Next Article in Journal
RutheniumII(η6-arene) Complexes of Thiourea Derivatives: Synthesis, Characterization and Urease Inhibition
Previous Article in Journal
A Class of Promising Acaricidal Tetrahydroisoquinoline Derivatives: Synthesis, Biological Evaluation and Structure-Activity Relationships
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bioresolution Production of (2R,3S)-Ethyl-3-phenylglycidate for Chemoenzymatic Synthesis of the Taxol C-13 Side Chain by Galactomyces geotrichum ZJUTZQ200, a New Epoxide-Hydrolase-Producing Strain

College of Biological and Environmental Engineering, Zhejiang University of Technology, Hangzhou 310014, China
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(6), 8067-8079; https://doi.org/10.3390/molecules19068067
Submission received: 3 May 2014 / Revised: 6 June 2014 / Accepted: 11 June 2014 / Published: 16 June 2014
(This article belongs to the Section Natural Products Chemistry)

Abstract

:
A newly isolated Galactomyces geotrichum ZJUTZQ200 strain containing an epoxide hydrolase was used to resolve racemic ethyl 3-phenylglycidate (rac-EPG) for producing (2R,3S)-ethyl-3-phenylglycidate ((2R,3S)-EPG). G. geotrichum ZJUTZQ200 was verified to be able to afford high enantioselectivity in whole cell catalyzed synthesis of this chiral phenylglycidate synthon. After the optimization of the enzymatic production and bioresolution conditions, (2R,3S)-EPG was afforded with high enantioselectivity (e.e.S > 99%, E > 49) after a 8 h reaction. The co-solvents, pH buffer solutions and substrate/cell ratio were found to have significant influences on the bioresolution properties of G. geotrichum ZJUTZQ200. Based on the bioresolution product (2R,3S)-EPG, taxol’s side chain ethyl (2R,3S)-3-benzoylamino-2-hydroxy-3-phenylpropionate was successfully synthesized by a chemoenzymatic route with high enantioselectivity (e.e.S > 95%).

1. Introduction

Optically pure EPG, an important pharmaceutical intermediate, has a wide variety of applications. (2R,3S)-EPG is a key intermediate in the synthesis of taxol’s C-13 side chain ethyl (2R,3S)-3-benzoylamino-2-hydroxy-3-phenylpropionate [1]. It is also used in the synthesis of the drug reboxetine which is an inhibitor of the reuptake of norepinephrine [2]. In addition, (2S,3R)-EPG can be applied to the synthesis of the nootropic drug clausenamide [3]. Analogues of EPG also have a wide range of applications in drug synthesis. (2R,3S)-3-(4-methylphenyl)glycidic acid methyl ester is an important intermediate in the synthesis of the cardiovascular drug diltiazem [4]. Moreoever, (2S,3R)-benzylepoxy acid esters are important intermediates in the synthesis of a series of anti-cancer drugs that act as aminopeptidase N inhibitors, such as bestatin, phebestin and probestin, etc. [5].
The preparation of chiral phenylglycidate synthons is mainly achieved by chemical synthesis or enzymatic asymmetric hydrolysis [6]. Among the enzymatic methods, the use of lipase has been extensively reported. Mithilesh et al. obtained (2R,3S)-EPG with e.e.S > 99% using Lecitase Ultra immobilized by gelatin as biocatalyst [7]. Zheng et al. prepared (2R,3S)-3-phenylglycidic acid methyl ester by hydrolyzing rac-3-phenylglycidic acid methyl ester using Pseudomonas putida cells as biocatalyst [8]. After the optimization of fermentation and catalytic conditions, the enantiomeric excess of the final product was greater than 99% with a yield of 48%. Manoucheh et al. obtained a series of chiral epoxy esters including chiral EPG through selective hydrolysis of the corresponding racemic compounds catalyzed by lipase from pig liver [9].
However, there are also several drawbacks in lipase-catalyzed asymmetric hydrolysis for producing chiral phenylglycidates, which may hinder its industrial application. For example, the unstable epoxy acid hydrolysis product may decarboxylate rapidly to the corresponding aldehyde [4], which affects the stability of the reaction system, and even leads to irreversible degenerative inactivation of lipase.
Another enzymatic option is epoxide hydrolase (EH) [10], which has none of the above drawbacks in the production of chiral phenylglycidates. In this field, EH-catalyzed hydrolysis of rac-EPG has been reported using the whole cells of Pseudomonas sp., yielding (2R,3S)-EPG with 95% e.e.S and 26% yield in 12 h from 0.2% (w/v) of the racemate [11]. Since there are few reports on the preparation of chiral EPG by EH [11,12], the exploration of new microbial catalysts and the improvement of EH enantioselectivity are thus necessary.
As a blockbuster anticancer drug, taxol represents one of the most clinically valuable natural products known to mankind in the recent past [1,13]. The dominant method for the preparation of taxol was the semisynthesis route by attaching a synthetic optically active C-13 side chain to natural baccatin III derived from renewable yew leaves. Chemoenzymatic synthesis protocols have been recognized as a very useful means to prepare optically active compounds [14,15]. Lipase-catalyzed asymmetric hydrolysis routes have been widely reported in the chemoenzymatic synthesis of taxol’s C-13 side chain [1]. In this work, we attempted to discover a new EH-producing microorganism that can be used to prepare chiral (2R,3S)-EPG as a synthon for taxol’s C-13 side chain. As a result, we obtained a newly isolated microorganism Galactomyces geotrichum ZJUTZQ200 to produce (2R,3S)-EPG with excellent enantioselectivity by whole-cell catalysis (Scheme 1). Furthermore, (2R,3S)-EPG prepared by EH was applied for the first time to the synthesis of the taxol C-13 side chain.
Scheme 1. Enantioselective hydrolysis of rac-EPG by G. geotrichum ZJUTZQ200.
Scheme 1. Enantioselective hydrolysis of rac-EPG by G. geotrichum ZJUTZQ200.
Molecules 19 08067 g004

2. Results and Discussion

2.1. Screening and Identification of a Strain Producing Enantioselective EH

We have successfully isolated more than 200 strains from soil samples using 3-phenylglycidol, an analogue of rac-EPG, as the sole carbon source. The selective hydrolysis capabilities of rac-EPG were checked by HPLC. More than 30 strains showed EH activity. Among them, ten strains could preferentially hydrolyze (2S,3R)-EPG with high enantioselectivity leading to (2R,3S)-EPG configuration retention. One of the fungi strains, ZJUTZQ200, afforded the highest e.e.S to produce (2R,3S)-EPG.
The ITS sequence analysis of the strain ZJUTZQ200 was carried out [16]. The sequence data had been submitted to GenBank under the accession NO. KJ534595. A phylogenetic tree (Figure 1) was further constructed, and this strain was closely clustered with Galactomyces geotrichum (GenBank accession No. JF262197), having 99% sequence identity. Based on the results of phylogenetic analysis and phenotypic tests, the isolated strain ZJUTZQ200 was designated as Galactomyces geotrichum ZJUTZQ200 and deposited in the China Center for Type Culture Collection (CCTCC M 2013114).
Figure 1. The phylogenetic tree based on 18S-ITS rDNA sequences, constructed by the neighbor-joining method, showing the relationship between the strain ZJUTZQ200 and the representatives of some related taxa.
Figure 1. The phylogenetic tree based on 18S-ITS rDNA sequences, constructed by the neighbor-joining method, showing the relationship between the strain ZJUTZQ200 and the representatives of some related taxa.
Molecules 19 08067 g001
G. geotrichum was usually reported as a lipase producer. Its lipase also probably hydrolyzed EPG in the bioresolution reaction, but our results indicated little EPG was hydrolyzed by the lipase in G. geotrichum ZJUTZQ200 (data not shown). Hitherto, EH produced by G. geotrichum ZJUTZQ200 has never been reported.

2.2. Substrate Specificity of Chiral Resolution by G. geotrichum ZJUTZQ200

The structure of the substrate can significantly affect enzyme selectivity [17,18]. In order to identify substrate specificity, we investigated the substrate scope of EH in G. geotrichum ZJUTZQ200. Nine racemic epoxides, including five phenylglycidate compounds, were resolved by the whole cells of G. geotrichum ZJUTZQ200. As seen in Table 1, G. geotrichum ZJUTZQ200 exhibited a relatively high enantioselectivity (e.e.S > 95%) towards phenylglycidate compounds 1, 2, 3 and 4, suggesting that it had potential in producing chiral phenylglycidate synthons. The enantioselectivity of G. geotrichum ZJUTZQ200 towards compound 5 with a methoxy group substituent on the phenyl ring was significantly lower than that of other phenylglycidate compounds. The epoxides 6, 7, 8 and 9 were resolved with low enantioselectivity. The results indicated that chiral resolution by EH of G. geotrichum ZJUTZQ200 was specific to EPG and its derivatives.
Table 1. Enantioselective hydrolysis of rac-EPG and its analogues by G. geotrichum ZJUTZQ200.
Table 1. Enantioselective hydrolysis of rac-EPG and its analogues by G. geotrichum ZJUTZQ200.
No.SubstrateReaction Time (h) aRetention Configuratione.e.S (%) bc (%)E Value c
1 Molecules 19 08067 i0014(2R,3S)99.46516
2 Molecules 19 08067 i0023(2R,3S)98.662.519
3 Molecules 19 08067 i0036(2R,3S)97.26810
4 Molecules 19 08067 i0043(2R,3S)95.162.16
5 Molecules 19 08067 i0055(2R,3S)60.32.83
6 Molecules 19 08067 i0066(2R,3S)7.363.41
7 Molecules 19 08067 i0076(2R,3S)41.364.12
8 Molecules 19 08067 i0086(2S,3R)17.324.14
9 Molecules 19 08067 i0094(2S,3R)72.4596
a Racemic epoxide (5 mM, dissolved in DMSO) was hydrolyzed using wet mycelium of G. geotrichum ZJUTZQ200 (1 g) in 10 mL KPB (100 mM, pH 7.0) at 30 °C, 200 rpm for several hours; b The e.e.S was calculated by the equation: e.e.S = (nR,SnS,R)/(nR,S + nS,R), where nR,S and nS,R are the concentrations of the (2R,3S)- and (2S,3R)-enantiomers, respectively; c The E value was enantiomeric selectivity which was calculated by the following equation: E = ln[(1 − c)(1 − e.e.S)]/ln[(1 − c)(1 + e.e.S)], c is the conversion rate of epoxide.

2.3. Optimization of Enzyme Production

In order to improve EH enzyme production, the culture conditions of G. geotrichum ZJUTZQ200, such as carbon and nitrogen sources, metal ions, the initial pH and culture time, etc. were optimized by single factor and orthogonal experiments. The optimal composition of enzyme production medium was as follows (per liter, pH 5.0): glycerin 10 g, soybean flour 15 g, NaNO3 3 g, MgSO4 0.5 g, KH2PO4 1 g, KCl 0.5 g. The biomass with EH was improved to 34.3 g/L from the initial 8.2 g/L.

2.4. Effect of Co-solvents on the Resolution of Rac-EPG

In enzymatic catalysis, a co-solvent can enhance the solubility of the substrate in water but also make it easier to get into the mycelium [19,20,21]. The effect of different organic solvents as co-solvents on the enantioselectivity was investigated at a concentration of 4‰ (v/v) (Figure 2). In comparison with the control without a supplementing co-solvent, the addition of different co-solvents all increased the e.e.S. Higher e.e.S and E values were obtained by the addition of DMSO, IPA and DMF. When acetone or ethanol was added, the conversion rate increased greatly so that E value was significantly decreased. In summary, DMSO as co-solvent afforded the best EH activity with the final e.e.S of 99.3% and E of 15. Therefore, DMSO was chosen as the optimal co-solvent.
Figure 2. Effects of co-solvents on the resolution of rac-EPG by G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 0.01 mL rac-EPG, 10 mL KPB (100 mM, pH 7.0), 30 °C, 200 rpm, 6 h.
Figure 2. Effects of co-solvents on the resolution of rac-EPG by G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 0.01 mL rac-EPG, 10 mL KPB (100 mM, pH 7.0), 30 °C, 200 rpm, 6 h.
Molecules 19 08067 g002
1: Dimethyl sulfoxide (DMSO), 2: tetrahydrofuran (THF), 3: acetone , 4: isopropanol (IPA), 5: ethanol, 6: N,N-dimethylformamide (DMF).

2.5. Effect of pH Buffer Solutions on Resolution of rac-EPG

The spatial conformation of the enzyme can be significantly affected by the type and pH value of the reaction buffer system [22], at the same time the dissociation situation of the substrate can also be influenced to some extent [23,24], thus affecting the binding of the enzyme and the substrate. To investigate the effects of pH and buffer type on the EH activity and enantioselectivity, the e.e.S and E values were examined in different buffer systems (citrate buffer (CB), KPB and Tris-HCl) with a pH range from 3.6 to 9.0 at 30 °C (Figure 3). The spontaneous hydrolysis rate of rac-EPG in different buffer systems was also examined. The results indicated that the enantioselectivity of the epoxide hydrolase produced by G. geotrichum ZJUTZQ200 was very sensitive to pH, which was consistent with Hellström’s report [25]. In acidic and alkaline buffer systems, the spontaneous hydrolysis rate was high, while it was relatively low around neutral pH. In citrate buffer, the e.e.S increased with pH and reached up to 94.7% at pH 6.4; meanwhile, the E value reached its maximum of 24 at pH 5.4. In KPB of pH 7.2, the e.e.S and E all reached their maxima of 98.5% and 17; besides, the spontaneous hydrolysis rate was in the lowest level (<10%). In Tris-HCl buffer, the e.e.S and E values were too low in comparison with the other two buffer systems, and the spontaneous hydrolysis rate was too high. As a result, KPB at pH 7.2 was chosen as the most favorable buffer solution.
Figure 3. Effects of buffer types and buffer pH on the resolution of rac-EPG by G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 0.01 mL rac-EPG, 10 mL buffer, 30 °C, 200 rpm and 6 h.
Figure 3. Effects of buffer types and buffer pH on the resolution of rac-EPG by G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 0.01 mL rac-EPG, 10 mL buffer, 30 °C, 200 rpm and 6 h.
Molecules 19 08067 g003

2.6. Effect of Substrate/Cell Ratio on Resolution of rac-EPG

Due to the toxicity of the epoxide moiety to microbial cells and the transport resistance of biocatalysts [26], the concentration of substrate and the amount of biocatalyst have great effects on the enantioselectivity and reactivity of EH [27]. As shown in Table 2, when the substrate/cell ratio (%, w/w) was less than 7%, the e.e.S of EPG could reach >99% after different bioconversion times. Under the condition of 10% mass ratio, the e.e.S of EPG only reached 78.1%, even if the reaction time was prolonged. E value increased with the mass ratio and reached its maximum value of 49 at the mass ratio of 5%. The possible reason is that when the substrate concentration is low (EPG/mycelium < 3%), the catalytic center of EH is not saturated, resulting in low catalytic efficiency manifested as a low E value. High substrate concentration (EPG/mycelium > 7%) may inhibit the activity of the enzyme’s active center, meanwhile the hydrolysis product can change the pH value of the buffer leading to a decrease of the EH activity, thus resulting in low enantiomeric selectivity (E value). High E value and e.e.S were all targets of biocatalysis optimization. e.e.S above 99.5% was especially favorable in the medicinal synthon preparation. Thus, we selected the mass ratio of 5% as an optimal substrate/cell ratio.
Table 2. Effect of substrate/cell ratio on enantioselectivity of epoxide hydrolase in G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 10 mL KPB (100 mM, pH 7.2), 30 °C, 200 rpm.
Table 2. Effect of substrate/cell ratio on enantioselectivity of epoxide hydrolase in G. geotrichum ZJUTZQ200. Reaction conditions: 1.0 g wet mycelium, 10 mL KPB (100 mM, pH 7.2), 30 °C, 200 rpm.
Sub./Cell (%)Reaction Time (h)e.e.S (%)c (%)E Vaule
1499.47511
2699.571.712
3899.872.528
5899.56949
7899.875.124
101378.174.73

2.7. Synthesis of Taxol’s C-13 Side Chain

Taxol’s C-13 side chain is of notable importance in the drug’s powerful antileukemic and tumor-inhibiting activity [28]. A new synthetic strategy towards taxol’s C-13 side chain ethyl (2R,3S)-3-benzoylamino-2-hydroxy-3-phenylpropionate (13) is shown in Scheme 2. Firstly, based on the optimal resolution process with G. geotrichum ZJUTZQ200, optically pure (2R,3S)-EPG (e.e.S > 99%) was obtained in a yield of 37.1%. Secondly, (2R,3S)-EPG was reacted with sodium azide in methanol to afford 11 in a yield of 95%. Subsequently, the reduction reaction with triphenylphosphine in THF to afford 12, followed by acylation with benzoyl chloride to obtain the desired ethyl (2R,3S)-3-benzoylamino-2-hydroxy-3-phenylpropionate (13, e.e.S > 95%) in a total yield of 33.8%. No racemization was observed during the whole formation of taxol’s C-13 side chain.
Scheme 2. Preparation of taxol C-13 side chain from rac-EPG.
Scheme 2. Preparation of taxol C-13 side chain from rac-EPG.
Molecules 19 08067 g005
Reaction conditions: (a) G. geotrichum ZJUTZQ200, KPB; (b) NaN3, NH4Cl, methanol, 65 °C; (c) triphenyl phosphine, THF, rt; (d) benzoyl chloride, dichloromethane, rt.

3. Experimental Section

3.1. Materials

rac-EPG 1 and microbial media components were purchased from J&K Scientific Ltd. (Shanghai, China). (2R,3S)-EPG was purchased from UHN Shanghai R&D Co., Ltd. (Shanghai, China). Other chemicals used were of analytical reagent grade and purchased from various commercial sources.

3.2. Preparation of Derivatives of EPG

Phenylglycidates 2, 3, 4 and 5 were synthesized by Darzens reaction [29]. Briefly, methyl chloroacetate (10 mmol) was dissolved in a solution of sodium methoxide (10 mmol) in dry methanol (10 mL) under ice bath cooling and stirred for 5 min, followed by the addition of corresponding aldehyde (5 mmol). The mixture was kept at 0 °C with stirring until completion and then the methanol was removed under reduced pressure. Dichloromethane (40 mL) was added and then washed with saturated sodium chloride solution (50 mL × 3), dried over anhydrous sodium sulfate and concentrated under vacuum and finally purified by thin layer chromatography (TLC) to give the desired derivatives of EPG. TLC was performed on F254 silica gel plates with the eluent petroleum ether/ethyl acetate (10:1, v/v). All the synthesized substrates were confirmed by LC-MS and 1H-NMR. NMR spectra were recorded on a Digital Avance 400 MHz spectrometer (Bruker) using tetramethylsilane as an internal standard.
Methyl 3-phenylglycidate (2): pale yellow liquid, yield 37%, 1H-NMR (CDCl3) δ 7.25 (s, 5H), 4.01 (s, 1H), 3.70 (d, J = 9.1 Hz, 3H), 3.42−3.39 (m, 1H). MS (ESI) m/z 178.0 (M)+.
Methyl 3-(2-methylphenyl) glycidate (3): pale yellow liquid, yield 42%, 1H-NMR (CDCl3) δ 7.05 (s, 2H), 6.72 (dd, J = 6.8, 1.9 Hz, 2H), 3.94 (s, 1H), 3.68–3.62 (m, 6H), 3.41 (s, 1H). MS (ESI) m/z 192.1 (M)+.
Methyl 3-(4-methylphenyl) glycidate (4): pale yellow liquid, yield 36%, 1H-NMR (CDCl3) δ 7.06 (s, 4H), 3.96 (d, J = 1.7 Hz, 1H), 3.96 (d, J = 1.7 Hz, 3H), 3.42 (s, 1H), 2.25 (s, 3H). MS (ESI) m/z 192.1 (M)+.
Methyl 3-(4-methoxyphenyl) glycidate (5): pale yellow liquid, yield 41%, 1H-NMR (CDCl3) δ 7.15–7.05 (m, 4H), 4.14 (s, 1H), 3.75 (s, 3H), 3.33 (s, 1H), 2.31 (s, 3H). MS (ESI) m/z 208.2 (M)+.

3.3. Screening of Strains and Enantioselective Hydrolysis to Rac-EPG

Soil samples were used for the screening of strains capable of utilizing 3-phenylglycidol as a sole carbon source. Isolated strains were transferred to 50 mL basal enzyme production medium and cultivated at 200 rpm, 30 °C for 48 h for bacteria, or 72 h for molds. After the cells were harvested, 1 g wet cells were washed and suspended in 10 mL 0.1 M, pH 7.0 potassium phosphate buffer (KPB). Cell suspensions were mixed with 50 μL substrate solution (EPG:DMSO = 2:8,V/V) and kept at 200 rpm, 30 °C for 6 h reaction, then the optical purity of the remaining substrate was determined by chiral HPLC. The screening medium contained (per liter, pH 6.5) 10 g soybean flour, 2 g NaNO3, 0.5 g KCl, 0.5 g MgSO4·7H2O, 1 g K2HPO4·3H2O, and 0.1% (v/v) 3-phenylglycidol. Basal enzyme production medium contained (per liter, pH 6.5) 10 g soybean flour, 2 g NaNO3, 0.5 g KCl, 0.5 g MgSO4·7H2O, 1 g K2HPO4·3H2O, and 30 g sucrose.

3.4. Identification of the Strain ZJUTZQ200

The isolated strain was preliminarily identified by morphological and microscopic observation. Furthermore, its 18S-internal transcribed spacer (ITS) regions of the rDNA were obtained through gene sequencing. The 18S-ITS region sequence was deposited in the GenBank database. Related sequences were obtained from GenBank database (National Center for Biotechnology Information) using the BLAST system. The 18S-ITS regions determined were aligned with the reference sequences obtained from GenBank databases using ClustalW ver.1.81 [30]. MEGA ver.5.1 was applied for the calculation of evolutionary distance and finally a phylogenetic tree was constructed using the neighbor-joining method [31,32].

3.5. Analytical Methods

The commercially available (2R,3S)-EPG was employed as a standard to identify the absolute configuration. The absolute configurations of the other products were assigned by comparing the measured specific rotations and retention times on chiral HPLC columns with the reported ones. EPG and its analogues were detected by HPLC (Waters 1525) equipped with a Daicel Chiralpack AS-H column (0.46 cm × 25 cm, 5 μm). The mobile phase was a mixture of n-hexane and isopropanol (8:2, v/v) at a flow rate of 0.8 mL/min. The value of e.e.S was expressed as enantiomeric excess of the remaining epoxide, which was calculated using the following equation: e.e.S = (nR,SnS,R)/(nR,S + nS,R) (nR,S and nS,R were the concentrations of the (2R,3S)- and (2S,3R)-enantiomer, respectively). E value was enantiomeric ratio of epoxides, which was calculated using the following equation: E = ln[(1 − c)(1 − e.e.S)]/ln[(1 − c)(1 + e.e.S)] (c, conversion rate) [33].

3.6. Synthesis of Taxol’s C-13 Side Chain

(2R,3S)-EPG (10): rac-EPG (1 mL) in DMSO (4 mL, 2:8, v/v) was dissolved in KPB (195 mL, 100 mM, pH 7.2) using wet mycelium of G. geotrichum ZJUTZQ200 (20 g) as biocatalyst at 30 °C (oil bathing) by stirring with a magnetic stir bar at 200 rpm. After 8 h biotransformation, the reaction solution was centrifuged to remove the mycelium and then extracted by n-butanol (60 mL × 3). The combined organic phase was concentrated under reduced pressure, followed by TLC purification with the eluent petroleum ether/ethyl acetate (10:1, v/v). 0.36 g (2R,3S)-EPG (e.e.S > 99%) was obtained. Molecules 19 08067 i010 = −148.6 (c 2, CHCl3) {lit. [34] Molecules 19 08067 i011 = −158.8 (c 1.06, CHCl3)}.
(2R,3S)-3-azido-2-hydroxy-benzenepropanoic acid ethyl ester (11): (2R,3S)-EPG (0.25 g, 1.3 mmol) was dissolved in methanol (10 mL) in round-bottom flask followed by addition of sodium azide (0.25 g, 3 eq.) and ammonium chloride (0.19 g, 3 eq.), then the mixture was stirred at 65 °C for 6 h. After completion of the reaction, the solvent was removed under reduced pressure and redissolved in ethyl acetate (30 mL). The organic phase was washed with saturated sodium chloride (2 × 30 mL) and dried over anhydrous sodium sulfate and then concentrated under reduced pressure to give 0.3 g (95% yield) of 11 as a pale yellow oil. 1H-NMR (CDCl3): δ 7.30 (s, 5H), 4.83 (s, 1H), 4.46 (s, 1H), 4.13 (s, 2H), 2.70 (s, 1H), 1.14 (s, 3H). MS (ESI) m/z 235.1 (M)+. Molecules 19 08067 i010 = +141.1 (c 1.8, CHCl3) {lit. [35] Molecules 19 08067 i012 = +133.5 (c 2, CH2Cl2)}.
(2R,3S)-3-amino-2-hydroxy-3-phenylpropionic acid ethyl ester (12): Compound 11 (0.3 g, 1.3 mmol) and triphenylphosphine (1 g, 3 eq.) were dissolved in tetrahydrofuran (30 mL) and then stirred at room temperature for 6 h. The solvent was removed under reduced pressure and then purified by TLC (eluent petroleum ether/ethyl acetate 2:1, v/v) to obtain 12 as pale yellow oil. 1H-NMR (CDCl3): δ 7.25–7.18 (m, 5H), 4.43 (s, 1H), 4.34 (s, 1H), 4.05–3.95 (m, 2H), 2.58 (s, 3H), 1.14 (s, 3H). MS (ESI) m/z 209.0 (M)+. Molecules 19 08067 i010 = −4 (c 1, CHCl3) {lit. [36] Molecules 19 08067 i010 = −2.9 (c 0.26, CHCl3)}.
Ethyl (2R,3S)-3-benzoylamino-2-hydroxy-3-phenylpropionate (13): Triethylamine (0.72 g, 1.5 eq.) was dissolved in a solution of 12 (0.1 g, 0.5 mmol) and benzoyl chloride (0.77 g, 1.5 eq.) in dichloromethane (20 mL) and then stirred at room temperature for 4 h. The solvent was removed under reduced pressure and then purified by TLC (eluent petroleum ether/ethyl acetate 2:1, v/v) to give 0.102 g of 13 (e.e.S > 95%) as white solid with a total yield of 33.8%. 1H-NMR (CDCl3): δ 7.68 (d, J = 8.4 Hz, 1H), 7.37–7.14 (m, 10H), 5.51 (dd, J = 8.5, 3.5 Hz, 1H), 4.55 (d, J = 3.5 Hz, 1H), 4.07–3.95 (m, 2H), 3.41 (s, 1H), 1.12 (t, J = 7.1 Hz, 3H). MS (ESI) m/z 336.1 (M)+. Molecules 19 08067 i010 = −11.4 (c 1, CHCl3) {lit. [37] Molecules 19 08067 i010 = –12 (c 2, CHCl3)}.

4. Conclusions

Bioresolution of rac-EPG is one of the important potential synthetic routes for industrial production of enantiopure (2R,3S)-EPG. In this work, we obtained a new strain capable of resolving rac-EPG to (2R,3S)-EPG with high enantioselectivity and identified it as G. geotrichum ZJUTZQ200. The use of this strain in enantioselective synthesis of (2R,3S)-EPG has not been previously reported. The substrate scope investigation results indicated G. geotrichum ZJUTZQ200 could afford high enantioselectivity in the synthesis of the chiral phenylglycidate synthon, suggesting that it had good potential in industrial applications. After the optimization of enzyme production and the whole cell catalysis process, G. geotrichum ZJUTZQ200 could resolve rac-EPG to (2R,3S)-EPG with e.e.S of 99.5% and E value of 49, both of which were higher than those reported in the literature for (2R,3S)-EPG preparation by enantioselective synthesis using whole cells containing EH [11,12]. Co-solvents, pH buffer solutions and substrate/cell ratio significantly affected the enantioselectivity of G. geotrichum ZJUTZQ200. Especially, the E value was improved to 49 from 17 at a substrate/cell ratio of 5%.
As a result, based on the optimized process of (2R,3S)-EPG synthesis by G. geotrichum ZJUTZQ200, we established a new chemoenzymatic synthesis route to produce taxol’s C-13 side chain, affording e.e.S of 95% and a total yield of 33.8%. In comparison with the chemoenzymatic synthesis route based on lipase-catalyzed asymmetric hydrolysis, our scheme is a promising alternative for the industrial preparation of taxol’s C-13 side chain due to EH’s advantages.

Acknowledgments

This work was supported by the National Natural Science Fund of China (No. 21272212) and the Natural Science Fund of Zhejiang Province, China (No. LY12B02018, LY12B06010).

Author Contributions

C.W. participated in designing the study. H.S. performed the experiments and collected the data. C.W. and J.L. analyzed the data and wrote the manuscript. Q.Z. directed the study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Borah, J.C.; Boruwa, J.; Barua, N.C. Synthesis of the C-13 side-chain of taxol. Curr. Org. Synth. 2007, 4, 175–199. [Google Scholar] [CrossRef]
  2. Vu, A.T.; Cohn, S.T.; Terefenko, E.A.; Moore, W.J.; Zhang, P.; Mahaney, P.E.; Trybulski, E.J.; Goljer, I.; Dooley, R.; Bray, J.A.; et al. 3-(Arylamino)-3-phenylpropan-2-olamines as a new series of dual norepinephrine and serotonin reuptake inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 2464–2467. [Google Scholar] [CrossRef]
  3. Li, X.; Zhu, C.; Li, C.; Wu, K.; Huang, D.; Huang, L. Synthesis of N-substituted clausenamide analogues. Eur. J. Med. Chem. 2010, 45, 5531–5538. [Google Scholar] [CrossRef]
  4. Gentile, A.; Giordano, C.; Fuganti, C.; Ghirotto, L.; Servi, S. The enzymic preparation of (2R, 3S)-phenyl glycidic acid esters. J. Org. Chem. 1992, 57, 6635–6637. [Google Scholar] [CrossRef]
  5. Pengsuparpapb, T.; Kingstond, D.G.; Neidighd, K.A.; Cordell, G.A.; Pezzuto, J.M. Evaluation of the cytotoxic mechanism mediated by baccatin III, the synthetic precursor of taxol. Chem. Biol. Interact. 1996, 101, 103–114. [Google Scholar] [CrossRef]
  6. Cusso, O.; Garcia-Bosch, I.; Ribas, X.; Lloret-Fillol, J.; Costas, M. Asymmetric epoxidation with H2O2 by manipulating the electronic properties of non-heme iron catalysts. J. Am. Chem. Soc. 2013, 135, 14871–14878. [Google Scholar] [CrossRef]
  7. Mishra, M.K.; Kumaraguru, T.; Sheelu, G.; Fadnavis, N.W. Lipase activity of Lecitase® Ultra: Characterization and applications in enantioselective reactions. Tetrahedron Asymmetry 2009, 20, 2854–2860. [Google Scholar] [CrossRef]
  8. Zheng, G.; Yuan, Q.; Yang, L.; Zhang, X.; Wang, J.; Sun, W. Preparation of (2S,3R)-methyl-3-phenylglycidate using whole cells of Pseudomonas putida. J. Mol. Catal. B Enzym. 2006, 43, 133–136. [Google Scholar] [CrossRef]
  9. Mamaghani, M.; Tabatabaeian, K.; Ghanadzadeh, A.; Habibi, F. An efficient enzymatic method for the separation of stereoisomeric cis and trans-glycidic esters synthesised via Darzen’s condensation reactions. Tetrahedron Lett. 2003, 44, 4775–4777. [Google Scholar] [CrossRef]
  10. Fretland, A.J.; Omiecinski, C.J. Epoxide hydrolases: Biochemistry and molecular biology. Chem. Biol. Interact. 2000, 129, 41–59. [Google Scholar] [CrossRef]
  11. Li, C.; Liu, Q.; Song, X.; Ding, D.; Ji, A.; Qu, Y. Epoxide hydrolase-catalyzed resolution of ethyl 3-phenylglycidate using whole cells of Pseudomonas sp. Biotechnol. Lett. 2003, 25, 2113–2116. [Google Scholar] [CrossRef]
  12. Zocher, F.; Enzelberger, M.M.; Bornscheuer, U.T.; Hauer, B.; Wohlleben, W.; Schmid, R.D. Epoxide hydrolase activity of Streptomyces strains. J. Biotechnol. 2000, 77, 287–292. [Google Scholar] [CrossRef]
  13. Kusari, S.; Singh, S.; Jayabaskaran, C. Rethinking production of Taxol (paclitaxel) using endophyte biotechnology. Trends Biotechnol. 2014, 32, 304–311. [Google Scholar] [CrossRef]
  14. Nakano, T.; Yagi, Y.; Miyahara, M.; Kaminura, A.; Kawatsura, M.; Itoh, T. Chemo-enzymatic synthesis of a multi-useful chiral building block molecule for the synthesis of medicinal compounds. Molecules 2011, 16, 6747–6757. [Google Scholar] [CrossRef]
  15. Vavříková, E.; Vacek, J.; Valentová, K.; Marhol, P.; Ulrichová, J.; Kuzma, M.; Křen, V. Chemo-enzymatic synthesis of silybin and 2,3-dehydrosilybin dimers. Molecules 2014, 19, 4115–4134. [Google Scholar] [CrossRef]
  16. Baldwin, B.G. Phylogenetic utility of the internal transcribed spacers of nuclear ribosomal DNA in plants: An example from the compositae. Mol. Phylogenet. Evol. 1992, 1, 3–16. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Sheng, Y.; Jiang, K.; Wang, Z.; Zheng, Y.; Zhu, Q. Bio-resolution of glycidyl (o,m,p)-methylphenyl ethers by Bacillus megaterium. Biotechnol. Lett. 2010, 32, 513–516. [Google Scholar] [CrossRef]
  18. Chang, C.S.; Ho, S.C. Enantioselective esterification of (R,S)-2-methylalkanoic acid with Carica papaya lipase in organic solvents. Biotechnol. Lett. 2011, 33, 2247–2253. [Google Scholar] [CrossRef]
  19. Pamperin, D.; Hopf, H.; Syldatk, C.; Pietzsch, M. Synthesis of planar chiral [2.2] paracyclophanes by biotransformations: Kinetic resolution of 4-formyl-[2.2]paracyclophane by asymmetric reduction. Tetrahedron Asymmetry 1997, 8, 319–325. [Google Scholar] [CrossRef]
  20. Leon, R.; Fernandes, P.; Pinheiro, H.M.; Cabral, J.M.S. Whole-cell biocatalysis in organic media. Enzym. Microb. Technol. 1998, 23, 483–500. [Google Scholar] [CrossRef]
  21. Zehentgruber, D.; Dragan, CA.; Bureik, M.; Lutz, S. Challenges of steroid biotransformation with human cytochrome P450 monooxygenase CYP21 using resting cells of recombinant Schizosaccharomyces pombe. J. Biotechnol. 2010, 146, 179–185. [Google Scholar] [CrossRef]
  22. Dong, H.P.; Wang, Y.J.; Zheng, Y.G. Enantioselective hydrolysis of diethyl 3-hydroxyglutarate to ethyl (S)-3-hydroxyglutarate by immobilized Candida antarctica lipase B. J. Mol. Catal. B Enzym. 2010, 66, 90–94. [Google Scholar]
  23. Long, W.S.; Kow, P.C.; Kamaruddin, A.H.; Bhatia, S. Comparison of kinetic resolution between two racemic ibuprofen esters in an enzymic membrane reactor. Process. Biochem. 2005, 40, 2417–2425. [Google Scholar] [CrossRef]
  24. Guo, J.L.; Mu, X.Q.; Xu, Y. Integration of newly isolated biocatalyst and resin-based in situ product removal technique for the asymmetric synthesis of (R)-methyl mandelate. Bioproc. Biosyst. Eng. 2010, 33, 797–804. [Google Scholar] [CrossRef]
  25. Hellström, H.; Steinreiber, A.; Mayer, S.F.; Faber, K. Bacterial epoxide hydrolase-catalyzed resolution of a 2,2-disubstituted oxirane: Optimization and upscaling. Biotechnol. Lett. 2001, 23, 169–173. [Google Scholar] [CrossRef]
  26. Siriphongphaew, A.; Pisnupong, P.; Wongkongkatep, J.; Inprakhon, P.; Vangnai, A.S.; Honda, K.; Ohtake, H.; Kato, J.; Ogawa, J.; Shimizu, S.; et al. Development of a whole-cell biocatalyst co-expressing the P450 monooxygenase and glucose dehydrogenase for synthesis of epoxyhexane. Appl. Microbiol. Biotechnol. 2012, 95, 357–367. [Google Scholar] [CrossRef]
  27. Wang, X.T.; Chen, X.H.; Xu, Y.; Lou, W.Y.; Wu, H.; Zong, M.H. Biocatalytic anti-Prelog stereoselective reduction of ethyl acetoacetate catalyzed by whole cells of Acetobacter sp. CCTCC M209061. J. Biotechnol. 2013, 163, 292–300. [Google Scholar] [CrossRef]
  28. Deng, L.; Jacobsen, E.N. A practical, highly enantioselective synthesis of the taxol side chain via asymmetric catalysis. J. Org. Chem. 1992, 57, 4320–4323. [Google Scholar] [CrossRef]
  29. Thijs, L.; Orskamp, J.J.; Adriaan, A.; Derks, M.P.W.; Feenstra, R.W.; Legters, J.; Zwanenburg, B. Synthesis of all four homochiral stereoisomers of methyl 3-phenyl-1h-aziridine-2-carboxylate. Tetrahedron 1990, 46, 2611–2622. [Google Scholar] [CrossRef]
  30. Thompson, J.D.; Higgins, D.G.; Gibson, T.J. CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, positions-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994, 22, 4673–4680. [Google Scholar] [CrossRef]
  31. Xu, F.; Deng, G.; Cheng, S.; Zhang, W.; Huang, X.; Li, L.; Cheng, H.; Rong, X.; Li, J. Molecular cloning, characterization and expression of the phenylalanine ammonia-lyase gene from Juglans Regia. Molecules 2012, 17, 7810–7823. [Google Scholar] [CrossRef]
  32. Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol. 2011, 28, 2731–2739. [Google Scholar] [CrossRef]
  33. Chen, C.S.; Fujimoto, Y.; Girdaukas, G.; Sih, C.J. Quantitative analyses of biochemical kinetic resolutions of enantiomers. J. Am. Chem. Soc. 1982, 104, 7294–7299. [Google Scholar] [CrossRef]
  34. Kakei, H.; Tsuji, R.; Ohshima, T.; Morimoto, H.; Matsunaga, S.; Shibasaki, M. Catalytic asymmetric epoxidation of α,β-unsaturated esters with chiral Yttrium–Biaryldiol complexes. Chem. Asian J. 2007, 2, 257–264. [Google Scholar] [CrossRef]
  35. Hönig, H.; Seufer-Wasserthal, P.; Weber, H. Chemo-enzymatic synthesis of all isomeric 3-phenylserines and-isoserines. Tetrahedron 1990, 46, 3841–3850. [Google Scholar] [CrossRef]
  36. Forró, E.; Fülöp, F. A new enzymatic strategy for the preparation of (2R,3S)-3-phenylisoserine: A key intermediate for the Taxol side chain. Tetrahedron Asymmetry 2010, 21, 637–639. [Google Scholar] [CrossRef]
  37. Feske, B.D.; Kaluzna, I.A.; Stewart, J.D. Enantiodivergent, biocatalytic routes to both taxol side chain antipodes. J. Org. Chem. 2005, 70, 9654–9657. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the synthetic compounds are available in milligram amounts from the authors.

Share and Cite

MDPI and ACS Style

Wei, C.; Ling, J.; Shen, H.; Zhu, Q. Bioresolution Production of (2R,3S)-Ethyl-3-phenylglycidate for Chemoenzymatic Synthesis of the Taxol C-13 Side Chain by Galactomyces geotrichum ZJUTZQ200, a New Epoxide-Hydrolase-Producing Strain. Molecules 2014, 19, 8067-8079. https://doi.org/10.3390/molecules19068067

AMA Style

Wei C, Ling J, Shen H, Zhu Q. Bioresolution Production of (2R,3S)-Ethyl-3-phenylglycidate for Chemoenzymatic Synthesis of the Taxol C-13 Side Chain by Galactomyces geotrichum ZJUTZQ200, a New Epoxide-Hydrolase-Producing Strain. Molecules. 2014; 19(6):8067-8079. https://doi.org/10.3390/molecules19068067

Chicago/Turabian Style

Wei, Chun, Jinlong Ling, Honglei Shen, and Qing Zhu. 2014. "Bioresolution Production of (2R,3S)-Ethyl-3-phenylglycidate for Chemoenzymatic Synthesis of the Taxol C-13 Side Chain by Galactomyces geotrichum ZJUTZQ200, a New Epoxide-Hydrolase-Producing Strain" Molecules 19, no. 6: 8067-8079. https://doi.org/10.3390/molecules19068067

Article Metrics

Back to TopTop