatly influence reaction outcomes (entries 12–23). In particular, the position of the substituent was found to be significant. ortho-Substituted substrates (entries 12–15) afforded the highest enantioselectivities with e.r.s comparable to those obtained with aliphatic esters. The stereoselectivity broadly decreased from ortho- (entries 12–15), to meta- (entries 16–18) and, finally, para-substituted analogues (entries 19–21). This outcome is most likely due to conformational effects, rather than to a direct inductive effect. No comparable pattern was apparent between the different series in terms of yields. Additionally, the choice of halogen substituent was unimportant. By contrast, aromatic esters bearing strongly electron-withdrawing groups (entry 22) or strongly electron-donating groups (entry 23) did behave differently, with the latter returning poorer e.r. values. Heteroatom-containing starting materials generally proved to be more challenging substrates (entries 24–26). Peroxidation of benzylether 29 proceeded in 26% yield and 82:18 e.r. (entry 24). The reactivity of 1,3-benzodioxole 30 was comparable and a similar outcome was observed in terms of isolated yields (entry 25). Unfortunately, we were unable to separate the enantiomers by chiral HPLC to determine the enantiomeric ratio. Furyl-substituted 31 proved to be an even less suitable substrate (entry 26) while δ-substituted substrate 32 failed to react (entry 27), a finding which mirrors that observed with other similar conjugated carbonyls [28,32]. Unlike their γ,δ-unsaturated β-keto ester precursors, the peroxide products existed primarily in the ketone form, with the exception of para-methoxyphenyl-substituted 54, where a keto:enol ratio of 71:29 was noted. For most other peroxides, only trace amounts of the enol tautomer of the δ-peroxy-β-keto ester were present in the 13C-NMR spectra. All of the δ-peroxy compounds were readily reduced to their corresponding δ-hydroxy β-keto esters in good to excellent yields, apart from the naphthyl-substituted 67 (entry 10) and ortho-iodophenyl 72 (entry 15). No over-reduction of either the ester or ketone moieties was observed under the conditions employed, although some dehalogenation of ortho-iodophenyl 72 did occur, which likely accounts for the reduced yield (entry 15).
The synthetic utility of these δ-peroxy-β-keto esters lies in their potential as precursors to 1,2-dioxolanes. We have previously demonstrated how tert-butylperoxyalkyl bromides can be transformed into the corresponding cycloperoxides on treatment with silver tetrafluoroborate [32,57]. Accordingly, 6 was first reduced with sodium borohydride to afford δ-peroxy-β-hydroxy ester 83 in 83% yield (Scheme 3). Attempted conversion of 83 to the corresponding δ-peroxy-β-bromo ester using the methodology of Khazdooz et al. did not proceed as planned [58]. Instead, an unexpected product was formed in <5% yield after 24 h which was subsequently identified as 1,2-dioxolane 86. Repeating the reaction in the absence of potassium bromide saw the direct conversion of 83 to 86, although starting material was still present by TLC analysis after 4 days. Increasing the loading of phosphorus pentoxide from two to six equivalents effected full conversion after 90 min, and 86 was recovered in 61% yield. The reaction time could be further shortened to 15 min on addition of 10 equivalents of phosphorus pentoxide, albeit with a lower yield of 43%.
Evidence for the formation of 1,2-dioxolane 86 was provided by mass spectrometry and NMR spectroscopy. An ion of mass 211.0934 was detected by HRMS, which corresponds to the sodium adduct of 86. Additionally, the chemical shifts of the C-3 peaks at 81.9 ppm and 82.5 ppm and the C-4 peaks at 44.9 ppm and 45.4 ppm are similar to those reported for comparable dioxolanes in the literature [59,60,61,62]. The cis- and trans-isomers may be differentiated by the distinct chemical shifts of the H-4 protons which appear at 1.89 ppm and 2.88 ppm in cis-86 due to their differing chemical environments (Figure 2). Additionally, geminal coupling of H-4, and further coupling to H-3 and H-5, gives rise to a characteristic ddd splitting pattern. By contrast, in the case of trans-86, the H-4 protons have a similar chemical shift of 2.41 ppm and result in a simpler dd splitting pattern. Based on the integrations of H-4 in the 1H-NMR spectrum of 86, the cis- and trans-1,2-dioxolanes were generated in a ratio of 44:56. δ-Peroxy-β-hydroxy esters 84 and 85 were also subjected to these conditions, and formation of the cycloperoxide products was again found to be successful. Cyclisation of phenyl-substituted 84 proceeded in a reduced yield of 33%, whereas the presence of the –CH2CH2– linker in 85 resulted in a slightly improved yield of 46%.
Given the Lewis acidic nature of phosphorus pentoxide, it is likely that the reaction proceeds in a manner similar to that of our previously established silver-mediated cyclisation of tert-butylperoxyalkyl bromides, i.e., by trapping of the oxygen lone pair by the incipient carbocation, resulting in loss of the tert-butyl group and formation of the target cycloperoxide (Figure 3) [18,40]. In a similar vein, Khazaei and colleagues have previously demonstrated the use of phosphorus pentoxide for the etherification of benzylic alcohols [63].

3. Materials and Methods

Synthesis of (4-ethoxy-2,4-dioxobutyl)triphenylphosphonium chloride: Ethyl 4-chloroacetoacetate (6.30 g, 38.2 mmol, 1.0 eq.) and triphenylphosphine (10.00 g, 38.2 mmol, 1.0 eq.) were dissolved in toluene and stirred at 50 °C for 24 h. Solvent was removed under reduced pressure to afford 15.49 g, 36.3 mmol (4-ethoxy-2,4-dioxobutyl)triphenylphosphonium chloride as a pale yellow solid in a 95% yield without the need for purification [64].
Wittig olefination: (4-Ethoxy-2,4-dioxobutyl)triphenylphosphonium chloride (2.56 g, 6.0 mmol, 1.2 eq.) was dissolved in anhydrous THF (20 mL) at 0 °C and sodium hydride (360 mg, 15.0 mmol, 3.0 eq.) was added and stirred for 30 min. The aldehyde (5.0 mmol, 1.0 eq.) was added dropwise, and the reaction was allowed to come to room temperature and stirred for 5 h. The reaction was quenched with dropwise addition of water (10 mL) and extracted with ethyl acetate (30 mL) and saturated ammonium chloride solution (30 mL). The organic layer was dried with magnesium sulfate, concentrated in vacuo and the residue was subjected to silica gel column chromatography with diethyl ether:hexane in gradient ratios (2–20%) to afford the target γ,δ-unsaturated β-ketoester [56].
Synthesis of (8S,9S)-9-amino-(9-deoxy)epiquinine (4): Quinine (2.00 g, 6.2 mmol, 1.0 eq.) and triphenylphosphine (1.93 g, 7.4 mmol, 1.2 eq.) were dissolved in 50 mL of THF at 0 °C under nitrogen. Diisopropyl azodicarboxylate (DIAD) (1.45 mL, 7.4 mmol, 1.2 eq.) was added dropwise via syringe giving a clear solution. After 5 min, diphenylphosphoryl azide (DPPA) (1.74 mL, 8.0 mmol, 1.2 eq.) was added dropwise via syringe, turning the reaction mixture yellow. The reaction was stirred overnight at room temperature under nitrogen. The temperature was increased to 50 °C for 2 h. Additional triphenylphosphine (2.10 g, 8.0 mmol, 1.4 eq.) was added and the reaction temperature was maintained at 50 °C until gas formation ceased (approximately 2 h). The reaction was allowed to cool to room temperature. Deionised water (2.00 mL) was added, and the reaction mixture was stirred overnight. The reaction mixture was concentrated in vacuo, and the residue was partitioned between dichloromethane (100 mL) and 2N HCl (100 mL). Following vigorous shaking, the aqueous phase was separated and washed with dichloromethane (3 × 20 mL). The aqueous phase was concentrated under reduced pressure. The crude product was recrystallized by dissolving in the minimum amount of methanol and adding ethyl acetate as an anti-solvent until the opalescence became persistent. The salt was basified with excess sodium carbonate and extracted with dichloromethane (3 × 20 mL). The organic layers were combined, dried with magnesium sulfate and reduced in vacuo to give 1.22 g of a sticky brown oil in a 62% yield [65,66].
Enantioselective peroxidation: γ,δ-Unsaturated β-ketoester (3.5 mmol, 1 eq.), pentafluorobenzoic acid (149 mg, 0.7 mmol, 0.2 eq.), quinine-derived catalyst 4 (113 mg, 0.35 mmol, 0.1 eq.) and tert-butyl hydroperoxide (1.15 mL of a 5.5 M solution in decane, 6.3 mmol, 1.8 eq.) were added to 1,2-dichloroethane (5 mL) and the mixture was stirred at room temperature for 96 h. The reaction mixture was then filtered through a silica gel pad and washed with diethyl ether (40 mL), and the solvent was removed in vacuo. The resulting residue was subjected to silica gel column chromatography with diethyl ether:hexane as the eluent in gradient ratios (1–10%) to afford the target peroxide.
Selective peroxide reduction: The appropriate δ-peroxy β-keto ester (0.2 mmol, 1 eq.) and zinc powder (171 mg, 2.8 mmol, 14 eq.) were added to a 50% aqueous solution of acetic acid (5 mL) and stirred at room temperature until the peroxide was fully consumed, as confirmed by TLC (7 h). The reaction was quenched with saturated sodium bicarbonate solution (15 mL). Ethyl acetate (20 mL) was added, and the organic layer was separated. The aqueous layer was extracted with ethyl acetate (2 × 15 mL). The combined organic layers were washed with brine (30 mL) and dried with magnesium sulfate before the solvent was removed in vacuo. The resulting residue was subjected to silica gel column chromatography with diethyl ether:hexane as eluent in gradient ratios (20–50%) to afford the target δ-hydroxy β-keto ester.
HPLC analysis: The enantiopurity of chiral δ-hydroxy β-keto esters was determined by chiral stationary phase high-performance liquid chromatography (HPLC). The stereoisomers were separated using a Phenomenex AMY-2 column and an eluent system consisting of 5–15% isopropyl alcohol in hexane; the exact conditions are outlined in the Supporting Information. HPLC analysis was performed on a Waters alliance 2695 separations module equipped with a Waters 2996 photodiode array detector. All chiral stationary phase HPLC analysis was carried out at ambient temperature unless otherwise stated.
Selective ketone reduction: δ-Peroxy β-keto ester (0.4 mmol, 1 eq.) and sodium borohydride (15 mg, 0.4 mmol, 1 eq.) were added to ethanol (5 mL) and stirred at room temperature until the starting material was fully consumed, as confirmed by TLC (approximately 2 h). The reaction mixture was passed through a pad of silica gel and washed with diethyl ether (15 mL), and the solvent was removed in vacuo. The resulting residue was subjected to silica gel column chromatography with diethyl ether: hexane as eluent in gradient ratios (5–25%).
P2O5-mediated cyclisation: δ-Peroxy β-hydroxy ester (0.4 mmol, 1 eq.) was dissolved in acetonitrile (5 mL), and phosphorus pentoxide (342 mg, 2.4 mmol, 6 eq.) was added. The reaction was stirred at room temperature until the δ-peroxy β-hydroxy ester was consumed, as confirmed by TLC (90 min). The reaction mixture was then filtered through a silica gel pad and washed with diethyl ether (20 mL), and the solvent was removed in vacuo. The resulting residue was subjected to silica gel column chromatography with diethyl ether: hexane as the eluent in gradient ratios (5–25%).

4. Conclusions

To date, literature methods for the synthesis of chiral peroxides remain relatively rare. In this article, we have demonstrated the use of a quinine-derived organocatalyst for the asymmetric peroxidation of a range of γ,δ-unsaturated β-keto esters. The resulting δ-peroxy-β-keto esters can be obtained in good to excellent enantiomeric ratios and in moderate to good yields. Both aliphatic and aromatic substrates are amenable to enantioselective peroxidation, with higher selectivity observed with substituted aromatic rings. This peroxidation methodology may also be applicable to other substrates, such as unsaturated α-keto esters [67]. Furthermore, the use of alternative cinchona-based catalysts may help in improving conversions [68]. The subsequent selective reduction of the δ-peroxy-β-keto ester products to their corresponding chiral δ-hydroxy-β-keto esters proceeds in high yields, without affecting the important β-keto ester functionality. We have also discovered a novel phosphorus pentoxide-mediated cyclisation of δ-peroxy-β-hydroxy esters to chiral 1,2-dioxolanes. Alkyl-, aryl- and homobenzyl-substituted δ-peroxy-β-hydroxy esters were successfully cyclised under optimised conditions. We surmise that reaction proceeds by trapping of the peroxide oxygen lone pair, resulting in loss of the tert-butyl group and formation of the target cycloperoxide. This chemistry provides a concise route to chiral 1,2-dioxolane-3-acetic acids which are incorporated into a wide range of bioactive natural products.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28114317/s1, Full experimental details, including compound characterisation and associated spectra, are available in the Supporting Information.

Author Contributions

Conceptualization, T.P.O.; methodology, M.C.H. and H.G.; investigation, M.C.H. and H.G.; writing—M.C.H.; writing—review and editing, T.P.O. and M.C.H.; supervision, T.P.O.; funding acquisition, T.P.O. All authors have read and agreed to the published version of the manuscript.

Funding

The Irish Research Council is gratefully acknowledged for funding PhD scholarships for M.C.H (GOIPG/2018/496) and H.G. (GOIPG/2013/113). This work was undertaken using equipment provided by Science Foundation Ireland though a research infrastructure award for process flow spectroscopy (ProSpect) (SFI 15/RI/3221).

Institutional Review Board Statement

Not acceptable.

Informed Consent Statement

Not acceptable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Casteel, D.A. Peroxy natural products. Nat. Prod. Rep. 1992, 9, 289–312. [Google Scholar] [CrossRef] [PubMed]
  2. Casteel, D.A. Peroxy natural products. Nat. Prod. Rep. 1999, 16, 55–73. [Google Scholar] [CrossRef]
  3. Liu, D.-Z.; Liu, J.-K. Peroxy natural products. Nat. Prod. Bioprospecting 2013, 3, 161–206. [Google Scholar] [CrossRef]
  4. Norris, M.D.; Perkins, M.V. Structural diversity and chemical synthesis of peroxide and peroxide-derived polyketide metabolites from marine sponges. Nat. Prod. Rep. 2016, 33, 861–880. [Google Scholar] [CrossRef]
  5. Kossuga, M.H.; Nascimento, A.M.; Reimão, J.Q.; Tempone, A.G.; Taniwaki, N.N.; Veloso, K.; Ferreira, A.G.; Cavalcanti, B.C.; Pessoa, C.; Moraes, M.O.; et al. Antiparasitic, Antineuroinflammatory, and Cytotoxic Polyketides from the Marine Sponge Plakortis angulospiculatus Collected in Brazil. J. Nat. Prod. 2008, 71, 334–339. [Google Scholar] [CrossRef]
  6. Dembitsky, V.M. Bioactive peroxides as potential therapeutic agents. Eur. J. Med. Chem. 2008, 43, 223–251. [Google Scholar] [CrossRef]
  7. Dembitsky, V.M. Astonishing Diversity of Natural Peroxides as Potential Therapeutic Agents. J. Mol. Genet. Med. 2015, 9, 1000163. [Google Scholar]
  8. Su, X.-Z.; Miller, L.H. The discovery of artemisinin and the Nobel Prize in Physiology or Medicine. Sci. China Life Sci. 2015, 58, 1175–1179. [Google Scholar] [CrossRef]
  9. WHO. WHO Guidelines for Malaria; World Health Organisation: Geneva, Switzerland, 2022; Available online: https://www.who.int/publications/i/item/guidelines-for-malaria (accessed on 10 April 2023).
  10. Chen, Y.; Killday, K.B.; McCarthy, P.J.; Schimoler, R.; Chilson, K.; Selitrennikoff, C.; Pomponi, S.A.; Wright, A.E. Three New Peroxides from the Sponge Plakinastrella Species. J. Nat. Prod. 2001, 64, 262–264. [Google Scholar] [CrossRef]
  11. Jiménez-Romero, C.; Ortiz, I.; Vicente, J.; Vera, B.; Rodríguez, A.D.; Nam, S.; Jove, R. Bioactive Cycloperoxides Isolated from the Puerto Rican Sponge Plakortis halichondrioides. J. Nat. Prod. 2010, 73, 1694–1700. [Google Scholar] [CrossRef]
  12. Terent’ev, A.O.; Borisov, D.A.; Vil’, V.A.; Dembitsky, V.M. Synthesis of five- and six-membered cyclic organic peroxides: Key transformations into peroxide ring-retaining products. Beilstein J. Org. Chem. 2014, 10, 34–114. [Google Scholar] [CrossRef] [PubMed]
  13. Gandhi, H.; O’Reilly, K.; Gupta, M.K.; Horgan, C.; O’Leary, E.M.; O’Sullivan, T.P. Advances in the synthesis of acyclic peroxides. RSC Adv. 2017, 7, 19506–19556. [Google Scholar] [CrossRef]
  14. Lassaletta, J.M. Spotting trends in organocatalysis for the next decade. Nat. Commun. 2020, 11, 3787. [Google Scholar] [CrossRef] [PubMed]
  15. Alemán, J.; Cabrera, S. Applications of asymmetric organocatalysis in medicinal chemistry. Chem. Soc. Rev. 2013, 42, 774–793. [Google Scholar] [CrossRef]
  16. Hughes, D.L. Asymmetric Organocatalysis in Drug Development—Highlights of Recent Patent Literature. Org. Process Res. Dev. 2018, 22, 574–584. [Google Scholar] [CrossRef]
  17. Grondal, C.; Jeanty, M.; Enders, D. Organocatalytic cascade reactions as a new tool in total synthesis. Nat. Chem. 2010, 2, 167–178. [Google Scholar] [CrossRef]
  18. Marqués-López, E.; Herrera, R.P.; Christmann, M. Asymmetric organocatalysis in total synthesis—A trial by fire. Nat. Prod. Rep. 2010, 27, 1138–1167. [Google Scholar] [CrossRef]
  19. Abbasov, M.E.; Romo, D. The ever-expanding role of asymmetric covalent organocatalysis in scalable, natural product synthesis. Nat. Prod. Rep. 2014, 31, 1318–1327. [Google Scholar] [CrossRef]
  20. Xiang, S.-H.; Tan, B. Advances in asymmetric organocatalysis over the last 10 years. Nat. Commun. 2020, 11, 3786. [Google Scholar] [CrossRef]
  21. García Mancheño, O.; Waser, M. Recent Developments and Trends in Asymmetric Organocatalysis. Eur. J. Org. Chem. 2023, 26, e202200950. [Google Scholar] [CrossRef]
  22. Han, B.; He, X.-H.; Liu, Y.-Q.; He, G.; Peng, C.; Li, J.-L. Asymmetric organocatalysis: An enabling technology for medicinal chemistry. Chem. Soc. Rev. 2021, 50, 1522–1586. [Google Scholar] [CrossRef] [PubMed]
  23. List, B. Introduction:  Organocatalysis. Chem. Rev. 2007, 107, 5413–5415. [Google Scholar] [CrossRef]
  24. MacMillan, D.W.C. The advent and development of organocatalysis. Nature 2008, 455, 304–308. [Google Scholar] [CrossRef] [PubMed]
  25. Marcelli, T. Organocatalysis: Cinchona catalysts. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 142–152. [Google Scholar] [CrossRef]
  26. Jiang, L.; Chen, Y.-C. Recent advances in asymmetric catalysis with cinchona alkaloid-based primary amines. Catal. Sci. Technol. 2011, 1, 354–365. [Google Scholar] [CrossRef]
  27. Marcelli, T.; Hiemstra, H. Cinchona Alkaloids in Asymmetric Organocatalysis. Synthesis 2010, 2010, 1229–1279. [Google Scholar] [CrossRef]
  28. Lu, X.; Liu, Y.; Sun, B.; Cindric, B.; Deng, L. Catalytic Enantioselective Peroxidation of α,β-Unsaturated Ketones. J. Am. Chem. Soc. 2008, 130, 8134–8135. [Google Scholar] [CrossRef]
  29. Reisinger, C.M.; Wang, X.; List, B. Catalytic Asymmetric Hydroperoxidation of α,β-Unsaturated Ketones: An Approach to Enantiopure Peroxyhemiketals, Epoxides, and Aldols. Angew. Chem. Int. Ed. 2008, 47, 8112–8115. [Google Scholar] [CrossRef]
  30. Lifchits, O.; Mahlau, M.; Reisinger, C.M.; Lee, A.; Farès, C.; Polyak, I.; Gopakumar, G.; Thiel, W.; List, B. The Cinchona Primary Amine-Catalyzed Asymmetric Epoxidation and Hydroperoxidation of α,β-Unsaturated Carbonyl Compounds with Hydrogen Peroxide. J. Am. Chem. Soc. 2013, 135, 6677–6693. [Google Scholar] [CrossRef]
  31. Hu, L.; Lu, X.; Deng, L. Catalytic Enantioselective Peroxidation of α,β-Unsaturated Aldehydes for the Asymmetric Synthesis of Biologically Important Chiral Endoperoxides. J. Am. Chem. Soc. 2015, 137, 8400–8403. [Google Scholar] [CrossRef]
  32. O’Reilly, K.; Gupta, M.K.; Gandhi, H.K.; Kumar, V.P.; Eccles, K.S.; Lawrence, S.E.; O’Sullivan, T.P. Cinchona-catalysed, Enantioselective Synthesis of β-Peroxycarboxylic Acids, β-Peroxyesters and β-Peroxyalcohols. Curr. Org. Chem. 2016, 20, 2633–2638. [Google Scholar] [CrossRef]
  33. O’Reilly, K.; Gupta, M.K.; Gandhi, H.; Kumar, V.P.; O’Sullivan, T.P. Asymmetric Peroxidation of α,β-Unsaturated Aldehydes under Diarylprolinol Ether Catalysis. Curr. Org. Chem. 2017, 21, 2013–2016. [Google Scholar] [CrossRef]
  34. Russo, A.; Lattanzi, A. Catalytic Asymmetric β-Peroxidation of Nitroalkenes. Adv. Synth. Catal. 2008, 350, 1991–1995. [Google Scholar] [CrossRef]
  35. Lu, X.; Deng, L. Catalytic Asymmetric Peroxidation of α,β-Unsaturated Nitroalkenes by a Bifunctional Organic Catalyst. Org. Lett. 2014, 16, 2358–2361. [Google Scholar] [CrossRef]
  36. Benetti, S.; Romagnoli, R.; De Risi, C.; Spalluto, G.; Zanirato, V. Mastering beta-Keto Esters. Chem. Rev. 1995, 95, 1065–1114. [Google Scholar] [CrossRef]
  37. Everaere, K.; Mortreux, A.; Carpentier, J.-F. Ruthenium(II)-Catalyzed Asymmetric Transfer Hydrogenation of Carbonyl Compounds with 2-Propanol and Ephedrine-Type Ligands. Adv. Synth. Catal. 2003, 345, 67–77. [Google Scholar] [CrossRef]
  38. Ratovelomanana-Vidal, V.; Girard, C.; Touati, R.; Tranchier, J.P.; Hassine, B.B.; Genêt, J.P. Enantioselective Hydrogenation of β-Keto Esters using Chiral Diphosphine-Ruthenium Complexes: Optimization for Academic and Industrial Purposes and Synthetic Applications. Adv. Synth. Catal. 2003, 345, 261–274. [Google Scholar] [CrossRef]
  39. Bariotaki, A.; Kalaitzakis, D.; Smonou, I. Enzymatic Reductions for the Regio- and Stereoselective Synthesis of Hydroxy-keto Esters and Dihydroxy Esters. Org. Lett. 2012, 14, 1792–1795. [Google Scholar] [CrossRef]
  40. Heravi, M.M.; Khaghaninejad, S.; Mostofi, M. Chapter one—Pechmann reaction in the synthesis of coumarin derivatives. In Advances in Heterocyclic Chemistry; Academic Press: Cambridge, MA, USA, 2014; Volume 112, pp. 1–50. [Google Scholar] [CrossRef]
  41. Hennessy, M.C.; O’Sullivan, T.P. Recent advances in the transesterification of β-keto esters. RSC Adv. 2021, 11, 22859–22920. [Google Scholar] [CrossRef]
  42. Khademi, Z.; Heravi, M.M. Applications of Claisen condensations in total synthesis of natural products. An old reaction, a new perspective. Tetrahedron 2022, 103, 132573. [Google Scholar] [CrossRef]
  43. Govender, T.; Arvidsson, P.I.; Maguire, G.E.M.; Kruger, H.G.; Naicker, T. Enantioselective Organocatalyzed Transformations of β-Ketoesters. Chem. Rev. 2016, 116, 9375–9437. [Google Scholar] [CrossRef] [PubMed]
  44. Piovesana, S.; Schietroma, D.M.S.; Tulli, L.G.; Monaco, M.R.; Bella, M. Unsaturated β-ketoesters as versatile electrophiles in organocatalysis. Chem. Commun. 2010, 46, 5160–5162. [Google Scholar] [CrossRef] [PubMed]
  45. Fujiwara, T.; Tsutsumi, T.; Nakata, K.; Nakatsuji, H.; Tanabe, Y. Asymmetric Total Syntheses of Two 3-Acyl-5,6-dihydro-2H-pyrones: (R)-Podoblastin-S and (R)-Lachnelluloic Acid with Verification of the Absolute Configuration of (−)-Lachnelluloic Acid. Molecules 2017, 22, 69. [Google Scholar] [CrossRef] [PubMed]
  46. Song, C.E. An overview of cinchona alkaloids in chemistry. In Cinchona Alkaloids in Synthesis and Catalysis; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009; pp. 1–10. [Google Scholar] [CrossRef]
  47. Boratyński, P.J.; Zielińska-Błajet, M.; Skarżewski, J. Chapter Two—Cinchona Alkaloids—Derivatives and Applications. Alkaloids Chem. Bio. 2019, 82, 29–145. [Google Scholar] [CrossRef]
  48. Dijkstra, G.D.H.; Kellogg, R.M.; Wynberg, H. Conformational analysis of some chiral catalysts of the cinchona and ephedra family. The alkaloid catalyzed addition of aromatic thiols to cyclic α,β-unsaturated ketones. Red. Trav. Chim. Pays-Bas 1989, 108, 195–204. [Google Scholar] [CrossRef]
  49. Dijkstra, G.D.H.; Kellogg, R.M.; Wynberg, H.; Svendsen, J.S.; Marko, I.; Sharpless, K.B. Conformational study of cinchona alkaloids. A combined NMR, molecular mechanics and X-ray approach. J. Am. Chem. Soc. 1989, 111, 8069–8076. [Google Scholar] [CrossRef]
  50. Bürgi, T.; Baiker, A. Conformational Behavior of Cinchonidine in Different Solvents:  A Combined NMR and ab Initio Investigation. J. Am. Chem. Soc. 1998, 120, 12920–12926. [Google Scholar] [CrossRef]
  51. Melchiorre, P. Cinchona-based Primary Amine Catalysis in the Asymmetric Functionalization of Carbonyl Compounds. Angew. Chem. Int. Ed. 2012, 51, 9748–9770. [Google Scholar] [CrossRef]
  52. Evans, G.J.S.; White, K.; Platts, J.A.; Tomkinson, N.C.O. Computational study of iminium ion formation: Effects of amine structure. Org. Biomol. Chem. 2006, 4, 2616–2627. [Google Scholar] [CrossRef]
  53. Tian, X.; Cassani, C.; Liu, Y.; Moran, A.; Urakawa, A.; Galzerano, P.; Arceo, E.; Melchiorre, P. Diastereodivergent Asymmetric Sulfa-Michael Additions of α-Branched Enones using a Single Chiral Organic Catalyst. J. Am. Chem. Soc. 2011, 133, 17934–17941. [Google Scholar] [CrossRef]
  54. Marigo, M.; Franzén, J.; Poulsen, T.B.; Zhuang, W.; Jørgensen, K.A. Asymmetric Organocatalytic Epoxidation of α,β-Unsaturated Aldehydes with Hydrogen Peroxide. J. Am. Chem. Soc. 2005, 127, 6964–6965. [Google Scholar] [CrossRef] [PubMed]
  55. Gandhi, H.; O’Sullivan, T.P. Preparation of γ,δ-unsaturated-β-ketoesters: Lewis acid-catalysed CH insertion of ethyl diazoacetate into α,β-unsaturated aldehydes. Tetrahedron Lett. 2017, 58, 3533–3535. [Google Scholar] [CrossRef]
  56. Pietrusiewicz, K.M.; Monkiewicz, J. Anionic activation of stabilized ylides. A highly Z-stereoselective wittig reaction of (3-ethoxycarbonyl-2-oxopropylidene)triphenyl-phosphorane with aliphatic aldehydes. Tetrahedron Lett. 1986, 27, 739–742. [Google Scholar] [CrossRef]
  57. Porter, N.A.; Mitchell, J.C. Intramolecular alkylation of peroxides and hydroperoxides; peroxide transfer via peroxonium intermediates. Tetrahedron Lett. 1983, 24, 543–546. [Google Scholar] [CrossRef]
  58. Khazdooz, L.; Zarei, A.; Aghaei, H.; Azizi, G.; Gheisari, M.M. An efficient and selective method for the iodination and bromination of alcohols under mild conditions. Tetrahedron Lett. 2016, 57, 168–171. [Google Scholar] [CrossRef]
  59. Nguyen, T.L.; Ferrié, L.; Figadère, B. Synthesis of 3,5-disubstituted-1,2-dioxolanes: Access to analogues of mycangimycin and some rearrangement products. Tetrahedron Lett. 2016, 57, 5286–5289. [Google Scholar] [CrossRef]
  60. Pinet, A.; Nguyen, L.T.; Figadère, B.; Ferrié, L. Synthesis of 3,5-Disubstituted 1,2-Dioxolanes. Eur. J. Org. Chem. 2020, 2020, 7407–7416. [Google Scholar] [CrossRef]
  61. Pinet, A.; Figadère, B.; Ferrié, L. Access to Functionalized 3,5-Disubstituted 1,2-Dioxolanes under Mild Conditions through Indium(III) Chloride/Trimethylsilyl Chloride or Scandium(III) Triflate Catalysis. Adv. Synth. Catal. 2020, 362, 1190–1194. [Google Scholar] [CrossRef]
  62. Pinet, A.; Cojean, S.; Nguyen, L.T.; Vásquez-Ocmín, P.; Maciuk, A.; Loiseau, P.M.; Le Pape, P.; Figadère, B.; Ferrié, L. Anti-protozoal and anti-fungal evaluation of 3,5-disubstituted 1,2-dioxolanes. Bioorganic Med. Chem. Lett. 2021, 47, 128196. [Google Scholar] [CrossRef]
  63. Khazaei, A.; Rad, M.N.S.; Borazjani, M.K.; Saednia, S.; Borazjani, M.K.; Golbaghi, M.; Behrouz, S. Highly Efficient Etherification and Oxidation of Aromatic Alcohols Using Supported and Unsupported Phosphorus Pentoxide as a Heterogeneous Reagent. Synth. Commun. 2011, 41, 1544–1553. [Google Scholar] [CrossRef]
  64. Xu, W.; Wang, X.-B.; Wang, Z.-M.; Wu, J.-J.; Li, F.; Wang, J.; Kong, L.-Y. Synthesis and evaluation of donepezil–ferulic acid hybrids as multi-target-directed ligands against Alzheimer’s disease. MedChemComm 2016, 7, 990–998. [Google Scholar] [CrossRef]
  65. Vakulya, B.; Varga, S.; Csámpai, A.; Soós, T. Highly Enantioselective Conjugate Addition of Nitromethane to Chalcones Using Bifunctional Cinchona Organocatalysts. Org. Lett. 2005, 7, 1967–1969. [Google Scholar] [CrossRef] [PubMed]
  66. Cassani, C.; Martín-Rapún, R.; Arceo, E.; Bravo, F.; Melchiorre, P. Synthesis of 9-amino(9-deoxy)epi cinchona alkaloids, general chiral organocatalysts for the stereoselective functionalization of carbonyl compounds. Nat. Protoc. 2013, 8, 325–344. [Google Scholar] [CrossRef] [PubMed]
  67. Deng, R.; Han, T.-J.; Gao, X.; Yang, Y.-F.; Mei, G.-J. Further developments of β,γ-unsaturated α-ketoesters as versatile synthons in asymmetric catalysis. iScience 2022, 25, 103913. [Google Scholar] [CrossRef] [PubMed]
  68. Nagy, S.; Fehér, Z.; Dargó, G.; Barabás, J.; Garádi, Z.; Mátravölgyi, B.; Kisszékelyi, P.; Dargó, G.; Huszthy, P.; Höltzl, T.; et al. Comparison of Cinchona Catalysts Containing Ethyl or Vinyl or Ethynyl Group at Their Quinuclidine Ring. Materials 2019, 12, 3034. [Google Scholar] [CrossRef]
Figure 1. Bioactive peroxide-containing natural products.
Figure 1. Bioactive peroxide-containing natural products.
Molecules 28 04317 g001
Scheme 1. Previously reported organocatalytic peroxidations of unsaturated carbonyl compounds [28,29,30,31,32].
Scheme 1. Previously reported organocatalytic peroxidations of unsaturated carbonyl compounds [28,29,30,31,32].
Molecules 28 04317 sch001
Scheme 2. Current Work.
Scheme 2. Current Work.
Molecules 28 04317 sch002
Scheme 3. P2O5-mediated cyclisation of δ-peroxy-β-hydroxy esters.
Scheme 3. P2O5-mediated cyclisation of δ-peroxy-β-hydroxy esters.
Molecules 28 04317 sch003
Figure 2. Comparison of cis- and trans-1,2 dioxolanes.
Figure 2. Comparison of cis- and trans-1,2 dioxolanes.
Molecules 28 04317 g002
Figure 3. Proposed cyclisation mechanism.
Figure 3. Proposed cyclisation mechanism.
Molecules 28 04317 g003
Table 1. Effect of solvent on yield and enantioselectivity.
Table 1. Effect of solvent on yield and enantioselectivity.
Molecules 28 04317 i001
EntrySolventConversion [a]Isolated Yielde.r.
1MeCN40%26%69:31
2THF50%31%92:8
3EtOAc64%33%92:8
4Tol73%42%95:5
5EtOHn.d. [b]13% (19%) [c]79:21
6H2O54%45%88:12
7CH2Cl275%52%93:7
81,2-DCE87%69%93:7
9CHCl373%49%96:4
10CCl461%42%94.5:5.5
[a] Conversion determined by comparison of α-proton integrals in starting material vs. product in the 1H-NMR spectrum. [b] Not determined due to overlapping 1H-NMR signals. [c] Yield of δ-ethoxy-β-keto ester.
Table 2. Effect of co-catalyst on yield and enantioselectivity.
Table 2. Effect of co-catalyst on yield and enantioselectivity.
Molecules 28 04317 i002
EntryCo-CatalystpKaTemp.Conversion [a]Isolated Yielde.r.
1Triflic acid−14.7r.t.60%41%57:43
2p-Toluenesulfonic acid−2.8r.t.15%10%94:6
3Methanesulfonic acid−1.9r.t.23%18%95.5:4.5
4Heptafluorobutyric acid0.4r.t.84%63%96:4
5Trifluoroacetic acid0.5r.t.77%58%96:4
6Pentafluorobenzoic acid1.5r.t.87%69%93:7
7Pentafluorobenzoic acid1.54 °C76%54%89:11
8Pentafluorobenzoic acid1.550 °C32%21%94:6
9Chloroacetic acid2.9r.t.25%17%86.5:13.5
10Tartaric acid2.9r.t.4%n.d.n.d.
112,4-Bis(trifluoromethyl)benzoic acid3.3r.t.42%29%86:14
124-(Trifluoromethyl)benzoic acid3.6r.t.22%18%81:19
13Boc-L-phenylglycine3.9r.t.22%18%88:12
14Boc-D-phenylglycine3.9r.t.25%20%87:13
15Benzoic acid4.2r.t.8%n.d.n.d.
[a] Conversion determined by comparison of α-proton integrals in starting material vs. product in the 1H-NMR spectrum.
Table 3. Peroxidation of γ,δ-unsaturated β-keto esters with TBHP and subsequent reduction with zinc and acetic acid.
Table 3. Peroxidation of γ,δ-unsaturated β-keto esters with TBHP and subsequent reduction with zinc and acetic acid.
Molecules 28 04317 i003
EntryRβ-Keto EsterConversion [a]PeroxideIsolated Yielde.r.MW (g/mol)AlcoholIsolated YieldMW (g/mol)
1CH3786%3379%91:9246.305887%174.20
2CH3CH2587%669%93:7260.335971%188.22
3CH3(CH2)2874%3452%93:7274.366077%202.25
4CH3(CH2)3946%3534%95:5288.386193%216.28
5CH3(CH2)41050%3643%95:5302.416293%230.30
6(CH3)2CH11n.d.3727%90:10274.366393%202.25
7(CH3)2CHCH21233%3830%92:8288.396497%216.28
8Cy1342%3935%92:8314.426591%242.32
9C6H51451%4049%74:26308.376683%236.27
102-Naphthyl1524%4117%67:33358.436757%286.33
11C6H5CH2CH21654%4247%92:8336.436874%264.32
122-FC6H41725%4322%83:17326.366980%254.26
132-ClC6H41829%4425%91:9342.827084%270.71
142-BrC6H41932%4529%92:8387.277197%315.16
152-IC6H42029%4620%93:7434.277234%362.16
163-FC6H42137%4726%78:22326.367389%254.26
173-ClC6H42242%4817%78:22342.827486%270.71
183-BrC6H42335%4928%80:20387.277583%315.16
194-FC6H42429%5028%73:27326.367683%254.26
204-ClC6H42526%5124%76:24342.827780%270.71
214-BrC6H42634%5218%78:22387.277890%315.16
224-CF3C6H42738%5331%67:33376.377994%304.27
234-MeOC6H42843%5438%57:43338.408074%266.30
24BnOCH22927%5526%82:18352.438161%280.32
25Molecules 28 04317 i0043025%5622%n.d [b]352.388294%280.28
262-Furyl317%576%n.d [b]298.34---
27Molecules 28 04317 i005320%-0%-----
[a] Conversion determined by comparison of α-proton integrals in starting material vs. product in the 1H-NMR spectrum. [b] Not separable by chiral HPLC.
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

Hennessy, M.C.; Gandhi, H.; O’Sullivan, T.P. Organocatalytic Asymmetric Peroxidation of γ,δ-Unsaturated β-Keto Esters—A Novel Route to Chiral Cycloperoxides. Molecules 2023, 28, 4317. https://doi.org/10.3390/molecules28114317

AMA Style

Hennessy MC, Gandhi H, O’Sullivan TP. Organocatalytic Asymmetric Peroxidation of γ,δ-Unsaturated β-Keto Esters—A Novel Route to Chiral Cycloperoxides. Molecules. 2023; 28(11):4317. https://doi.org/10.3390/molecules28114317

Chicago/Turabian Style

Hennessy, Mary C., Hirenkumar Gandhi, and Timothy P. O’Sullivan. 2023. "Organocatalytic Asymmetric Peroxidation of γ,δ-Unsaturated β-Keto Esters—A Novel Route to Chiral Cycloperoxides" Molecules 28, no. 11: 4317. https://doi.org/10.3390/molecules28114317

Article Metrics

Back to TopTop