Next Article in Journal
Coincidences of the Concave Integral and the Pan-Integral
Previous Article in Journal
Focus Assessment Method of Gaze Tracking Camera Based on ε-Support Vector Regression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of (R)-Modafinil via Organocatalyzed and Non-Heme Iron-Catalyzed Sulfoxidation Using H2O2 as an Environmentally Benign Oxidant

Department of Chemistry and Pharmacy, Organic Chemistry Chair I and Interdisciplinary Center for Molecular Materials (ICMM), University of Erlangen-Nuremberg, Henkestraße 42, 91054 Erlangen, Germany
*
Author to whom correspondence should be addressed.
Symmetry 2017, 9(6), 88; https://doi.org/10.3390/sym9060088
Submission received: 11 May 2017 / Revised: 31 May 2017 / Accepted: 7 June 2017 / Published: 16 June 2017

Abstract

:
The first organocatalyzed sulfoxidation reaction towards enantioenriched (R)-modafinil (Armodafinil®), a drug against narcolepsy, is reported here. A series of chiral organocatalysts, e.g., different chiral BINOL-phosphates, or a fructose-derived N-substituted oxazolidinone ketone (Shi catalyst) were applied for the sulfoxidation reaction with environmentally friendly H2O2 as a convenient oxygen transferring agent. Furthermore, the potential of a biomimetic catalytic system consisting of FeCl3 and a dipeptide-based chiral ligand was demonstrated, which constitutes the most successful asymmetric non-heme iron-catalyzed synthesis of (R)-modafinil so far.

1. Introduction

Sulfoxides demonstrate versatile properties and are ubiquitous in bioactive compounds and drugs [1]. Being of medicinal interest, chiral sulfoxides serve as building blocks for the generation of biologically active compounds [2]. With the discovery of their presence in naturally chiral sulfoxide metabolites [3], such as cysteine sulfoxide derivatives, the demand for new synthetic routes was steadily growing, especially since the pioneering work of Kagan and Modena [4,5,6].
Optically active sulfoxides are important representatives of therapeutically used drugs. There are several examples that playing an important role in medicinal and pharmaceutical chemistry [2], including Esomeprazole [7,8], which acts as proton-pump inhibitor; the selective NK2 antagonist ZD7944 [9]; and the anti-cancer agent Sulindac® (Figure 1) [10].
Modafinil (also known as Provigil® [11], Figure 2), another important sulfoxide, is administered as a drug against narcolepsy [12] with significant advantages compared to dexamphetamine [13] and several other stimulants [14] because of a lower abuse potential [15,16].
It is furthermore used for the treatment of depression [17], attention-deficit hyperactivity disorder [18], Parkinson’s disease [19], and epilepsy [20]. Moreover, it is a memory-improving and mood-brightening psychostimulant [21].
After the first synthesis was developed in 1979 by Lafon [22], several methods have been envisioned for its racemic generation using hydrogen peroxide as an abundant, environmentally benign oxidant [23,24,25,26,27,28,29].
The fact that there is a continued demand for novel synthetic methods to attain modafinil is further demonstrated by a recently published one-pot parallel synthetic approach [29] as well as another strategy, established by Cibulka and co-workers, employing electron-deficient heteroarenium salts for the activation of hydrogen peroxide [30].
The quantitative potency of the two enantiomers of modafinil, however, is indeed influenced by the stereogenic sulfoxide group. (R)-modafinil has a longer half-life in the human body compared to (S)-modafinil [31], which is metabolized three times faster [32].
An ester, amide, or carboxylic acid moiety close to the sulfide function, however, can adversely affect enantioselectivity [33]. Given that no directing group is available adjacent to the sulfur atom, compared, for instance, to omeprazole [7], most known procedures aim to circumvent an enantioselective oxidation. As an alternative, racemic resolution of diastereomers [34,35], preferential crystallization or chiral chromatography can be applied [36]. Thus, the variety of enantioselective synthetic approaches is limited to a small selection, namely, a microbial sulfoxidation [37], an advanced Kagan method patented by Cephalon [33], and an oxygen-transfer by a chiral oxaziridine in ionic liquid [38].
These methods, however, suffer from the need for hazardous oxidizing agents, expensive solvents, and metal catalysts. Apart from a vanadium-based catalytic system, developed in our group, which was able to provide (R)-modafinil in excellent yield and moderate enantiomeric excess within a very short reaction time [39], all other attempts making use of convenient metal complexes generally resulted in a drastically decreased performance of the catalyst [38].
For the pharmaceutical industry it is a permanent issue to free biologically active compounds from metal traces. Thus, the development of a metal-free sulfoxidation process is in high demand.
A class of powerful organocatalysts is represented by Brønsted acid catalysts such as BINOL-derived phosphoric acids, which have already been applied in numerous highly enantioselective transformations [40,41,42,43,44,45,46,47,48,49] since the pioneering studies by the groups of Akiyama et al. [50] and Uraguchi and Terada [51].
Wang, Tao and co-workers [52] were the first to develop a sulfoxidation reaction catalyzed by BINOL-phosphates, and List et al. [53,54] impressively demonstrated the potential of a new family of chiral phosphoric acids with results analogous to the best obtained with metal catalysts. Having already shown a very good performance in a patented, efficient, and direct route towards Sulindac® (Figure 1) [53,54], BINOL-derived phosphoric acids have surprisingly never been applied for the enantioselective synthesis of modafinil.
Following our previous successes in applying Lewis base/Brønsted acid catalysts for various organic transformations [55,56,57], we decided to test BINOL-phosphates (chiral bifunctional organocatalysts, Scheme 1) and additional catalytic systems. To the best of our knowledge, we present herein the first organocatalyzed synthesis of enantiomerically enriched modafinil via sulfoxidation.

2. Materials and Methods

Solvents were purified by standard procedures and distilled prior to use. Reagents obtained from commercial sources were used without further purification. Thin layer chromatography (TLC) chromatography was performed on pre-coated aluminum silica gel ALUGRAM SIL G/UV254 plates (Macherey, Nagel & Co., Düren, Germany). Flash chromatography was performed using silica gel ACROS 60 Å, (particle size 0.035–0.070 mm, Thermofischer ACROS Organics, Janssen-Pharmaceuticalaan 3a, 2440 Geel, Belgium). Proton nuclear magnetic resonance spectroscopy (1H-NMR) spectra were recorded in CDCl3 with Bruker Avance 300 or 400 (Bruker, Billerica, MA, USA). Enantioselectivities were determined by chiral high pressure liquid chromatography (HPLC) analysis in comparison with authentic racemic material.

2.1. (Methylsulfinyl)Benzene (1a):

The corresponding BINOL-phosphate catalyst (0.048 mmol) was dissolved in CH2Cl2 (0.5 mL) at −10 °C. After the addition of the sulfide (0.24 mmol, 29.8 mg) aqueous 30% H2O2 (1.2 equiv, 0.29 mmol, 29.4 µL) was added in one portion. The reaction mixture was stirred at room temperature for 24 h. The product was obtained via direct purification by column chromatography (SiO2, EtOAc/PE 4:1). The enantiomeric excess of the product was determined by chiral HPLC analysis. 1H-NMR (300 MHz, CDCl3): δ = 2.70 (s, 3H), 7.47–7.54 (m, 3H), 7.60–7.64 (m, 2H).

2.2. 1-(Methylsulfinyl)-4-Nitrobenzene (1b):

The corresponding BINOL-phosphate catalyst (0.048 mmol) was dissolved in CH2Cl2 (0.5 mL) at −10 °C. The sulfide (0.24 mmol, 40.6 mg) was added, followed by the addition of aqueous 30% H2O2 (1.2 equiv, 0.29 mmol, 29.4 µL) in one portion. The reaction mixture was stirred at room temperature for 24 h. The product was directly purified by column chromatography (SiO2, EtOAc/PE 4:1). The enantiomeric excess of the product was determined by chiral HPLC analysis. 1H-NMR (300 MHz, CDCl3): δ = 2.77 (s, 3H), 7.82 (d, J = 8.9 Hz, 2H), 8.38 (d, J = 8.9 Hz, 2H).

2.3. 2-(Benzhydrylsulfinyl)Acetamide (2):

BINOL-phosphate V (0.048 mmol, 36 mg) was dissolved in CH2Cl2 (0.5 mL) at −10 °C, followed by the addition of aqueous 30% H2O2 (1.2 equiv, 0.29 mmol, 29.4 µL) in one portion. Then, the sulfide (0.24 mmol, 61.8 mg) was added to the reaction mixture, which was stirred at −10 °C. The product was directly purified by column chromatography (SiO2, EtOAc). The isolated enantiomerically enriched (R)-modafinil was identified through comparison of 1H-NMR spectra with literature data [38]. The enantiomeric excess of the product was determined by chiral HPLC analysis (Daicel Chiralpak AS, n-hexane/i-PrOH (60:40), flow 0.9 mL/min, 25 °C, 31 bar). 1H-NMR (300 MHz, DMSO-d6): δ = 3.21 (d, J = 13.5 Hz, 1H), 3.36 (d, J = 13.5 Hz, 1H), 5.34 (s, 1H), 7.32–7.43 (m, 7H), 7.50–7.52 (m, 4H), 7.68 (s, 1H).

3. Results and Discussion

Our initial work started with the screening of different BINOL-phosphates for the sulfoxidation of thioanisole as a model reaction. The required organocatalysts IVII for the reaction were prepared by conventional means [58].
Initially, the reaction was performed in dichloromethane using 20 mol % of the organocatalyst at −10 °C within 24 h. In case of BINOL-phosphate II, only moderate yield was observed (44% yield, entry 1, Table 1) and the enantiomeric excess was rather low (10% ee). A further reduction of the amount of catalyst from 20 mol % to 10 mol % resulted in a decreased yield, but the enantiomeric ratio remained constant (26% yield, 10% ee, entry 2, Table 1). Surprisingly, BINOL-phosphate III provided the corresponding sulfoxide in high yield (98%) with an enantiomeric excess of 20% (entry 3, Table 1) and we continued to investigate catalyst I. To our delight, the enantiomeric excess increased to 36% ee, but the yield was lowered to 68% (entry 4, Table 1). In the case of BINOL-phosphate with a more sterically hindered moiety, a complete loss of activity was the consequence (entry 5, Table 1). With catalyst V, on the other hand, the product could be isolated with 73% yield and an enantiomeric excess of 30% (entry 6, Table 1). Furthermore, carrying out one reaction without an external catalyst, we could preclude a background reaction (entry 7, Table 1).
We investigated the oxidation of a thioanisole derivative bearing a nitro moiety in 4-position (entries 8–11, Table 1). We applied BINOL-phosphate VI, and found that the product could be isolated with a low yield of 11% and a moderate enantiomeric excess of 33% (entry 8, Table 1). Carrying out the reaction with a lower amount of hydrogen peroxide (0.27 equivalent) under otherwise identical conditions, the isolated product had a significantly higher enantiomeric excess of 59% (entry 9, Table 1). With catalyst VII, bearing an electron withdrawing substituent on the phenyl moiety in para position, the desired product was generated only in traces (entry 10, Table 1).
Subsequently, we studied the sulfoxidation of the prochiral sulfide in the presence of 20 mol % of BINOL-phosphate V, which had previously performed best. The reaction proceeded at −10 °C, in CH2Cl2 as a solvent (Table 2). Initially, we applied the optimized reaction conditions, and the product could be isolated after 24 h in quantitative yield with an enantiomeric excess of 10% (entry 1, Table 2) In order to examine the temperature effect on the reaction, we carried out the sulfoxidation at 0 °C. The product was obtained quantitatively but with a slightly lower ee value of 7% (entry 2, Table 2). A shorter reaction time of 12 h resulted in a lower yield of 61%, and no enantioselectivity could be detected (entry 3, Table 2). A higher initial concentration of the sulfide had a positive influence on the stereoselectivity of the oxidation. Thus, following the previous trend concerning the oxidation of thioanisole (entry 9, Table 1), we observed this supporting effect on the reaction outcome for modafinil as well. Due to an accelerated reaction, the desired compound could be isolated in quantitative yield already after 12 h and the ee value slightly increased to 13% (entry 4, Table 2).
The Shi’s N-substituted oxazolidinone ketone VIII belongs to the attractive class of fructose-derived asymmetric organocatalysts, generally employed for the epoxidation of unfunctionalized terminal alkenes [59,60]. Having successfully applied it to the one-pot, two-step synthesis of β-adrenergic blockers via epoxidation as a key enantioselective step [60], we decided to evaluate its activity for the sulfoxidation reaction as well. Modafinil was generated with a yield of 38% and an enantioselectivity of 26% ee (entry 5, Table 2).
In order to minimize the number of reactants, we next intended to combine the oxidizing agent and the chiral source. Therefore, we applied the synthesized chiral hydroperoxide TADOOH IX, which is generally known for its high efficiency for sulfoxidation reactions [61]. A striking feature of this method is that no further external catalyst is needed. With an amount of 1.5 equivalent of the oxidant in tetrahydrofuran (THF), the product was generated in 66% yield with an enantioselectivity of 16% (entry 6, Table 2).
Following our previous interest in the application of the non-heme iron-catalyzed sulfoxidation reaction, which exhibited high activity in the oxidation of the thioanisol but, up to now, showed unsatisfactory results for the modafinil precursor [39], we decided to investigate a dipeptide-based chiral ligand as an alternative chiral source (Scheme 2).
Under these conditions, modafinil could be isolated in good yield of 75% and with an ee value of 24%. This represents the most successful asymmetric non-heme iron-catalyzed synthesis of (R)-modafinil so far.

4. Conclusions

We report here the first organocatalytic synthesis of enantiomerically enriched (R)-modafinil (Armodafinil®), accomplished with the chiral bifunctional BINOL-phosphate V, a fructose-derived N-substituted oxazolidinone ketone VIII, or the chiral peroxide TADOOH IX, respectively. The desired compound could be obtained via sulfoxidation using H2O2 with excellent yield up to >99% and ee values up to 23%. This protocol includes all the advantages of metal-free asymmetric organocatalysis.
We additionally showed the successful application of a dipeptide-based ligand in the biomimetic non-heme iron-catalyzed sulfoxidation towards enantiomerically enriched (R)-modafinil, which could be isolated in good yield of 75% with an enantioselectivity of 24% ee. The sulfoxidation reported here makes use of mild reaction conditions using aqueous hydrogen peroxide as an environmentally friendly oxidant.

Supplementary Materials

The following are available online at www.mdpi.com/2073-8994/9/6/88/s1: Experimental Procedures, Analytical and Spectroscopic Data of Products.

Acknowledgments

We gratefully acknowledge the financial support from Deutsche Forschungsgemeinschaft (DFG) by grant TS 87/16-3. We also thank the Interdisciplinary Center for Molecular Materials (ICMM), the Graduate School Molecular Science (GSMS) for research support, as well as the Wilhelm Sander-Stiftung (Grant No. 2014.019.1) for funding.

Author Contributions

Felix E. Held conducted the sulfoxidation reactions. Kerstin A. Stingl contributed to the catalyst screening of the model reaction. Svetlana B. Tsogoeva conceived and directed the research, supervised the synthetic experiments. The manuscript was written through contribution of all authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pulis, A.P.; Procter, D.J. C–H coupling reactions directed by sulfoxides: Teaching an old functional group new tricks. Angew. Chem. Int. Ed. 2016, 55, 9842–9860. [Google Scholar] [CrossRef] [PubMed]
  2. Legros, J.; Dehli, J.R.; Bolm, C. Applications of catalytic asymmetric sulfide oxidations to the syntheses of biologically active sulfoxides. Adv. Synth. Catal. 2005, 347, 19–31. [Google Scholar] [CrossRef]
  3. Bauder, C.; Martínez, J.; Salom-Roig, X. Chiral sulfoxides as building blocks for enantiopure 1,3-diol precursors in the synthesis of natural products. Curr. Org. Synth. 2014, 10, 885–902. [Google Scholar] [CrossRef]
  4. Pitchen, P.; Dunach, E.; Deshmukh, M.N.; Kagan, H.B. An efficient asymmetric oxidation of sulfides to sulfoxides. J. Am. Chem. Soc. 1984, 106, 8188–8193. [Google Scholar] [CrossRef]
  5. Di Furia, F.; Modena, G.; Seraglia, R. Synthesis of chiral sulfoxides by metal-catalyzed oxidation witht-butyl hydroperoxide. Synthesis 1984, 1984, 325–326. [Google Scholar] [CrossRef]
  6. Pitchen, P.; Kagan, H.B. An efficient asymmetric oxidation of sulfides to sulfoxides. Tetrahedron Lett. 1984, 25, 1049–1052. [Google Scholar] [CrossRef]
  7. Cotton, H.; Elebring, T.; Larsson, M.; Li, L.; Sörensen, H.; von Unge, S. Asymmetric synthesis of esomeprazole. Tetrahedron 2000, 11, 3819–3825. [Google Scholar] [CrossRef]
  8. Caron, S.; Dugger, R.W.; Ruggeri, S.G.; Ragan, J.A.; Ripin, D.H. Large-scale oxidations in the pharmaceutical industry. Chem. Rev. 2006, 106, 2943–2989. [Google Scholar] [CrossRef] [PubMed]
  9. Jacobs, R.T.; Shenvi, A.B.; Mauger, R.C.; Ulatowski, T.G.; Aharony, D.; Buckner, C.K. 4-alkylpiperidines related to SR-48968: Potent antagonists of the neurokinin-2 (NK2) receptor. Bioorg. Med. Chem. Lett. 1998, 8, 1935–1940. [Google Scholar] [CrossRef]
  10. Katoumas, K.; Nikitakis, N.; Perrea, D.; Dontas, I.; Sklavounou, A. In vivo antineoplastic effects of the nsaid sulindac in an oral carcinogenesis model. Cancer Prev. Res. 2015, 8, 642–649. [Google Scholar] [CrossRef] [PubMed]
  11. Donovan, J.L.; Malcolm, R.J.; Markowitz, J.S.; DeVane, C.L. Chiral analysis of d- and l-modafinil in human serum: Application to human pharmacokinetic studies. Ther. Drug Monit. 2003, 25, 197–202. [Google Scholar] [CrossRef] [PubMed]
  12. Shreeram, S.S.; McDonald, T.; Dennison, S. Psychosis after ultrarapid opiate detoxification. Am. J. Psychiatry 2001, 158, 970. [Google Scholar] [CrossRef] [PubMed]
  13. Jacobs, R.T.; Miller, S.C.; Shenvi, A.B.; Ohnmacht, C.J.; Veale, C.A. Substituted piperidinabuty1 nitrogen-containing heterocyclic compounds and analogues thereof as neurokinin antagonist. U.S. Patent US 5,739,149, 4 April 1998. [Google Scholar]
  14. Simon, P.; Panissaud, C.; Costentin, J. The stimulant effect of modafinil on wakefulness is not associated with an increase in anxiety in mice. Psychopharmacology 1994, 114, 597–600. [Google Scholar] [CrossRef] [PubMed]
  15. Willie, J.T.; Renthal, W.; Chemelli, R.M.; Miller, M.S.; Scammell, T.E.; Yanagisawa, M.; Sinton, C.M. Modafinil more effectively induces wakefulness in orexin-null mice than in wild-type littermates. Neuroscience 2005, 130, 983–995. [Google Scholar] [CrossRef] [PubMed]
  16. Mitler, M.M.; Harsh, J.; Hirshkowitz, M.; Guilleminault, C. Long-term efficacy and safety of modafinil (provigil®) for the treatment of excessive daytime sleepiness associated with narcolepsy. Sleep Med. 2000, 1, 231–243. [Google Scholar] [CrossRef]
  17. Menza, M.A.; Kaufman, K.R.; Castellanos, A.M. Modafinil augmentation of antidepressant treatment in depression. J. Clin. Psychiatry 2000, 61, 378–381. [Google Scholar] [CrossRef] [PubMed]
  18. Taylor, F.B.; Russo, J. Efficacy of modafinil compared to dextroamphetamine for the treatment of attention deficit hyperactivity disorder in adults. J. Child Adolesc. Psychopharmacol. 2000, 10, 311–320. [Google Scholar] [CrossRef] [PubMed]
  19. Jenner, P.; Zeng, B.Y.; Smith, L.A.; Pearce, R.K.B.; Tel, B.; Chancharme, L.; Moachon, G. Antiparkinsonian and neuroprotective effects of modafinil in the mptp-treated common marmoset. Exp. Brain Res. 2000, 133, 178–188. [Google Scholar] [CrossRef] [PubMed]
  20. Chatterjie, N.; Stables, J.P.; Wang, H.; Alexander, G.J. Anti-narcoleptic agent modafinil and its sulfone: A novel facile synthesis and potential anti-epileptic activity. Neurochem. Res. 2004, 29, 1481–1486. [Google Scholar] [CrossRef] [PubMed]
  21. Kumar, R. Approved and investigational uses of modafinil. Drugs 2008, 68, 1803–1839. [Google Scholar] [CrossRef] [PubMed]
  22. Lafon, L. Acetamide derivatives. U.S. Patent US 4,177,290, 12 April 1979. [Google Scholar]
  23. Fornaroli, M.; Velardi, F.; Colli, C.; Baima, R. Process for the synthesis of modafinal. U.S. Patent US 20040106829, 2004. [Google Scholar]
  24. Castaldi, G.; Lucchini, V.; Tarquini, A. Process for the preparation of modafinal. U.S. Patent US 20050154063, 6 June 2006. [Google Scholar]
  25. De Risi, C.; Ferraro, L.; Pollini, G.P.; Tanganelli, S.; Valente, F.; Veronese, A.C. Efficient synthesis and biological evaluation of two modafinil analogues. Bioorg. Med. Chem. 2008, 16, 9904–9910. [Google Scholar] [CrossRef] [PubMed]
  26. Cao, J.; Prisinzano, T.E.; Okunola, O.M.; Kopajtic, T.; Shook, M.; Katz, J.L.; Newman, A.H. Structure-activity relationships at the monoamine transporters for a novel series of modafinil (2-[(diphenylmethyl)sulfinyl]acetamide) analogues. ACS Med. Chem. Lett. 2010, 2, 48–52. [Google Scholar] [CrossRef] [PubMed]
  27. Jung, J.C.; Lee, Y.; Son, J.Y.; Lim, E.; Jung, M.; Oh, S. Simple synthesis of modafinil derivatives and their anti-inflammatory activity. Molecules 2012, 17, 10446–10458. [Google Scholar] [CrossRef] [PubMed]
  28. Lari, A.; Karimi, I.; Adibi, H.; Aliabadi, A.; Firoozpour, L.; Foroumadi, A. Synthesis and psychobiological evaluation of modafinil analogs in mice. DARU J. Pharm. Sci. 2013, 21, 67. [Google Scholar] [CrossRef] [PubMed]
  29. Bogolubsky, A.V.; Moroz, Y.S.; Mykhailiuk, P.K.; Ostapchuk, E.N.; Rudnichenko, A.V.; Dmytriv, Y.V.; Bondar, A.N.; Zaporozhets, O.A.; Pipko, S.E.; Doroschuk, R.A.; et al. One-pot parallel synthesis of alkyl sulfides, sulfoxides, and sulfones. ACS Comb. Sci. 2015, 17, 348–354. [Google Scholar] [CrossRef] [PubMed]
  30. Sturala, J.; Bohacova, S.; Chudoba, J.; Metelkova, R.; Cibulka, R. Electron-deficient heteroarenium salts: An organocatalytic tool for activation of hydrogen peroxide in oxidations. J. Org. Chem. 2015, 80, 2676–2699. [Google Scholar] [CrossRef] [PubMed]
  31. Robertson, P., Jr.; Hellriegel, E.T. Clinical pharmacokinetic profile of modafinil. Clin. Pharmacokinet. 2003, 42, 123–137. [Google Scholar] [CrossRef]
  32. Wong, Y.N.; Wang, L.; Hartman, L.; Simcoe, D.; Chen, Y.; Laughton, W.; Eldon, R.; Markland, C.; Grebow, P. Comparison of the single-dose pharmacokinetics and tolerability of modafinil and dextroamphetamine administered alone or in combination in healthy male volunteers. J. Clin. Pharmacol. 1998, 38, 971–978. [Google Scholar] [CrossRef]
  33. Courvoisier, L.; Duret, G.; Graf, S.; Prat-Lacondemine, L.; Rebiere, F.; Sabourault, N. Process for enantioselective synthesis of signal enantiomers of modafinil by asymmetric oxidation. U.S. Patent 8,759,583, 24 June 2014. [Google Scholar]
  34. Prisinzano, T.; Podobinski, J.; Tidgewell, K.; Luo, M.; Swenson, D. Synthesis and determination of the absolute configuration of the enantiomers of modafinil. Tetrahedron 2004, 15, 1053–1058. [Google Scholar] [CrossRef]
  35. Osorio-Lozada, A.; Prisinzano, T.; Olivo, H.F. Synthesis and determination of the absolute stereochemistry of the enantiomers of adrafinil and modafinil. Tetrahedron 2004, 15, 3811–3815. [Google Scholar] [CrossRef]
  36. Hauck, W.; Adam, P.; Bobier, C.; Landmesser, N. Use of large-scale chromatography in the preparation of armodafinil. Chirality 2008, 20, 896–899. [Google Scholar] [CrossRef] [PubMed]
  37. Olivo, H.F.; Osorio-Lozada, A.; Peeples, T.L. Microbial oxidation/amidation of benzhydrylsulfanyl acetic acid. Synthesis of (+)-modafinil. Tetrahedron 2005, 16, 3507–3511. [Google Scholar] [CrossRef]
  38. Ternois, J.; Guillen, F.; Plaquevent, J.-C.; Coquerel, G. Asymmetric synthesis of modafinil and its derivatives by enantioselective oxidation of thioethers: Comparison of various methods including synthesis in ionic liquids. Tetrahedron 2007, 18, 2959–2964. [Google Scholar] [CrossRef]
  39. Stingl, K.A.; Weiß, K.M.; Tsogoeva, S.B. Asymmetric vanadium- and iron-catalyzed oxidations: New mild (R)-modafinil synthesis and formation of epoxides using aqueous H2O2 as a terminal oxidant. Tetrahedron 2012, 68, 8493–8501. [Google Scholar] [CrossRef]
  40. Held, F.E.; Grau, D.; Tsogoeva, S.B. Enantioselective cycloaddition reactions catalyzed by BINOL-derived phosphoric acids and N-triflyl phosphoramides: Recent advances. Molecules 2015, 20, 16103–16126. [Google Scholar] [CrossRef] [PubMed]
  41. Parmar, D.; Sugiono, E.; Raja, S.; Rueping, M. Complete field guide to asymmetric BINOL-phosphate derived bronsted acid and metal catalysis: History and classification by mode of activation; bronsted acidity, hydrogen bonding, ion pairing, and metal phosphates. Chem. Rev. 2014, 114, 9047–9153. [Google Scholar] [CrossRef] [PubMed]
  42. Terada, M. Binaphthol-derived phosphoric acid as a versatile catalyst for enantioselective carbon-carbon bond forming reactions. Chem. Commun. 2008, 35, 4097–4112. [Google Scholar] [CrossRef] [PubMed]
  43. Akiyama, T. Stronger bronsted acids. Chem. Rev. 2007, 107, 5744–5758. [Google Scholar] [CrossRef] [PubMed]
  44. Terada, M. Chiral phosphoric acids as versatile catalysts for enantioselective carbon–carbon bond forming reactions. Bull. Chem. Soc. Jpn. 2010, 83, 101–119. [Google Scholar] [CrossRef]
  45. Kampen, D.; Reisinger, C.M.; List, B. Chiral bronsted acids for asymmetric organocatalysis. Top. Curr. Chem. 2010, 291, 395–456. [Google Scholar] [CrossRef] [PubMed]
  46. Zamfir, A.; Schenker, S.; Freund, M.; Tsogoeva, S.B. Chiral BINOL-derived phosphoric acids: Privileged bronsted acid organocatalysts for C–C bond formation reactions. Org. Biomol. Chem. 2010, 8, 5262–5276. [Google Scholar] [CrossRef] [PubMed]
  47. Tsogoeva, S.; Zamfir, A. Towards a catalytic asymmetric version of the [3+2] cycloaddition between hydrazones and cyclopentadiene. Synthesis 2011, 2011, 1988–1992. [Google Scholar] [CrossRef]
  48. Serdyuk, O.V.; Zamfir, A.; Hampel, F.; Tsogoeva, S.B. Combining in situ generated chiral silicon Lewis acid and chiral brønsted acid catalysts for [3+2] cycloadditions: Cooperative catalysis as a convenient enantioselective route to pyrazolidines. Adv. Synth. Catal. 2012, 354, 3115–3121. [Google Scholar] [CrossRef]
  49. Zamfir, A.; Tsogoeva, S.B. Asymmetric hydrocyanation of hydrazones catalyzed by in situ formed o-silylated BINOL-phosphate: A convenient access to versatile alpha-hydrazino acids. Org. Lett. 2010, 12, 188–191. [Google Scholar] [CrossRef] [PubMed]
  50. Akiyama, T.; Itoh, J.; Yokota, K.; Fuchibe, K. Enantioselective mannich-type reaction catalyzed by a chiral bronsted acid. Angew. Chem. Int. Ed. 2004, 43, 1566–1568. [Google Scholar] [CrossRef] [PubMed]
  51. Uraguchi, D.; Terada, M. Chiral bronsted acid-catalyzed direct mannich reactions via electrophilic activation. J. Am. Chem. Soc. 2004, 126, 5356–5357. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Z.-M.; Zhao, H.; Li, M.-Q.; Lan, Y.-B.; Yao, Q.-B.; Tao, J.-C.; Wang, X.-W. Chiral phosphoric acid-catalyzed asymmetric oxidation of aryl alkyl sulfides and aldehyde-derived 1,3-dithianes: Using aqueous hydrogen peroxide as the terminal oxidant. Adv. Synth. Catal. 2012, 354, 1012–1022. [Google Scholar] [CrossRef]
  53. List, B.; Coric, I.; Liao, S. Process for the asymmetric oxidation of organic compounds with peroxides in the presence of a chiral acid catalyst. U.S. Patent WO 2013104605, 17 July 2013. [Google Scholar]
  54. Liao, S.; Coric, I.; Wang, Q.; List, B. Activation of H2O2 by chiral confined bronsted acids: A highly enantioselective catalytic sulfoxidation. J. Am. Chem. Soc. 2012, 134, 10765–10768. [Google Scholar] [CrossRef]
  55. Tsogoeva, S.; Wei, S.; Stingl, K.; Weiß, K. Thieme chemistry journal awardees—Where are they now? Bifunctional organocatalysis with N-formyl-l-proline: A novel approach to epoxide ring opening and sulfide oxidation. Synlett 2010, 2010, 707–711. [Google Scholar] [CrossRef]
  56. Stingl, K.A.; Tsogoeva, S.B. Recent advances in sulfoxidation reactions: A metal-free approach. Tetrahedron 2010, 21, 1055–1074. [Google Scholar] [CrossRef]
  57. Baudequin, C.; Zamfir, A.; Tsogoeva, S.B. Highly enantioselective organocatalytic formation of a quaternary carbon center via chiral bronsted acid catalyzed self-coupling of enamides. Chem. Commun. 2008, 38, 4637–4639. [Google Scholar] [CrossRef] [PubMed]
  58. Yamanaka, M.; Itoh, J.; Fuchibe, K.; Akiyama, T. Chiral bronsted acid catalyzed enantioselective mannich-type reaction. J. Am. Chem. Soc. 2007, 129, 6756–6764. [Google Scholar] [CrossRef] [PubMed]
  59. Goeddel, D.; Shu, L.; Yuan, Y.; Wong, O.A.; Wang, B.; Shi, Y. Effective asymmetric epoxidation of styrenes by chiral dioxirane. J. Org. Chem. 2006, 71, 1715–1717. [Google Scholar] [CrossRef]
  60. Held, F.E.; Wei, S.; Eder, K.; Tsogoeva, S.B. One-pot route to β-adrenergic blockers via enantioselective organocatalysed epoxidation of terminal alkenes as a key step. RSC Adv. 2014, 4, 32796–32801. [Google Scholar] [CrossRef]
  61. Aoki, M.; Seebach, D. Preparation of tadooh, a hydroperoxide from taddol, and use in highly enantioface- and enantiomer-differentiating oxidations. Helv. Chim. Acta 2001, 84, 187–207. [Google Scholar] [CrossRef]
Figure 1. Examples of chiral biologically active sulfoxides.
Figure 1. Examples of chiral biologically active sulfoxides.
Symmetry 09 00088 g001
Figure 2. The hippocampal activator in its racemic (modafinil) and optically pure (Armodafinil®) forms.
Figure 2. The hippocampal activator in its racemic (modafinil) and optically pure (Armodafinil®) forms.
Symmetry 09 00088 g002
Scheme 1. Proposed Lewis base/Brønsted acid catalysis—BINOL-phosphates as bifunctional organocatalysts in the oxidation of sulfide with hydrogen peroxide as an oxidant.
Scheme 1. Proposed Lewis base/Brønsted acid catalysis—BINOL-phosphates as bifunctional organocatalysts in the oxidation of sulfide with hydrogen peroxide as an oxidant.
Symmetry 09 00088 sch001
Scheme 2. Application of a dipeptide-based ligand X to the biomimetic iron-catalyzed synthesis of (R)-modafinil.
Scheme 2. Application of a dipeptide-based ligand X to the biomimetic iron-catalyzed synthesis of (R)-modafinil.
Symmetry 09 00088 sch002
Table 1. BINOL-phosphate catalyzed oxidation of thioanisole derivatives.
Table 1. BINOL-phosphate catalyzed oxidation of thioanisole derivatives.
Symmetry 09 00088 i001
EntryProductBINOL-PhosphateYield, a %ee, % (S)
11aII4410 b
21aII d2610 b
31aIII9820 b
41aI6836 b
51aIVTracesn.d.
61aV7330 b
71aTracesn.d.
81bVI1133 c
91bVI e1459 c
101bVIITraces18 c
a Yield of isolated product; b determined by HPLC (OD, n-hexanes/iPrOH (93:7), flow 1.0 mL/min); c determined by HPLC (IA, n-hexanes/iPrOH (90:10), flow 1.0 mL/min); d 10 mol % catalyst loading; e c (educt) = 0.8 mol/l; H2O2 = 0.27 equiv. For details of the reaction procedures and product characterizations, see Supplementary Information.
Table 2. Organocatalytic synthesis of modafinil.
Table 2. Organocatalytic synthesis of modafinil.
Symmetry 09 00088 i002
EntryCatalystTime, hYield, a %ee, b % (R)
1V24>9910
2V c24>997
3V1261rac
4 dV12>9913
5VIII243826
6IX e966616
a Yield of isolated product; b determined by chiral HPLC (AS, n-hexanes/iPrOH (60:40), flow 0.9 mL/min, 25 °C, 31 bar); c T = 0 °C; d c (sulfide) = 0.8 mol/l; e without H2O2, 1.5 equiv of IX, T = −30 oC > room temperature (RT), tetrahydrofuran (THF). For details of the reaction procedures and product characterizations, see Supplementary Information.

Share and Cite

MDPI and ACS Style

Held, F.E.; Stingl, K.A.; Tsogoeva, S.B. Synthesis of (R)-Modafinil via Organocatalyzed and Non-Heme Iron-Catalyzed Sulfoxidation Using H2O2 as an Environmentally Benign Oxidant. Symmetry 2017, 9, 88. https://doi.org/10.3390/sym9060088

AMA Style

Held FE, Stingl KA, Tsogoeva SB. Synthesis of (R)-Modafinil via Organocatalyzed and Non-Heme Iron-Catalyzed Sulfoxidation Using H2O2 as an Environmentally Benign Oxidant. Symmetry. 2017; 9(6):88. https://doi.org/10.3390/sym9060088

Chicago/Turabian Style

Held, Felix E., Kerstin A. Stingl, and Svetlana B. Tsogoeva. 2017. "Synthesis of (R)-Modafinil via Organocatalyzed and Non-Heme Iron-Catalyzed Sulfoxidation Using H2O2 as an Environmentally Benign Oxidant" Symmetry 9, no. 6: 88. https://doi.org/10.3390/sym9060088

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