Next Article in Journal
Predicting Protein-Protein Interactions Using BiGGER: Case Studies
Next Article in Special Issue
Efficient Enzyme-Free Biomimetic Sensors for Natural Phenol Detection
Previous Article in Journal
The Complete Chloroplast Genome Sequence of the Medicinal Plant Swertia mussotii Using the PacBio RS II Platform
Previous Article in Special Issue
Chemo-Enzymatic Synthesis of Chiral Epoxides Ethyl and Methyl (S)-3-(Oxiran-2-yl)propanoates from Renewable Levoglucosenone: An Access to Enantiopure (S)-Dairy Lactone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemo-Enzymatic Synthesis of Oligoglycerol Derivatives

1
Department of Chemistry, University of Delhi, Delhi 110007, India
2
Université de Technologie de Compiègne (UTC), CS 60319, Compiègne Cedex 60203, France
3
Institut für Chemie und Biochemie, Freie Universität Berlin, Takustraße 3, Berlin 14195, Germany
*
Authors to whom correspondence should be addressed.
Molecules 2016, 21(8), 1038; https://doi.org/10.3390/molecules21081038
Submission received: 20 July 2016 / Revised: 4 August 2016 / Accepted: 5 August 2016 / Published: 9 August 2016

Abstract

:
A cleaner and greener method has been developed and used to synthesize 14 different functionalized oligomer derivatives of glycerol in moderate 29%–39% yields over three steps. After successive regioselective enzymatic acylation of the primary hydroxyl groups, etherification or esterification of the secondary hydroxyl groups and chemoselective enzymatic saponification, the target compounds can efficiently be used as versatile building blocks in organic and supramolecular chemistry.

Graphical Abstract

1. Introduction

Glycerol is a considered green feedstock due to its bioavailability. An enormous quantity of glycerol is being produced by the oleochemical and biodiesel industry and is used as a base chemical for the production of value added products [1,2,3,4,5,6,7,8,9,10,11,12,13]. It has enormous applications in the food industry, pharmaceutical and personal care preparations. Oligomerization of glycerol and the physical and chemical properties of its oligomers have also been well studied. In particular the low molecular weight oligomers such as di-, tri-, and tetraglycerol are more hydrophilic than higher ones and thus have better solubility in polar solvents. These oligomers are used in personal care formulations for their mild humectant properties and ability to enhance fragrance/flavor impact and longevity. Glycerol oligomers also act as plasticizers in PVA films and starch-based biodegradable thermoplastic compositions. The oligomer derivatives have been explored for various applications, e.g., thickners, emulsifiers, antifogging agents, etc. [14,15,16,17,18,19]. Glycerol oligomers may also be considered as superior building blocks for polymerization or polycondensation reactions in comparison to the monomer as the latter leads to low molecular weight reaction products which have a considerable effect on the properties of the macromolecular compounds. The preparation of polymers based on the conversion of oligomerss into macromolecular compounds is emerging as an interesting line of development in the synthesis of polymers [20,21]. Such a method is associated with the synthesis of oligomers with reactive groups at the ends of the molecules.
In order to provide more extensive data for a wider structure-activity relationship (SAR), analogues having 2-O-alkyl and 2-O-acyl groups have to be synthesized selectively via full or partial green chemistry. Unfortunately, glycerol and its oligomers have two types of hydroxyl groups having similar pKa and consequently, regioselective differentiation of the primary and secondary hydroxyl groups is difficult. In this regard, a protection-functionalization-deprotection strategy has to be realized. Our preliminary work reported that an enzymatic method using immobilized Candida antarctica lipase (Novozym 435) [22,23,24] can distinguish the primary and secondary hydroxyl groups in glycerol and polyglycerol moieties to synthesize a wide variety of polymeric and dendritic architectures for biomedical applications [25,26,27,28,29]. Starting from dimers and trimers of glycerol, regio-selective enzymatic protection of the primary hydroxyl groups by acetylation could be envisaged followed by 2-O-alkylation or 2-O-acylation of the secondary hydroxyl groups, then followed by deprotection of primary hydroxyls. Herein, we report the synthesis of 14 new building blocks based on diglycerol and triglycerol wherein aromatic/azido groups have been incorporated in the secondary carbon via an ether or an ester so as to provide versatility besides facilitating the monitoring of their reactions and product purification. The presence of aromatic moiety may provide the possibility of additional π-π interactions in the macromolecules and thus controlling the aggregation phenomenon and encapsulation behavior. The incorporation of the azido group and an alkynyl group on the other hand provide a site for the 1,3-dipolar cycloaddition reaction under click conditions [30,31,32].

2. Results

Commercially available diglycerol and triglycerol were subjected to Novozym 435-catalyzed acylation by following the literature procedure [26]. Since the commercially available oligomers are not in pure form and rather are mixtures of glycerol, diglycerol and triglycerol, consequently, a mixture of products is obtained after acylation using Novozym 435 (20 wt %) and vinyl acetate in THF (Scheme 1).
The reaction progress was monitored by TLC (methanol/chloroform, 1:9, v/v) and on completion of the reaction the desired diacylated product was purified through column chromatography over silica gel using CHCl3/MeOH as eluent It should be noted that compounds 2 and 10 contain two and three stereocenters, respectively. In our hands, both of these compounds were obtained as a mixture of three and seven stereoisomers, respectively and this was confirmed by the observance of the peak multiplicity in the 13C-NMR spectrum of compounds 2 and 10 (see Supplementary Materials Figures S1 and S8). This suggests that the enzymatic conditions produced the protected glycerol derivatives regioselectively but not stereoselectively. In the 1H-NMR spectrum of compounds 2 and 10, the methyl protons of the acetyl group appeared at 2.08 ppm, whereas the methylene and methine protons appeared in the range of 3.43 to 4.24 ppm (Figure 1 and Figure 2).
The resulting diacyl oligomers 2 and 10 were further functionalized at the secondary hydroxyl via ether/ester linkage (Scheme 1). The 2-O-alkylation of compounds 2 and 10 was realized via Mitsunobu reaction using 4-hydroxybenzoic acid ethyl ester and furnished the two ethers 3 and 11 in 70% and 65% yield, respectively [33]. The 2-O-acylation of compounds 2 and 10 using benzoic acid derivative in the presence of EDC gave the corresponding esters 7 and 15 in 75% and 70% yield, respectively [25,34].
Furthermore, the azido group was also incorporated in a stepwise manner i.e., first carrying out mesylation of the secondary hydroxyl group followed by treatment with sodium azide in DMF. The formation of products 5 and 13 was monitored by the appearance of an azide peak in the IR spectra. Our conditions permitted us to produce regioselectively, in the secondary position, the ethers 3 and 11, esters 7 and 15 and azido derivatives 5 and 13. Saponification and transesterification were not observed, meaning that the regioselective protection of the primary hydroxyl groups supported the described nucleophilic conditions.
In the next step the resulting diacyl oligomer derivatives 3, 5, 7, 11, 13, and 15 were subjected to Novozym 435-catalyzed deacetylation in the presence of excess n-butanol in THF [35]. After column chromatography, the target compounds 4, 6, 8, 12, 14 and 16 were obtained in good yields. A high chemo-selectivity was observed for deacylation of compounds 3, 7, 11 and 15, the aromatic acid ester remained intact and only the selective hydrolysis of aliphatic acid ester was observed. Moreover, in our conditions, no transesterification of the aromatic esters was observed from the secondary hydroxyl group to the primary hydroxyl group. All the compounds were characterized by 1H-, 13C-NMR and HRMS analysis. The methine proton in all of these compounds undergoes a characteristic chemical shift on functionalization i.e., O-alkylation of secondary hydroxyl groups in compounds 4, 8, 12 and 16 led to a shift from 3.90–4.02 to 4.50–4.55 ppm (Figure 1 and Figure 2). However, on acylation of the secondary hydroxyl group, the corresponding methine proton undergoes a more significant shift to 5.20–5.44 ppm (Figure 1 and Figure 2). The azido compounds 6 and 14 exhibited a characteristic peak for azide group at about 2100 cm−1 in the IR spectrum. While in the compounds 8 and 16 the characteristic acetylinic proton was observed at 2.55–2.56 ppm (Figure 1 and Figure 2) in the 1H-NMR spectrum and the acetylinc (-C≡C-) moiety led to the observance of a peak in the rang 2100–2200 cm−1 in the IR spectrum.

3. Experimental Section

3.1. General Information

All the compounds were characterized by their physical and spectral data. Infrared spectra were recorded on a Perkin-Elmer FT-IR Model 9 Spectrophotometer (Perkin-Elmer, Singapore). The 1H- and 13C-NMR spectra (400 MHz/100.5 MHz) were recorded on Jeol-400 NMR spectrometer (Jeol, Tokyo, Japan) using TMS as an internal standard. The chemical shift values are measured on δ scale and the coupling constant values (J) are reported in Hz. The HRMS data were recorded on Agilent-6530, Q-TOF LCMS (Agilent, Singapore).
The diglycerol and triglycerol were obtained from Solvay Bruxelles (Solvay, Bruxelles, Belgium) and Sigma Aldrich (Saint Louis, MO, USA) respectively. All other chemicals and solvents used were purchased from the Spectrochem Pvt. Ltd. (Mumbay, India) and SD Fine Chemicals Pvt. Ltd. (Mumbay, India). All the solvents were distilled prior to their use. Novozym 435 (immobilized Candida antarctica lipase) was purchased from Novo Nordisk A/S, Bagsvaerd, Denmark. Reactions were monitored by pre-coated TLC plates (Merck silica gel 60 F254, Darmstadt, Germany), by visualizing the spot in ceric solution stain and iodine. All the compounds were purified by column chromatography using silica gel (100–200 mesh).

3.2. Synthesis and Characterization

3.2.1. Synthesis of Oxybis(2-hydroxypropan-3,1-diyl) Diacetate (2)

In a 250 mL RB flask diglycerol (1, 30.1 mmol) was dissolved in THF (150 mL) followed by the addition of Novozym 435 (20 wt % of monomers). After stirring for 10 min vinyl acetate (69.3 mmol, 5.96 g) was added, the reaction mixture was placed in an incubator shaker at 230 rpm for 6 h at 30 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The organic solvent was evaporated under reduced pressure. The obtained crude product was purified by column chromatography using CHCl3–MeOH to give the desired compound 2 as a viscous liquid (75%); IR (neat) νmax: 3778, 3698, 2983, 1743, 1712 cm−1; 1H-NMR (CDCl3): δ 4.24–3.97 (m, 6H, H-3, H-4, H-7, H-8), 3.67–3.53 (4H, m, H-5, H-6), 3.22 (br s, 1H, OH), 3.00 (br s, 1H, OH), 2.08 (s, 6H, H-1, H-10) ppm; 13C-NMR (CDCl3): δ 171 (C-2, C-9), 72 (C-5, C-6), 68 (C-4, C-7), 65 (C-3, C-8), 20 (C-1, C-10) ppm; HRMS (positive, acetonitrile): m/z calcd. for C10H18O7: 250.1053; found [M + Na]+: 273.0946.

3.2.2. Synthesis of ((2-Hydroxypropane-1,3-diyl)bis(oxy))bis(2-hydroxypropane-3,1-diyl) Diacetate (10)

In a 250 mL RB flask triglycerol (9, 30.1 mmol) was dissolved in THF (150 mL) followed by the addition of Novozym 435 (20 wt % of monomers). After stirring for 10 min vinyl acetate (69.3 mmol, 5.96 g) was added, the reaction mixture was placed in an incubator shaker at 230 rpm for 6 h at 30 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The organic solvent was evaporated under reduced pressure. The obtained crude product was purified by column chromatography using CHCl3-MeOH to give the desired compound 10 as a viscous liquid (70%); IR (neat) νmax: 3399, 2880, 1724, 1694 1443 cm−1; 1H-NMR (CDCl3): δ 4.14–4.08 (m, 4H, H-3, H-11), 4.02–3.93 (m, 3H, H-4, H-7, H-10), 3.58–3.43 (m, 8H, H-5, H-6, H-8, H-9), 2.08 (s, 6H, H-1, H-13) ppm; 13C-NMR (CDCl3): δ 171 (C-2, C-12), 72 (C-5, C-6, C-8, C-9), 69, 68 (C-4, C-7, C-10), 65 (C-3, C-11), 21 (C-1, C-13) ppm; HRMS (positive, acetonitrile): m/z calcd. for C13H24O9: 324.1420; found [M + Na]+: 347.1909.

3.2.3. Synthesis of Diethyl 4,4′-((Oxybis(1-acetoxypropane-3,2-diyl))bis(oxy))dibenzoate (3)

To a stirred solution of compound 2 (4.0 mmol), ethyl,4-hydroxybenzoate (8.4 mmol, 1. 39 g) and triphenylphosphine (12.0 mmol, 3.15 g) in THF (20 mL), DIAD (10 mmol, 2.02 g) in THF (5 mL) was added dropwise. The reaction mixture was stirred for 15 h at 40 °C. The progress of the reaction was monitored by TLC (ethyl acetate–petroleum ether, 1:1, v/v). On completion of the reaction, the reaction mixture was concentrated under reduced pressure and the desired compound was extracted with ethyl acetate (3 × 30 mL). The combined organic layer was dried over anhydrous Na2SO4 and concentrated in vacuuo. The obtained crude product was purified through column chromatography using petroleum ether-ethyl acetate to give the desired compound 3 as a viscous liquid (70%); IR (neat) νmax: 2981, 1740, 1708, 1603, 1506 cm−1; 1H-NMR (CDCl3): δ 7.97–7.94 (m, 4H, H-3′, H-5′), 6.95–6.93 (m, 4H, H-2′, H-6′), 4.69–4.64 (m, 2H, H-4, H-7), 4.34–4.28 (m, 8H, H-3, H-8, H-8′), 3.77–3.70 (m, 4H, H-5, H-6), 2.02 (s, 6H, H-1, H-10), 1.36 (t, 6H, J = 8.0, H-9′) ppm; 13C-NMR (CDCl3): δ 170 (C-2, C-9), 166 (C-7′), 161 (C-1′), 131 (C-3′, C-5′), 123 (C-4′), 115 (C-2′, C-6′), 74 (C-4, C-7), 70 (C-5, C-6), 63 (C-3, C-8), 60 (C-8′), 20 (C-1, C-10), 14 (C-9′) ppm; HRMS (positive, acetonitrile): m/z calcd. for C28H34O11: 546.2101; found [M + H]+: 547.2162.

3.2.4. Synthesis of Diethyl 4,4′-((9-(4-(Ethoxycarbonyl)phenoxy)-2,16-dioxo-3,7,11,15-tetraoxahepta-decane-5,13-diyl)bis(oxy))dibenzoate (11)

To a stirred solution of compound 10 (4.0 mmol), ethyl,4-hydroxybenzoate (8.4 mmol, 1.39 g) and triphenylphosphine (12.0 mmol, 3.15 g) in THF (20 mL), DIAD (10 mmol, 2.02 g) in THF (5 mL) was added drop wise. The reaction mixture was stirred for 15 h at 40 °C. The progress of the reaction was monitored by TLC (ethyl acetate–petroleum ether, 1:1, v/v). On completion of the reaction, the reaction mixture was concentrated under reduced pressure and the desired compound was extracted with ethyl acetate (3 × 30 mL). The combined organic layer was dried over anhydrous Na2SO4 and concentrated in vacuuo. The obtained crude product was purified through column chromatography using petroleum ether–ethyl acetate to give the desired compound 11 as a viscous liquid (65%); IR (neat) νmax: 2992, 1741, 1707, 1603, 1507 cm−1; 1H-NMR (CDCl3): δ 7.97–7.90 (m, 6H, H-3′, H-5′), 6.95-6.85 (m, 6H, H-2′, H-6′), 4.67–4.54 (m, 3H, H-4, H-7, H-10), 4.36–4.31 (m, 6H, H-8′), 4.28–4.25 (m, 4H, H-3, H-11), 3.76–3.66 (m, 8H, H-5, H-6, H-8, H-9), 2.06, 2.02 (s, 6H, H-1, H-13), 1.39–1.35 (m, 9H, H-9′) ppm; 13C-NMR (CDCl3): δ 170 (C-2, C-12), 166 (C-7′), 161 (C-1′), 131 (C-3′, C-5′), 123 (C-4′), 115 (C-2′, C-6′), 76, 74 (C-4, C-7, C-10), 70 (C-5, C-6, C-8, C-9), 63 (C-3, C-11), 60 (C-8′), 20 (C-1, C-13), 14 (C-9′) ppm; HRMS (positive, acetonitrile): m/z calcd for C40H48O15: 768.2993; found [M + NH4]+: 786.3316.

3.2.5. Synthesis of Oxybis(2-azidopropane-3,1-diyl) Diacetate (5)

A solution of compound 2 (4.0 mmol) in DCM (30 mL) was cooled under nitrogen atmosphere over ice bath, triethylamine (16.0 mmol, 1.62 g) and methanesulfonyl chloride (12 mmol, 1.37 g) were then added with maintaining the temperature of the reaction mixture at 0 °C. The solution was then stirred at 30 °C for 2.5 h, the progress of the reaction was monitored by TLC (methanol–chloroform, 1:19, v/v). On completion of the reaction, the salt was filtered and the solvent evaporated under reduced pressure. To the mesylated product (4 mmol) so obtained, sodium azide (24 mmol, 1.56 g) and DMF (30 mL) were added, the reaction mixture was heated at 90 °C for 15 h. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:19, v/v). On completion of the reaction, DMF was removed under reduced pressure, the residue obtained was extracted with ethyl acetate (3 × 30 mL). The combined organic layer was dried over anhydrous sodium sulphate followed by evaporation of solvent. The crude product was purified by column chromatography using petroleum ether-ethyl acetate to give the desired compound 5 as a viscous liquid (75%); IR (neat) νmax: 2922, 2875, 2094, 1739, 1449 cm−1; 1H-NMR (CDCl3): δ 4.24–4.06 (m, 4H, H-3, H-8), 3.78-3.67 (m, 6H, H-4, H-5, H-6, H-7), 2.09, 2.08 (s, 6H, H-1, H-10) ppm; 13C-NMR (CDCl3): δ 170 (C-2, C-9), 71 (C-5, C-6), 63 (C-3, C-8), 59 (C-4, C-7), 20 (C-1, C-10) ppm; HRMS (positive, acetonitrile): m/z calcd. for C10H16N6O5: 300.1182; found [M + H]+: 301.1409.

3.2.6. Synthesis of ((2-Azidopropane-1,3-diyl)bis(oxy))bis(2-azidopropane-3,1-diyl) Diacetate (13)

A solution of compound 10 (4.0 mmol) in DCM (30 mL) was cooled under nitrogen atmosphere over ice bath, triethylamine (16.0 mmol, 1.62 g) and methanesulfonyl chloride (12 mmol, 1.37 g) were then added with maintaining the temperature of the reaction mixture at 0 °C. The solution was then stirred at 30 °C for 2.5 h, the progress of the reaction was monitored by TLC (methanol–chloroform, 1:19, v/v). On completion of the reaction, the salt was filtered and the solvent evaporated under reduced pressure. To the mesylated product (4 mmol) so obtained, sodium azide (24 mmol, 1.56 g) and DMF (30 mL) were added, the reaction mixture was heated at 90 °C for 15 h. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:19, v/v). On completion of the reaction, DMF was removed under reduced pressure, the residue obtained was extracted with ethyl acetate (3 × 30 mL). The combined organic layer was dried over anhydrous sodium sulphate followed by evaporation of solvent. The crude product was purified by column chromatography using petroleum ether-ethyl acetate to give the desired compound 13 as a viscous liquid (70 %); IR (neat) νmax: 2919, 2101, 1737, 1450 cm−1; 1H-NMR (CDCl3): δ 4.24–4.09 (m, 4H, H-3, H-11), 3.82–3.49 (m, 11H, H-4, H-5, H-6, H-7, H-8, H-9, H-10), 2.08 (s, 6H, H-1, H-13) ppm; 13C-NMR (CDCl3): δ 170 (C-1, C-13), 71 (C-5, C-6, C-8, C-9,), 70 (C-3, C-11), 63 (C-7), 59 (C-4, C-10), 20 (C-1, C-13) ppm; HRMS (positive, acetonitrile): m/z calcd. for C10H18O7: 399.1615; found [M + NH4]+: 417.1971.

3.2.7. Synthesis of Oxybis(1-acetoxypropane-3,2-diyl) bis(4-(prop-2-yn-1-yloxy)benzoate) (7)

Compound 2 (4.0 mmol) and 4-(prop-2-yn-1-yloxy)benzoic acid (8.4 mmol) were dissolved in a mixture of DCM and DMF in 4:1 ratio (30 mL). The reaction mixture was stirred at 0 °C, then EDC.HCl (10 mmol, 1.91 g) and DMAP (6 mmol, 0.73 g) were added, the reaction mixture was stirred for 15 h at 40 °C. The progress of the reaction was monitored by TLC (petroleum ether–ethyl acetate, 1:1, v/v). On completion of the reaction, the solvent was removed under reduced pressure and the residue obtained was extracted with ethyl acetate (3 × 30 mL), the organic layer was dried over anhydrous Na2SO4 and solvent evaporated under reduced pressure. The resulting crude product was purified by column chromatography using petroleum ether–zethyl acetate to give the desired compound 7 as a viscous liquid (75%); IR (neat) νmax: 3284, 2932, 2878, 2123, 1712, 1603, 1508 cm−1; 1H-NMR (CDCl3): δ 7.98–7.96 (m, 4H, H-3′, H-7′), 6.98–6.95 (m, 4H, H-4′, H-6′), 5.44–5.38 (m, 2H, H-4, H-7), 4.75–4.74 (m, 4H, H-8′), 4.41–4.29 (m, 4H, H-3, H-8), 3.82–3.70 (m, 4H, H-5, H-6), 2.56 (m, H-10′), 2.03 (s, 6H, H-1, H-10) ppm; 13C-NMR (CDCl3): δ 171 (C-2, C-9), 165 (C-1′), 161 (C-5′), 132 (C-3′, C-7′), 122 (C-2′), 114 (C-4′, C-6′), 77 (C-9′), 76 (C-10′), 70 (C-4, C-7), 69 (C-5, C-6), 62 (C-3, C-8), 55 (C-8′), 21 (C-1, C-10) ppm; HRMS (positive, acetonitrile): m/z calcd. for C30H30O11: 566.1788; found [M + NH4]+: 584.2123.

3.2.8. Synthesis of 2,16-Dioxo-9-((4-(prop-2-yn-1-yloxy)benzoyl)oxy)-3,7,11,15-tetraoxaheptadecane-5,13-diyl bis(4-(prop-2-yn-1-yloxy)benzoate) (15)

Compound 10 (4.0 mmol) and 4-(prop-2-yn-1-yloxy)benzoic acid (8.4 mmol) were dissolved in a mixture of DCM and DMF in 4:1 ratio (30 mL). The reaction mixture was stirred at 0 °C, then EDC.HCl (10 mmol, 1.91 g) and DMAP (6 mmol, 0.73 g) were added, the reaction mixture was stirred for 15 h at 40 °C. The progress of the reaction was monitored by TLC (petroleum ether–ethyl acetate, 1:1, v/v). On completion of the reaction, the solvent was removed under reduced pressure and the residue obtained was extracted with ethyl acetate (3 × 30 mL), the organic layer was dried over anhydrous Na2SO4 and solvent evaporated under reduced pressure. The resulting crude product was purified by column chromatography using petroleum ether–ethyl acetate to give the desired compound 15 as a viscous liquid (70%); IR (neat) νmax: 3286, 2923, 2122, 1714, 1604, 1507 cm−1; 1H-NMR (CDCl3): δ 8.18–7.95 (m, 6H, H-3′, H-7′), 7.26–6.94 (m, 6H, H-4′, H-6′), 5.39–5.24 (m, 3H, H-4, H-7, H-10), 4.80–4.73 (m, 6H, H-8′), 4.37–4.27 (m, 4H, H-3, H-11), 3.90–3.62 (m, 8H, H-5, H-6, H-8, H-9), 2.58–2.50 (m, 3H, H-10′), 2.03 (s, 6H, H-1, H-13) ppm; 13C-NMR (CDCl3): δ 170 (C-2, C-12), 165 (C-1′), 161 (C-5′), 131 (C-3′, C-7′), 123 (C-2′), 114 (C-4′, C-6′), 77 (C-9′), 76 (C-10′), 70 (C-4, C-7, C-10), 70 (C-5, C-6, C-8, C-9), 62 (C-3, C-11), 55 (C-8′) 20 (C-1, C-13) ppm; HRMS (positive, acetonitrile): m/z calcd. for C43H42O15: 798.2524; found [M + NH4]+: 816.2855.

3.2.9. Synthesis of 4,4′-((Oxybis(1-hydroxypropane-3,2-diyl))bis(oxy))dibenzoate (4)

Compound 3 (0.91 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The filtrate was concentrated under reduced pressure. The crude product so obtained was purified by column chromatography using CHCl3–MeOH to give the desired compound 4 as a viscous liquid (75%); IR (neat) νmax: 3440, 2935, 17033, 1603, 1507 cm−1; 1H-NMR (CDCl3): δ 7.95 (d, 4H, J = 8.0, H-3′, H-5′), 6.93 (d, 4H, J = 8.0, H-2′, H-6′), 4.56–4.54 (m, 2H, H-2, H-5), 4.36–4.30 (m, 4H, H-8′), 4.00–4.72 (m, 8H, H-1, H-3, H-4, H-6), 2.39 (br s, 2H, OH), 1.37 (t, 6H, H-9′) ppm; 13C-NMR (CDCl3): δ 166 (C-7′), 161 (C-1′), 131 (C-3′, C-5′), 123 (C-4′), 115 (C-2′, C-6′), 77 (C-2, C-5), 70 (C-3, C-4), 62 (C-1, C-6), 61 (C-8′), 14 (C-9′) ppm; HRMS (positive, acetonitrile): m/z calcd. for C24H30O9: 462.1890; found [M + H]+: 463.1964.

3.2.10. Synthesis of Diethyl 4,4′-((((2-(4-(ethoxycarbonyl)phenoxy)propane-1,3-diyl)bis(oxy))bis(1-hydroxypropane-3,2-diyl))bis(oxy))dibenzoate (12)

Compound 11 (0.91 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The filtrate was concentrated under reduced pressure. The crude product so obtained was purified by column chromatography using CHCl3–MeOH to give the desired compound 12 as a viscous liquid (65%); IR (neat) νmax: 3465, 2980, 2928, 1702, 1603, 1507 cm−1; 1H-NMR (CDCl3, 25 °C): δ 7.97–7.94 (m, 6H, H-3′, H-5′), 6.98–6.89 (m, 6H, H-2′, H-6′), 4.58–4.53 (m, 3H, H-2, H-5, H-8), 4.37–4.31 (m, 6H, H-8′), 3.92–3.70 (m, 12H, H-1, H-3, H-4, H-6, H-7, H-9), 2.27 (br s, 2H, OH), 1.39 (m, 9H, H-9′) ppm; 13C-NMR (CDCl3): δ 166 (C-7′), 161 (C-1′), 131 (C-3′, C-5′), 124 (C-4′), 115 (C-2′, C-6′), 75, 70 (C-2, C-5, C-8), 62 (C-3, C-4, C-6, C-7), 62 (C-1, C-9), 60 (C-8′), 14 (C-9′) ppm; HRMS (positive, acetonitrile): m/z calcd. for C36H44O13: 684.2773; found [M + NH4]+: 702.3113.

3.2.11. Synthesis of 3,3′-Oxybis(2-azidopropan-1-ol) (6)

Compound 5 (1.6 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The filtrate was concentrated under reduced pressure. The crude product obtained was purified by column chromatography using CHCl3–MeOH as an eluent to give the desired compound 6 as a viscous liquid (70%); IR (neat) νmax: 3357, 2926, 2978, 2088, 1638, 1464 cm−1; 1H-NMR (CDCl3): δ 3.78–3.65 (m, 10H, H-1, H-2, H-3, H-4, H-5, H-6), 2.45 (br s, 2H, OH) ppm; 13C-NMR (CDCl3): δ 71 (C-3, C-4), 62 (C-2, C-5), 62 (C-1, C-6) ppm. HRMS (positive, acetonitrile): m/z calcd. for C6H12N6O3: 216.0971; found [M + H]+: 217.1044.

3.2.12. Synthesis of 3,3′-((2-Azidopropane-1,3-diyl)bis(oxy))bis(2-azidopropan-1-ol) (14)

Compound 13 (1.6 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered off and washed with methanol. The filtrate was concentrated under reduced pressure. The crude product obtained was purified by column chromatography using CHCl3–MeOH as an eluent to give the desired compound 14 as a viscous liquid (60%); IR (neat) νmax: 3390, 2915, 1988, 2105, 1748, 1480 cm−1; 1H-NMR (CDCl3): δ 3.78–3.55 (m, 15H, H-1–H-9) ppm; 13C-NMR (CDCl3, 25 °C): δ 71 (C-4, C-6), 70 (C-3, C-7), 62 (C-1, C-9), 62, 60 (C-2, C-5, C-8) ppm; HRMS (positive, acetonitrile): m/z calcd. for C9H17N9O4: 315.14.04; found [M + H]+: 338.1298.

3.2.13. Synthesis of Oxybis(1-hydroxypropane-3,2-diyl) bis(4-(prop-2-yn-1-yloxy)benzoate) (8)

Compound 7 (0.88 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered and washed with methanol. The filtrate was concentrated under reduced pressure. The obtained crude product was purified by column chromatography using CHCl3–MeOH to give the desired compound 8 as a viscous liquid (70%); IR (neat) νmax: 3459, 3289, 2932, 2880, 2123, 1701, 1603, 1508 cm−1; 1H-NMR (CDCl3): δ 8.01–7.98 (m, 4H, H-3′, H-7′), 6.98–6.95 (m, 4H, H-4′, H-6′), 5.26–5.22 (m, 2H, H-2, H-5), 4.74 (m, 4H, H-8′), 3.92–3.73 (m, 8H, H-1, H-3, H-4, H-6), 2.56 (m, 2H, H-10′), 2.43 (br s, 2H, OH) ppm; 13C-NMR (CDCl3): δ 166 (C-1′), 161 (C-5′), 132 (C-3′, C-7′), 123 (C-2′), 114 (C-4′, C-6′), 77 (C-9′), 76 (C-10′), 73 (C-2, C-5), 70 (C-3, C-4), 62 (C-1, C-6), 55 (C-8′) ppm; HRMS (positive, acetonitrile): m/z calcd. for C26H26O9: 482.1577; found [M + Na]+: 505.1459.

3.2.14. Synthesis of ((2-((4-(Prop-2-yn-1-yloxy)benzoyl)oxy)propane-1,3-diyl)bis(oxy))bis(1-hydroxypropane-3,2-diyl) bis(4-(Prop-2-yn-1-yloxy)benzoate) (16)

Compound 15 (0.88 mmol) was dissolved in THF (1 mL), then n-butanol (1 mL) and Novozym 435 (50 wt % of monomer) were added. The reaction mixture was kept in an incubator shaker at 230 rpm for 72 h at 60 °C. The progress of the reaction was monitored by TLC (methanol–chloroform, 1:9, v/v). On completion of the reaction, the enzyme was filtered and washed with methanol. The filtrate was concentrated under reduced pressure. The obtained crude product was purified by column chromatography using CHCl3–MeOH to give the desired compound 16 as a viscous liquid (65%); IR (neat) νmax: 3286, 2955, 2878, 2123, 1713, 1603, 1508 cm−1; 1H-NMR (CDCl3): δ 8.17–7.94 (m, 6H, H-3′, H-7′), 7.08–6.94 (m, 6H, H-4′, H-6′), 5.39–5.18 (m, 3H, H-2, H-5, H-8), 4.79–4.73 (m, 6H, H-8′), 3.96–3.67 (m, 12H, H-1, H-3, H-4, H-6, H-7, H-9), 2.66 (br s, 2H, OH), 2.59–2.55 (m, 3H, H-10′) ppm; 13C-NMR (CDCl3): δ 165 (C-1′), 161 (C-5′), 131 (C-3′, C-7′), 123 (C-2′), 114 (C-4′, C-6′), 77 (C-9′), 76 (C-10′), 73 (C-2, C-5, C-8), 70 (C-3, C-4, C-6, C-7), 62 (C-1, C-9), 55 (C-8′) ppm; HRMS (positive, acetonitrile): m/z calcd. for C39H38O13: 714.2303; found [M + Na]+: 737.2198.

4. Conclusions

A series of 13 novel glycerol oligomer derivatives 38 and 1016 have been synthesized using chemo-enzymatic approach that can be used further for the synthesis of supramolecular architectures. The enzyme catalyzed approach exhibits high chemo- and regioselectivity as the primary hydroxyl group can be selectively acylated in the presence of secondary hydroxyl group. The acetyl groups of the hydroxymethyl of polymers 2 and 10 were stable during the acylation, etherification and azidation. Moreover selective enzymatic deprotection of the primary hydroxyl groups was efficient without saponification and trans-esterification of the ester in the secondary hydroxyl group.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/21/8/1038/s1.

Acknowledgments

The authors acknowledge the Science & Engineering Research Board (DST-SERB), Government of India for financial support.

Author Contributions

All authors contributed equally to this work. Rainer Haag, Sunil K. Sharma and Christophe Len conceived and designed the experiments; Abhishek K. Singh, Remi Nguyen and Nicolas Galy performed the experiments; Abhishek K. Singh, Remi Nguyen and Nicolas Galy analyzed the data; Rainer Haag, Sunil K. Sharma and Christophe Len contributed reagents/materials/analysis tools; Rainer Haag, Sunil K. Sharma and Christophe Len wrote the paper.”

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
DMF
N,N-dimethylformamide
EDC
1-ethyl-3-(3-dimethylaminopropyl)carbodiimide
FT-IR
Fourier transform infrared
HRMS
high-resolution mass spectrometry
IR
infrared
MeOH
methanol
NMR
nuclear magnetic resonance
pKa
acid dissociation constant
ppm
parts per million
PVA
polyvinyl acetate
Q-TOF LCMS
quadrupoletime of flight liquid chromatography mass spectrometry
RB
round-bottom
rpm
round per min
SAR
structure-activity relationship
THF
tetrahydrofuran
TLC
thin layer chromatography
TMS
tetramethylsilane

References

  1. Katryniok, B.; Kimura, H.; Skrzynska, E.; Girardon, J.; Fongarland, P.; Capron, M.; Ducoulombier, R.; Mimura, N.; Paul, S.; Dumeignil, F. Selective catalytic oxidation of glycerol: perspective for high value chemicals. Green Chem. 2011, 13, 1960–1979. [Google Scholar] [CrossRef]
  2. Cintas, P.; Tagliapietra, S.; Gaudino, E.C.; Palmisano, G.; Cravotto, G. Glycerol: A solvent and a building block of choice for microwave and ultrasound irradiation procedure. Green Chem. 2014, 16, 1056–1065. [Google Scholar] [CrossRef]
  3. Behr, A.; Eilting, J.; Irawadi, K.; Leschinski, J.; Lindner, F. Improved utilization of renewable ressources: New important derivatives of glycerol. Green Chem. 2008, 10, 13–30. [Google Scholar] [CrossRef]
  4. García, J.; Marín, H.; Pires, E. Glycerol based solvents: Synthesis, properties and applications. Green Chem. 2014, 16, 1007–1033. [Google Scholar] [CrossRef]
  5. Diaz-Alvarez, A.E.; Francos, J.; Lastra-Barreira, B.; Crochet, P.; Cadierno, V. Glycerol and derived solvents: New sustainable reaction media for organic synthesis. Chem. Commun. 2011, 47, 6208–6227. [Google Scholar] [CrossRef] [PubMed]
  6. Len, C.; Merlot, A.S.; Postel, D.; Ronco, G.; Villa, P.; Goubert, C.; Jeufrault, E.; Mathon, B.; Simon, H. Synthesis and antifungal activity of novel bis (dithiocarbamate) derivatives of glycerol. J. Agric. Food Chem. 1996, 44, 2856–2858. [Google Scholar] [CrossRef]
  7. Len, C.; Postel, D.; Ronco, G.; Villa, P.; Goubert, C.; Jeufrault, E.; Mathon, B.; Simon, H. Synthesis of carbamic esters derivatives of itols: Antifungal activity against various crop diseases. J. Agric. Food Chem. 1997, 45, 3–6. [Google Scholar] [CrossRef]
  8. Rafin, C.; Veignie, E.; Sancholle, M.; Postel, D.; Len, C.; Villa, P.; Ronco, G. Synthesis and antifungal activity of novel bisdithiocarbamate derivatives of carbohydrates against Fusarium oxysporum f. sp. Lini. J. Agric. Food Chem. 2000, 48, 5283–5287. [Google Scholar] [CrossRef] [PubMed]
  9. DeSousa, R.; Thurier, C.; Len, C.; Pouilloux, Y.; Barrault, J.; Jerome, F. Regioselective functionalization of glycerol with a dithiocarbamate moiety: An environmentally friendly route to safer fungicides. Green Chem. 2011, 13, 1129–1132. [Google Scholar] [CrossRef]
  10. Saggadi, H.; Luart, D.; Thiebault, N.; Polaert, I.; Estel, L.; Len, C. Toward the synthesis of 6-hydroxyquinoline starting from glycerol via improved microwave-assisted modified Skraup reaction. Catal. Commun. 2014, 44, 15–18. [Google Scholar] [CrossRef]
  11. Len, C.; Luque, R. Continuous flow transformations of glycerol to valuable products: An overview. Sustain. Chem. Process. 2014, 2. [Google Scholar] [CrossRef]
  12. Saggadi, H.; Luart, D.; Thiebault, N.; Polaert, I.; Estel, L.; Len, C. Quinoline and phenanthroline preparation starting from glycerol via improved microwave-assisted modified Skraup reaction. RSC Adv. 2014, 4, 21456–21464. [Google Scholar] [CrossRef]
  13. Saggadi, H.; Polaert, I.; Luart, D.; Len, C.; Estel, L. Microwaves under pressure for the continuous production of quinolone from glycerol. Catal. Today 2015, 255, 66–74. [Google Scholar] [CrossRef]
  14. Zhou, C.H.; Beltramini, J.N.; Fana, Y.X.; Lu, G.Q. Chemoselective catalytic conversion of glycerol as a biorenewable source to valuable commodity chemicals. Chem. Soc. Rev. 2008, 37, 527–549. [Google Scholar] [CrossRef] [PubMed]
  15. Martin, A.; Richter, M. Oligomerization of glycerol: A critical review. Eur. J. Lipid Sci. Technol. 2011, 113, 100–117. [Google Scholar] [CrossRef]
  16. Shi, Y.; Dayoub, W.; Chen, G.R.; Lemaire, M. Selective synthesis of 1-O-alkyl glycerol and diglycerol ethers by reductive alkylation of alcohols. Green Chem. 2010, 12, 2189–2195. [Google Scholar] [CrossRef]
  17. Cassel, S.; Debaig, C.; Benvegnu, T.; Chaimbault, P.; Lafosse, M.; Plusquellec, D.; Rollin, P. Original synthesis of linear, branched and cyclic oligoglycerol standards. Eur. J. Org. Chem. 2001, 875–896. [Google Scholar] [CrossRef]
  18. Sutter, M.; Dayoub, W.; Métay, E.; Raoul, Y.; Lemaire, M. 1-O-Alkyl (di)glycerol ethers synthesis from methyl esters and triglycerides by two pathways: catalytic reductive alkylation and transesterification/reduction. Green Chem. 2013, 15, 786–797. [Google Scholar] [CrossRef]
  19. Katryniok, B.; Paul, S.; Baca, V.; Reye, P. Glycerol dehydration to acrolein in the context of new uses of glycerol. Green Chem. 2010, 12, 2079–2098. [Google Scholar] [CrossRef]
  20. Urata, K. Unique block molecules based on glycerol skeleton as C-3 building blocks for liquid crystals formation by self-assembly and their future potential for the “nano-chemistry”. Eur. J. Lipid Sci. Technol. 2003, 105, 542–556. [Google Scholar] [CrossRef]
  21. Sonnati, M.O.; Amigoni, S.; Taffin de Givenchy, E.P.; Darmanin, T.; Chouletb, O.; Guittard, F. Glycerol carbonate as a versatile building block for tomorrow: Synthesis, reactivity, properties and applications. Green Chem. 2013, 15, 283–306. [Google Scholar] [CrossRef]
  22. Kirk, O.; Christensen, M.W. Lipases from Candida antartica: Unique biocatalysts from a unique origin. Org. Proc. Res. Dev. 2002, 6, 446–451. [Google Scholar] [CrossRef]
  23. Anderson, E.M.; Larsson, K.M.; Krik, O. One biocatalyst-many applications: The use of Candida antartica B-lipase in organic synthesis. Biocatal. Biotransform. 1998, 16, 181–204. [Google Scholar] [CrossRef]
  24. Krik, O.; Bjorkling, F.; Godtfredsen, S.E.; Larsen, T.O. Fatty acid specificity in lipase-catalyzed synthesis of glucoside esters. Biocatalysis 1992, 6, 127–134. [Google Scholar] [CrossRef]
  25. Gupta, S.; Jalal, S.; Kumar, S.; Haag, R.; Sharma, S.K. A simple and convenient chemoenzymatic approach for the synthesis of valuable triacylglycerol-based dendritic building blocks. Indian J. Chem. 2012, 51B, 1376–1387. [Google Scholar]
  26. Gupta, S.; Schade, B.; Kumar, S.; Böttcher, C.; Sharma, S.K.; Haag, R. Dendronized multiamphiphilic polymers as non-ionic nanotransporters for biomedical applications: Synthesis, aggregation behavior and transport properties. Small 2013, 9, 894–904. [Google Scholar] [CrossRef] [PubMed]
  27. Kumari, M.; Gupta, S.; Böttcher, C.; Khandare, J.; Sharma, S.K.; Haag, R. Dendronized multifunctional amphiphilic polymers as efficient nanocarriers for biomedical applications. Macromol. Rapid Commun. 2015, 36, 254–251. [Google Scholar] [CrossRef] [PubMed]
  28. Kumari, M.; Billamboz, M.; Leonard, E.; Len, C.; Böttcher, C.; Prasad, A.K.; Haag, R.; Sharma, S.K. Self-assembly, photoresponsive behavior and transport potential of azobenzene grafted dendronized polymeric amphiphiles. RSC Adv. 2015, 5, 48301–48310. [Google Scholar] [CrossRef]
  29. Kumar, S.; Achazi, K.; Böttcher, C.; Licha, K.; Haag, R.; Sharma, S.K. Encapsulation and cellular internalization of cyanine dye using amphiphilic dendronized polymers. Eur. Polym. J. 2015, 69, 416–428. [Google Scholar] [CrossRef]
  30. Thirumurugan, P.; Matosiuk, D.; Jozwiak, K. Click chemistry for drug development and diverse chemical-biology applications. Chem. Rev. 2013, 113, 4905–4979. [Google Scholar] [CrossRef] [PubMed]
  31. Legros, V.; Vanhaverbecke, C.; Souard, F.; Len, C.; Desire, J. β-Cyclodextrin-glycerol dimers: Synthesis and NMR conformational analysis. Eur. J. Org. Chem. 2013, 2013, 2583–2590. [Google Scholar] [CrossRef]
  32. Espeel, P.; du Prez, F.E. “Click”-inspired chemistry in macromolecular science: Matching recent progress and user expectations. Macromolecules 2015, 48, 2–14. [Google Scholar] [CrossRef]
  33. Saku, O.; Ishida, H.; Atsumi, E.; Sugimoto, Y.; Kodaira, H.; Kato, Y.; Shirakura, S.; Nakasato, Y. Discovery of novel 5,5-diarylpentadienamides as orally available transient receptor potential vanilloid 1 (TRPV1) antagonists. J. Med. Chem. 2012, 55, 3436–3451. [Google Scholar] [CrossRef] [PubMed]
  34. Zhu, Y.Y.; Yin, T.T.; Li, X.L.; Su, M.; Xue, Y.X.; Yu, Z.P.; Liu, N.; Yin, J.; Wu, Z.Q. Chiroptical and thermotropic properties of helical styrenic polymers: effect of achiral group. Macromolecules 2014, 47, 7021–7029. [Google Scholar] [CrossRef]
  35. Sharma, D.; Khandelwal, A.; Sharma, R.K.; Bhatia, S.; Reddy, L.C.; Olsen, C.E.; Wengel, J.; Parmar, V.S.; Prasad, A.K. Synthesis and regioselective deacylation studies on peracylated 2′-azido arabino- and ribo-thymine nucleosides: towards 5′-O,2′-N-linked oligonucleotides. Indian J. Chem. 2009, 48B, 1712–1720. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds are not available from the authors.
Scheme 1. Synthesis of diglycerol/triglycerol based building blocks. Reagents and Conditions: (i) Novozym 435, vinyl acetate, THF, 6 h, 30 °C; (ii) DIAD, TPP, THF, ethyl,4-hydroxybenzoate, 15 h, 40 °C (iii) (a) MsCl, DCM, TEA, 2.5 h, 0 to −5 °C. (b) DMF, NaN3, 90 °C, 15 h. (iv) EDC.HCl, DCM:DMF (4:1), 4-(prop-2-yn-1-yloxy)benzoic acid, 15 h, 30 °C. (v) Novozym 435, n-butanol, THF, 72 h, 60 °C.
Scheme 1. Synthesis of diglycerol/triglycerol based building blocks. Reagents and Conditions: (i) Novozym 435, vinyl acetate, THF, 6 h, 30 °C; (ii) DIAD, TPP, THF, ethyl,4-hydroxybenzoate, 15 h, 40 °C (iii) (a) MsCl, DCM, TEA, 2.5 h, 0 to −5 °C. (b) DMF, NaN3, 90 °C, 15 h. (iv) EDC.HCl, DCM:DMF (4:1), 4-(prop-2-yn-1-yloxy)benzoic acid, 15 h, 30 °C. (v) Novozym 435, n-butanol, THF, 72 h, 60 °C.
Molecules 21 01038 sch001
Figure 1. 1H-NMR spectrum of compound 2, 4, and 8.
Figure 1. 1H-NMR spectrum of compound 2, 4, and 8.
Molecules 21 01038 g001
Figure 2. 1H-NMR spectrum of compound 10, 12 and 16.
Figure 2. 1H-NMR spectrum of compound 10, 12 and 16.
Molecules 21 01038 g002

Share and Cite

MDPI and ACS Style

Singh, A.K.; Nguyen, R.; Galy, N.; Haag, R.; Sharma, S.K.; Len, C. Chemo-Enzymatic Synthesis of Oligoglycerol Derivatives. Molecules 2016, 21, 1038. https://doi.org/10.3390/molecules21081038

AMA Style

Singh AK, Nguyen R, Galy N, Haag R, Sharma SK, Len C. Chemo-Enzymatic Synthesis of Oligoglycerol Derivatives. Molecules. 2016; 21(8):1038. https://doi.org/10.3390/molecules21081038

Chicago/Turabian Style

Singh, Abhishek K., Remi Nguyen, Nicolas Galy, Rainer Haag, Sunil K. Sharma, and Christophe Len. 2016. "Chemo-Enzymatic Synthesis of Oligoglycerol Derivatives" Molecules 21, no. 8: 1038. https://doi.org/10.3390/molecules21081038

Article Metrics

Back to TopTop