Next Article in Journal
Core Oligosaccharide of Plesiomonas shigelloides PCM 2231 (Serotype O17) Lipopolysaccharide — Structural and Serological Analysis
Previous Article in Journal
Cytotoxic Effect of Clerosterol Isolated from Codium fragile on A2058 Human Melanoma Cells
Previous Article in Special Issue
Marine Cyanobacteria Compounds with Anticancer Properties: A Review on the Implication of Apoptosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Novel One-Pot Green Synthesis of Indolizines Biocatalysed by Candida antarctica Lipases

1
Department of Chemistry, Physics and Environment, Faculty of Science and Environment, "Dunarea de Jos" University of Galati, 111 Domneasca Street, Galati 800201, Romania
2
Department of Food Science, Food Engineering and Applied Biotechnology, Faculty of Food Science and Engineering, "Dunarea de Jos" University of Galati, 111 Domneasca Street, Galati 800201, Romania
3
Pharmacochimie Moléculaire, UMR 5063 & FR 2607, CNRS/Université de Grenoble, 38041 Grenoble cedex 9, France
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2013, 11(2), 431-439; https://doi.org/10.3390/md11020431
Submission received: 29 December 2012 / Revised: 29 January 2013 / Accepted: 29 January 2013 / Published: 6 February 2013
(This article belongs to the Special Issue Bioactive Compounds from Marine Microorganisms)

Abstract

:
Marine microorganisms are of considerable interest as a promising source of enzymes with unsuspected potentials as catalysts for chemical synthesis. We describe here an efficient method for one-pot indolizine synthesis that has been developed using lipase A and lipase B from Candida antarctica as biocatalysts. As showed by HPLC/MS analysis, the yield in indolizines was higher in the presence of the biocatalyst than in absence of enzyme. Lipase A, from Candida antarctica, showed high catalytic activity and selectivity for the cycloaddition reactions. When the reactions were performed under ultrasound irradiation, the Candida antarctica lipase catalyzed reactions yielded pure indolozines, in good yields and in very short time.

1. Introduction

The indolizine ring is an important structural system frequently found in natural products and has been used as an important skeleton in pharmaceutics because of its interesting and promising biological properties [1,2,3]. Both natural and synthetic derivatives have been identified as anticancer, anti-tuberculosis, analgesic, and antioxidant agents. Several data point also to the potential of indolizine derivatives as antimicrobial agents [4,5], calcium entry blockers [6], antioxidants [7] or inhibitors of 15-lipooxygenase [8]. As a result, different approaches have been reported in the literature for their synthesis [3,9]. The cycloaddition of ylides with alkynes emerges as one of the most general, and widely used reactions in indolizine ring formation [10]. However, long reaction time, elevated thermal conditions, expensive and/or toxic metals catalysts, and environmentally hazardous organic solvents are generally required.
We are interested in the synthesis of a variety of indolizine derivatives, and in the study of their biological, physical, and spectroscopic properties [11,12,13]. More recently we have initiated a project dealing with the use of microorganisms, or isolated enzymes, to design greener alternatives to known chemical processes.
Indeed, enzymes such as lipases are well-known both at the lab-bench and for industrial processes, and their use is widespread, especially when regio- or stereoselectivity is required. Microorganisms from marine sources have been, so far, little investigated in the search of new biocatalysts [14,15]. However, organisms originating from the Antarctic seawater are known to produce enzymes that are able to function at extremely low temperatures, and are therefore useful as novel tools in biotechnology, particularly for pharmaceuticals and the food industry. Antarctic yeast Candida antarctica is probably one of the most useful microorganism, which has been cited in numerous patents, and has also been used in a variety of commercial applications. The lipases A and B isolated from Candida antarctica (and their immobilized forms CAL A and CAL B) catalyze a variety of chemical reactions including ester hydrolysis, transesterification, and amide synthesis [16,17,18]. These enzymes are used advantageously in organic synthesis or biotechnology because of their low cost, good stability in organic solvents, absence of co-factors requirement, and large pH operating range. CAL A and CAL B are able to work with various types of substrates, and may be stereo-, chemo-, and regioselective.
We report here an original and efficient method for the synthesis of substituted bis-indolizines via an enzyme catalyzed three-component reaction starting from 4,4′-bipyridine, α-bromo carbonyl reagents and activated alkynes. The reactions were carried out in an aqueous buffer solution in the presence of Candida antarctica enzymes (CAL A or CAL B). Our one-pot transformation involved the cycloaddition of a dipolarophile with a stable ylide, issued from the 4,4′-bipyridinium diquaternary salt, formed in situ.

2. Results and Discussion

Several methods have been developed for the synthesis of indolizine derivatives, including recent one-pot multicomponent reactions [19,20]. In a previous paper, we reported the cycloaddition reaction of 4,4′-bipyridinium diquaternary salts with various alkynes that gave the corresponding fluorescent dyes in good yields [11,12]. The conditions are presented in Figure 1 as the classical reaction.
Figure 1. Synthesis of bisindolizines.
Figure 1. Synthesis of bisindolizines.
Marinedrugs 11 00431 g001
We designed the first example of a one-pot biocatalytic bis-indolizine ring formation and compared this process to the classical reaction. The reactions were realized starting, as previously reported, from 4,4′-bipyridine, halide derivatives (ω-bromacetophenones), and ethyl propiolate, in a pH 7 phosphate buffer solution, and stirred at 40 °C, either using a thermoshaker or under ultrasound activation (Figure 1). The one-pot reactions were carried out in the presence and in the absence of lipases to evidence the biocatalytic effect. Both classical and catalyzed reactions have given the same indolizine products 1ac.
Indeed, the choice of the reaction parameters (nature of the solvent, temperature, time, and substrate concentration) is an important step in biocatalysis. Water is the solvent of choice as the enzyme-bound water is essential for catalysis and serves as a lubricant for the enzyme. The temperature can be used to influence the viscosity of the medium, which in turn influences the solubility of substrates and products, enzyme activity, and even the biocatalyst denaturation. The molar ratio of reagents also modulates the product yield. Thus, for the in situ formation of the 4,4′-bipyridinium diquaternary salts from 4,4′-bipyridine and ω-bromacetophenones, and for the subsequent cycloaddition, excesses of alkylating agent (ω-bromacetophenones) and dipolarophile (ethyl propiolate) were used. The selected reaction conditions for the cycloaddition using Candida antarctica enzymes were using water as the solvent and excess (three equivalents) of ω-bromacetophenones and ethyl propiolate.
The one-pot transformations were monitored by high performance liquid chromatography (HPLC) using a system equipped with a diode array detector. Figure 2a displays the chromatogram profiles of the reactions of the 4,4′-bipyridine with ω-bromacetophenone and ethyl propiolate, with and without enzymes, after 48 h of stirring at 40 °C in a thermoshaker. The chromatograms showed two important peaks: a peak eluting at 1.8 min, probably corresponding to the intermediary diquaternary salt, and a peak at 3.4 min, identified as the final indolizine 1a. This last peak showed, in its corresponding UV-Vis spectra (Figure 2b), several intense absorption bands (231 nm, 277 nm, 352 nm, 411 nm) that are characteristics of the indolizine ring. This observation confirmed that the cycloaddition reaction occurred.
Figure 2. Hplc chromatograms of the one-pot reaction mixtures after 48 h of shaking in a thermoshaker (a) and UV-VIS spectra (b) corresponding to the compound eluting at 3.4 min and corresponding to the indolizine 1a.
Figure 2. Hplc chromatograms of the one-pot reaction mixtures after 48 h of shaking in a thermoshaker (a) and UV-VIS spectra (b) corresponding to the compound eluting at 3.4 min and corresponding to the indolizine 1a.
Marinedrugs 11 00431 g002
The chromatograms also showed that the concentration in indolizinic compound was greater in the presence of the biocatalysts. The conversion ratio in indolizine 1a, calculated from HPLC chromatograms, is illustrated in Figure 3.
Figure 3. Conversion percentage for enzymatic and non-enzymatic one-pot synthesis of indolizine 1a.
Figure 3. Conversion percentage for enzymatic and non-enzymatic one-pot synthesis of indolizine 1a.
Marinedrugs 11 00431 g003
After 48 h of reaction in phosphate buffer solution at 40 °C, it was observed that starting materials were converted to indolizine compounds in all samples, but better yields of indolizines 1ac were obtained when CAL A and CAL B were used as biocatalysts. Comparing the two enzymes, CAL A showed better activity for the cycloaddition reaction with higher yields in the desired products, 1ac may be due to the lower thermostability and higher acidity of CAL B compared with CAL A [21]. The good yield observed in the presence of the Candida antarctica enzymes, and the poor results obtained in the absence of enzymes, were attributed to the ability of the enzymes to perform the cycloaddition reactions under these buffer conditions.
The enzymatic and non-catalyzed one-pot syntheses were then carried out with stirring and heating at 40 °C in an ultrasound bath, for two hours. As shown in Figure 4, ultrasounds activated the reactions, as similar conversion rates were obtained but with reaction times significantly reduced (from 48 h to 2 h).
Figure 4. Conversion percentage to indolizine 1a in enzymatic and non-enzymatic one-pot synthesis, carried out with and without ultrasounds.
Figure 4. Conversion percentage to indolizine 1a in enzymatic and non-enzymatic one-pot synthesis, carried out with and without ultrasounds.
Marinedrugs 11 00431 g004
These conditions were then applied to the lipase catalyzed synthesis of indolizines 1b and 1c, with similar results as for indolizine 1a. In the HPLC analysis, indolizines 1b and 1c appeared at 6.45 min and at 6.03 min respectively.
As a conclusion, this biocatalytic method used inexpensive and readily available solvents (water), and allowed N-alkylation of 4,4′-bipyridine and subsequent cycloaddition reaction to occur in one-pot, thus limiting the number of experimental steps. The reactions took place in a more environmentally friendly way than the classical two-step reactions. It is worth mentioning that both classical and enzyme catalyzed reactions gave the same indolizine products 1ac, thus indicating that the ester group present in 1ac (R=COOEt) is not a substrate for the studied lipases in these conditions. Indeed, in a separate experiment, we checked that the ester group present in indolizine 1a was not hydrolysed in the presence of CAL A.
This procedure represents an eco-friendly regioselective approach to the indolizine core formation.
Our synthetic approach, which involves the cycloaddition of in situ generated ylides, in aqueous medium, turned out to be the first example of a one-pot bio-preparation for indolizines.

3. Experimental Section

3.1. Enzymes and Chemicals

The lipases used were immobilized lipases from Candida antarctica, CAL A (Novozyme 735, 1.6 U/mg) and CAL B (Novozyme 435, ≥10000 U/g). All the other reagents used for synthesis were purchased from Aldrich, Fluka, and Merck companies. The solvents used in high performance liquid chromatography are of analytic grade from Merck®. A Milli-Q System® (Bedford, MA, USA) was used to purify the water.

3.2. Equipment and Chemical Analysis

The biocatalyzed reactions were performed in an orbital flasks thermoshaker (Biosan) or in an ultrasound bath (Bandelin Sonrex Digitech). The progress of each reaction was monitored using an HPLC Thermo Scientific system, using a C18 BDS column and acetonitrile as mobile phase, with a flow rate of 1 mL/min and injection volume of 10 μL. Melting points were recorded with a Büchi Melting Point B-540. 1H NMR and 13C NMR spectra were recorded with a Bruker 400 Ultrashield (400 MHz) spectrometer operating at room temperature. Deuterated CDCl3 was used as solvent. Abbreviations for data quoted are: s, singlet; d, doublet; t, triplet; q, quartet; dd, doublet of doublets; m, multiplet. IR spectra were recorded from 4000 to 650 cm−1 with a Perkin-Elmer Spectrum 100 instrument by total reflectance on a CdSe crystal. Elemental analyses (C, H, N) were performed with a Fisons Instruments 1108 CHNS–O elemental analyzer. The atmosphere pressure chemical ionization (APCI) mass spectra were measured on Thermo Scientific HPLC/MSQ Plus.

3.3. Procedure for One-Pot Non-Enzymatic Reactions

The 4,4′-bipyridine (1.2 mmol), halide derivatives (3.6 mmol), and ethyl propiolate (3.6 mmol), were added to Erlenmeyer flasks (50 mL) containing phosphate buffer solutions (250 mM, pH 7, 12 mL). The mixtures were stirred in a rotary shaker (200 rpm) at 40 °C for 48 h or in ultrasonic bath for 2 h.

3.4. Procedure for One-Pot Enzymatic Reaction

To a 50 mL Erlenmeyer flask containing phosphate buffer solutions (250 mM, pH 7, 12 mL) and enzyme (CAL A or CAL B) (10 mg) was added the 4,4′-bipyridine (1.2 mmol), halide derivatives (3.6 mmol), and ethyl propiolate (3.6 mmol). The reaction mixture was stirred in a rotary shaker (200 rpm) at 40 °C until maximum conversion of the starting material (48 h) or under ultrasounds (2 h).

3.5. Procedure for Extraction of Indolizines

All reactions were monitored by HPLC and, after the appropriate conversion ratio, the mixture was filtered and the aqueous phase was extracted using chloroform. The yellow organic phase was dried over MgSO4 and then filtered. The solvent was evaporated under reduced pressure and the residue was precipitated with methanol giving the bis-indolizine products in 67% (1a), 60% (1b), 66% (1c) yields for CAL A and 28% (1a), 25% (1b), 27% (1c) yields for CAL B.

3.6. Analysis of the Final Indolizine Products 1a–c

The products of biocatalyzed reactions were analyzed by IR, NMR, MS, elemental analysis and the obtained data were identical with those of the substituted 7,7′-bis-indolizines previously obtained and published by us [11].
1,1′-diethyldicarboxylate-3,3′-benzoyl-7,7′-bis(indolizine) (1a): yellow crystals, mp 273–275 °C; IR (ATR, cm−1): 3139 (l, CHarom); 2979 (l, CHalif); 1700 (s, C=Oester); 1645 (s, C=O); 1524, 1512, 1479, 1459, 1427 (s, C=N, C=Carom); 1341 (s, COester) 1216, 1167 (s, C–O–C); 1071 (s, C–N); 1H-NMR (400 MHz, CDCl3, TMS) δ/ppm: 9.98 (d, J = 7.5, 2H); 8.77(d, J = 1.2, 2H); 7.78-7.88 [m, 6H: 4H, 2H (7.82)]; 7.42–7.73 [m, 8H: 4H, 2H, 2H (7.52, J = 7.5, J = 1.2)]; 4,4 (q, J = 7.2, 4H); 1,43 (t, J = 7.2, 6H); 13C-NMR (400 MHz, CDCl3, TMS δ/ppm): 185.67 (2C=O), 163.97 (2C=Oester), 139.78 (2C), 139.70 (2C), 136.61 (2C), 131.74 (2CH), 129.53 (2CH), 129.33 (2CH), 129.05 (4CH), 128.52 (4CH), 122.90 (2C), 116.96 (2C), 113.53 (2CH), 107.55 (2C), 60.37 (2CH2), 14.59 (2CH3); MS (APCI+): m/z: 585 [M + H]+; anal. C 73.72; H 5.14; N 4.76%, calcd. for C36H28N2O6, C 73.96, H 4.83, N 4.79%.
1,1′-diethyldicarboxylate-3,3′-bis(p-methoxy-benzoyl)-7,7′-bis(indolizine) (1b): yellow crystals, mp 290–291 °C; IR (ATR, cm−1): 3140 (l, CHarom); 2977 (l, CHalif); 1700 (s, C=Oester); 1665 (s, C=O); 1597, 1508, 1463 (s, C=N, C=Carom); 1339 (s, COester) 1249, 1220, 1167 (s, C–O–C); 1077 (s, C–N); 1H-NMR (400 MHz, CDCl3, TMS) δ/ppm: 9.98 (d, J = 7.5, 2H); 8.80(d, J = 1.2, 2H); 7.79 (d, J = 7.5, 2H); 7.76 (s, 2H); 7.33 (d, J = 8.8, 4H); 7.03 (d, J = 8.8, 4H); 4.43 (q, J = 7.2, 4H); 3,93 (s, 6H) 1,43 (t, J = 7.2, 6H); 13C-NMR (400 MHz, CDCl3, TMS δ/ppm ): 186.06 (2C=O), 165.28 (2C=Oester), 163.71 (2C), 138.57 (2C), 136.57 (4CH), 132.42 (2C), 131.81 (2CH), 129.04 (2CH), 122.2 (4CH), 122 (2CH), 117.72 (2CH), 114 (4CH), 113.98 (2CH), 107 (2C), 61.09 (CH2), 56,03 (OCH3), 14,73 (CH3); MS (APCI+): m/z: 645 [M + H]+; anal. C 70.92; H 4.88; N 4.28%, calcd. for C38H32N2O8, C 70.80, H 5.00, N 4.35%.
1,1′-diethyldicarboxylate-3,3′-bis(p-nitro-benzoyl)-7,7′-bis(indolizine) (1c): yellow crystals, mp 271–272 °C; IR (ATR, cm−1): 3359 (l, CHarom); 2978 (l, CHalif); 1700 (s, C=Oester); 1656 (s, C=O); 1464, 1428 (s, C=N, C=Carom); 1521, 1338 (s, NO2); 1208, 1169 (s, C–O–C); 1077 (s, C–N); 1H-NMR (400 MHz, CDCl3, TMS) δ/ppm: 10.13 (d, J = 7.5, 2H); 8.99(d, J = 1.2, 2H); 8.03 (d, J = 8.8, 4H); 7.8 (d, J = 8.8, 4H); 7.48 (d, J = 7.5, 2H ); 7.42 (s, 2H); 4.4 (q, J = 7.2, 4H); 1.43 (t, J = 7.2, 6H); 13C-NMR (400 MHz, CDCl3, TMS δ/ppm ): 184.15 (2C=O), 164.52 (2C=Oester), 150.16 (2C), 143.06 (2C); 138.14 (2C); 132.60 (2C); 129.86 (4CH); 129.14 (2CH); 127.46 (2CH); 123.52 (4CH); 122.74 (2C); 118.28 (2CH); 113.07 (2CH); 108.16 (2Cquat); 60.53 (2CH2); 14.29 (2CH3); MS (APCI+): m/z: 675 [M + H]+; anal. C 64.15; H 4.11; N 8.26%, calcd. for C36H26N4O10, C 64.09, H 3.88, N 8.31%.

4. Conclusions

Our synthetic approach is the first example for biocatalyzed one-pot synthesis for indolizines, which involves cycloadditions of ethyl propiolate with the in situ generated ylides, in aqueous media.
The best results were obtained when the Candida antarctica lipase catalyzed the biotransformation of 4,4′-bipyridine, halide derivatives, and ethyl propiolate to form indolizine compounds. A better yield of indolizine was obtained when CAL A was used as a biocatalyst. The catalytic reactions with lipase from Candida antarctica produced almost pure indolizines with good yields and in very short time under ultrasound activation. As mentioned above, the lipase catalyzed synthesis was performed in water, a green solvent compared to the organic solvents usually required for the classical methodologies.
The increasing importance of safety, health, and environmental issues as well as initiatives in green and sustainable chemistry, white and industrial biotechnology, and the new perspectives of molecular economy plainly justify the work done to develop this biocatalytic approach using enzymes for indolizine synthesis. Further developments of this methodology are under investigations. In particular, the process will be extended to the synthesis of variously substituted mono- and bis-indolizines of biological interest.

Acknowledgements

This work was supported by a grant of the Romanian National Authority for Scientific Research, CNCS—UEFISCDI, project number PN-II-ID-PCE-2011-3-0226.

References

  1. Flitsch, W. Comprehensive Heterocyclic Chemistry; Katritzky, A.R., Rees, C.W., Eds.; Pergamon Press: Oxford, UK, 1984; Volume 4, p. 443. [Google Scholar]
  2. Michael, J.P. Indolizidine and quinolizidine alkaloids. Nat. Prod. Rep. 1989, 6, 139–165. [Google Scholar]
  3. Singh, G.S.; Mmatli, E.E. Recent progress in synthesis and bioactivity studies of indolizines. Eur. J. Med. Chem. 2011, 46, 5237–5257. [Google Scholar] [CrossRef]
  4. Gundersen, L.-L.; Charnock, C.; Negussie, A.H.; Rise, F.; Teklu, S. Synthesis of indolizine derivatives with selective antibacterial activity against Mycobacterium tuberculosis. Eur. J. Pharm. Sci. 2007, 30, 26–35. [Google Scholar]
  5. Medda, S.; Jaisankar, P.; Manna, R.; Pal, B.; Giri, V.S.; Basu, M.K. Phospholipid microspheres: A novel delivery mode for targeting antileishmanial agent in experimental leishmaniasis. J. Drug Target 2003, 11, 123–128. [Google Scholar] [CrossRef]
  6. Gupta, S.P.; Mathur, A.N.; Nagappa, A.N.; Kumar, D.; Kumaran, S. A quantitative structure-activityrelationship study on a novel class of calcium-entry blockers: 1-[(4-(Aminoalkoxy)phenyl)sulphonyl]indolizine. Eur. J. Med. Chem. 2003, 38, 867–873. [Google Scholar] [CrossRef]
  7. Ostby, O.B.; Dalhus, B.; Gundersen, L.-L.; Rise, F.; Bast, A.; Haenen, G.R.M.M. Synthesis of 1-substituted 7-cyano-2,3-diphenylindolizines and evaluation of antioxidant properties. Eur. J. Org. Chem. 2000, 9, 3763–3770. [Google Scholar]
  8. Teklu, S.; Gundersen, L.-L.; Larsen, T.; Malterud, K.E.; Rise, F. Indolizine 1-sulfonates as potent inhibitors of 15-lipoxygenase from soybeans. Bioorg. Med. Chem. 2005, 13, 3127–3139. [Google Scholar] [CrossRef]
  9. Prostakov, N.S.; Baktibaev, O.B. Indolizines. Russ. Chem. Rev. 1975, 44, 748–766. [Google Scholar]
  10. Kakehi, A. Reactions of pyridinium N-ylides and their related pyridinium salts. Heterocycles 2012, 85, 1529–1577. [Google Scholar] [CrossRef]
  11. Druta, I.; Dinica, R.; Bacu, E.; Humelnicu, I. Synthesis of 7,7′-bis-Indolizines by the reaction of 4,4′-bipyridinium-ylides with activated alkynes. Tetrahedron 1998, 54, 10811–10818. [Google Scholar] [CrossRef]
  12. Dinica, R.M.; Pettinari, C. Synthesis of substituted, 7,7′-bis-Indolizines via 1,3-dipolar cycloaddition under microwave irradiation. Synlett 2000, 7, 1013–1015. [Google Scholar]
  13. Furdui, B.; Dinica, R.M.; Druta, I.; Demeunynck, M. Improved synthesis of cationic pyridinium-substituted indolizines. Synthesis 2006, 16, 2640–2642. [Google Scholar]
  14. Zhang, C.; Kim, S.-K. Research and application of marine microbial enzymes: Status and prospects. Mar. Drugs 2010, 8, 1920–1934. [Google Scholar] [CrossRef]
  15. Sanchez-Amat, A.; Solano, F.; Lucas-Elío, P. Finding new enzymes from bacterial physiology: A successful approach illustrated by the detection of novel oxidases in Marinomonas mediterranea. Mar. Drugs 2010, 8, 519–541. [Google Scholar] [CrossRef]
  16. Ferreira, H.V.; Rocha, L.C.; Severino, R.P.; Porto, A.L.M. Syntheses of enantiopure aliphatic secondary alcohols and acetates by bioresolution with lipase B from Candida antarctica. Molecules 2012, 17, 8955–8967. [Google Scholar] [CrossRef]
  17. Raminelli, C.; Kagohara, E.; Pellizari, V.H.; Comasseto, J.V.; Andrade, L.H.; Porto, A.L.M. Biotransformations of Mannich bases and propiophenones by Brazilian microorganisms and enzymatic resolution of phenylpropanols by lipase from Candida antarctica (Novozym 435). Enzyme Microb. Technol. 2007, 40, 362–369. [Google Scholar] [CrossRef]
  18. Cardillo, G.; Gennari, A.; Gentilucci, L.; Mosconi, E.; Tolomelli, A.; Troisi, S. Synthesis of chiral non-racemic intermediates and Arg-Gly-Asp mimetics by CaLB-catalyzed resolution. Tetrahedron Asymmetry 2010, 2, 96–102. [Google Scholar]
  19. Georgescu, E.; Caira, M.R.; Georgescu, F.; Draghici, B.; Popa, M.M.; Dumitrascu, F. One-pot, three-component synthesis of a library of new pyrrolo[1,2-a]quinoline derivatives. Synlett 2009, 11, 1795–1799. [Google Scholar]
  20. Bora, U.; Saikia, A.; Boruah, R.C. A novel microwave-mediated one-pot synthesis of indolizines via a three-component reaction. Org. Lett. 2003, 5, 435–438. [Google Scholar] [CrossRef]
  21. Siddiqui, S.K.; Cavicchioli, R. Improved thermal stability and activity in the cold-adapted lipase B from Candida antarctica following chemical modification with oxidized, polysaccharides. Extremophiles 2005, 9, 471–476. [Google Scholar] [CrossRef]
  • Samples Availability: Available from the authors.

Share and Cite

MDPI and ACS Style

Dinica, R.M.; Furdui, B.; Ghinea, I.O.; Bahrim, G.; Bonte, S.; Demeunynck, M. Novel One-Pot Green Synthesis of Indolizines Biocatalysed by Candida antarctica Lipases. Mar. Drugs 2013, 11, 431-439. https://doi.org/10.3390/md11020431

AMA Style

Dinica RM, Furdui B, Ghinea IO, Bahrim G, Bonte S, Demeunynck M. Novel One-Pot Green Synthesis of Indolizines Biocatalysed by Candida antarctica Lipases. Marine Drugs. 2013; 11(2):431-439. https://doi.org/10.3390/md11020431

Chicago/Turabian Style

Dinica, Rodica Mihaela, Bianca Furdui, Ioana Otilia Ghinea, Gabriela Bahrim, Simon Bonte, and Martine Demeunynck. 2013. "Novel One-Pot Green Synthesis of Indolizines Biocatalysed by Candida antarctica Lipases" Marine Drugs 11, no. 2: 431-439. https://doi.org/10.3390/md11020431

Article Metrics

Back to TopTop