Next Article in Journal
A Study of the Effect of Sintering Conditions of Mg0.95Ni0.05Ti3 on Its Physical and Dielectric Properties
Previous Article in Journal
A Compressive Review about Taxol®: History and Future Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Synthesis of New Fluorinated β-Amino Acid Enantiomers through Lipase-Catalyzed Hydrolysis

by
Sayeh Shahmohammadi
1,
Ferenc Fülöp
1,2 and
Enikő Forró
1,2,*
1
Institute of Pharmaceutical Chemistry, University of Szeged, H-6720 Szeged, Hungary
2
Stereochemistry Research Group of the Hungarian Academy of Sciences, University of Szeged, H-6720 Szeged, Hungary
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(24), 5990; https://doi.org/10.3390/molecules25245990
Submission received: 27 October 2020 / Revised: 14 December 2020 / Accepted: 14 December 2020 / Published: 17 December 2020

Abstract

:
An efficient and novel enzymatic method has been developed for the synthesis of β-fluorophenyl-substituted β-amino acid enantiomers through lipase PSIM (Burkholderia cepasia) catalyzed hydrolysis of racemic β-amino carboxylic ester hydrochloride salts 3ae in iPr2O at 45 °C in the presence of Et3N and H2O. Adequate analytical methods were developed for the enantio-separation of racemic β-amino carboxylic ester hydrochlorides 3ae and β-amino acids 2ae. Preparative-scale resolutions furnished unreacted amino esters (R)-4ae and product amino acids (S)-5ae with excellent ee values (≥99%) and good chemical yields (>48%).

Graphical Abstract

1. Introduction

In recent years, enantiomerically pure β-aryl-substituted β-amino acids have been intensively investigated due to their pharmacological significance, unique and remarkable biological properties [1], their utility in synthetic chemistry [2], and drug research [3]. Therefore, this class of compounds has been documented as a crucial scaffold in the design and synthesis of conceivable pharmaceutical drugs. For instance, 3-amino-3-phenylpropionic acid, which is a key pharmaceutical building block, is present in anticancer agents, such as Taxol [4]. It can also find application as a fundamental component in the synthesis of novel antibiotics [5] and analgesic endomorphine-1 analogue tetrapeptides [6].
On the other hand, tremendous achievements in the development of fluorinated amino acid drugs verified the high importance of this type of compounds in pharmaceutical chemistry. It is known that the occurrence of fluorine in biologically active natural compounds is extremely low. In turn, the number of fluorine-containing drugs on the market is rising continuously. The reasons are the unique characteristics of the fluorine atom in terms of its high electronegativity and the polarity of a carbon–fluorine bond [7,8]. Thus, incorporation of fluorine into β-amino acids has gained increasing attention in recent decades. For example, Januvia (sitagliptin phosphate) acts as an antidiabetic agent via inhibition of dipeptidyl peptidase IV [9], whereas (±)-Eflornithine was used for the treatment of trypanosomiasis [10] and against facial hirsutism in women [11].
There are different approaches for the synthesis of optically active β-aryl-β-amino acids [12,13,14]. The utilization of enzymes in these reactions gained special attention, which is due to their ability to conduct the reactions enantio-selectively. For example, lipases are stable. They work under mild conditions and many of them are commercially available. They can be applied on an industrial scale [15,16]. Lipase-catalyzed methods for the resolution of both cyclic [17] and acyclic [18] β-amino carboxylic esters through hydrolysis are known in the literature. Various enzymatic procedures have been developed by our research group for the preparation of biologically active β-aryl-substituted, β-heteroaryl-substituted, and β-arylalkyl-substituted β-amino acid enantiomers through enantioselective (E > 200) hydrolysis of the corresponding β-amino carboxylic esters both in H2O or in an organic solvent catalyzed by lipase (Pseudomonas cepacia) PS[19,20,21]. Catalyzed kinetic and dynamic kinetic resolution of β-amino carboxylic esters or their hydrochloride salts with tetra-hydro-isoquinoline and tetra-hydro-β-carboline skeleton through hydrolysis have been performed. Catalysts used include Candida antarctica lipase B (CAL-B) (in aqueous NH4OAc buffer at pH 8.5 and or in iPr2O in the presence of 1 equiv of H2O), Alcalase (in borate buffer at pH 8), and lipase PS (in iPr2O with 4 equiv of added H2O) [22,23,24,25].
A number of N-benzylated β2-, β3-, and β2,3-amino acids were prepared through CAL-B-catalyzed hydrolysis of the corresponding racemic amino carboxylic esters with H2O in n-hexane or 2-methyl-2-butanol, under stirring [26] or utilizing high-speed ball-milling conditions [27]. Covalently immobilized lipase AK (Pseudomonas fluorescens) and lipase PS were used as efficient stereoselective catalysts for the kinetic resolution of exotic and variously substituted rac-(5-phenylfuran-2-yl)-β-alanine ethyl ester hydrochlorides through hydrolysis (E > 146) in NH4OAc buffer (20 mM, pH 5.8) at 30 °C [28]. Gotor et al. reported the kinetic resolution (E > 200) of a large number of 3-amino-3-phenylpropanoate esters through lipase PS-catalyzed hydrolysis with H2O in 1,4-dioxane at 45 °C [18]. The method was successfully used for the synthesis of (S)-3-amino-3-phenylpropionic acid, which is a key precursor for the preparation of (S)-dapoxetine, and a potent selective serotonin reuptake inhibitor (SSRI) used for the treatment of depression, bulimia, or anxiety [29]. In addition, very recently, Zhang et al. summarized a review of the most facile catalytic enantioselective strategies to construct the fluorinated α-amino and β-amino acids [30].
Herein, in view of the importance of fluorinated β-amino acids, our aim was to synthesize (±)-β-amino carboxylic ester hydrochloride salts 3ae (Scheme 1), then to devise a suitable enzymatic protocol for the synthesis of new fluorinated β-amino acids via enantioselective hydrolysis of 3ae (Scheme 2) and provide an adequate characterization of the enantiomeric products.

2. Results

2.1. Synthesis of Ethyl 3-Amino-3-Arylpropanoate Hydrochloride Salts (±)-3ae

Racemic β-amino acids (±)-2ae were synthesized by modified Rodionov synthesis through the reaction of the corresponding aldehydes with malonic acid in the presence of NH4OAc in EtOH at a reflux temperature (Scheme 1) [31,32]. Subsequently, the β-amino carboxylic ester hydrochloride salts (±)-3ae were prepared with yields ranging from 76% to 98% by esterification of the corresponding β-amino acids in the presence of SOCl2 in EtOH.

2.2. Enzyme-Catalyzed Hydrolysis of (±)-3ae

2.2.1. Preliminary Experiments

On the basis of the results achieved on the enzyme-mediated enantioselective hydrolysis of β-amino carboxylic esters [17,18], the hydrolysis of model compound (±)-3a (Scheme 2) was conducted with 5 equiv of Et3N and 0.5 equiv of H2O in the presence of 30 mg mL–1 enzyme in iPr2O at 45 °C (Table 1, entry 1). In the frame of enzyme screening, lipase AY (Candida rugosa), lipase AK, PPL (Porcine pancreatic lipase), and CAL-B (Table 1, entries 2–5) showed activity in enzymatic hydrolysis. However, with the exception of lipase AK affording an eep value of 75% and a moderate E (8) (19% conversion in 10 min, entry 3), low reactivities and low enantio-selectivities were achieved (entries 2, 4, and 5). It is noteworthy that PPL catalyzed the reaction with opposite enantio-selectivity. Lipase PSIM, in contrast, provided an E value of 108 (entry 1) and, consequently, it was selected for further studies.
Next, we analyzed the effect of solvent on enantio-selectivity and reaction rate. Very different E and reaction rate data were observed in the green solvents tested (Table 2). The hydrolytic reactions of 3a in the ether-type solvents were rapid (conv. 52%, 51%, and 54% after 10 min, E = 59, 113, and 27, entries 1, 2, and 4), while, in EtOAc, the reaction proceeded relatively slowly with low enantioselectivity (conv. 11%, after 10 min, E = 3, entry 3). On the basis of our earlier results [33], the reaction was also performed under solvent-free conditions when, in harmony with our earlier observation, a reasonable enantioselectivity (E 74) was observed in addition to a rapid transformation (conv. 49% after 10 min, entry 5). For the reason of economy (taking into account that 2-Me-THF is the most expensive selected solvent), despite the highest E (113), iPr2O, with a slightly lower E (108), but as significantly less expensive was identified as the most suitable solvent.
In order to follow up the progress of the reaction while maintaining high enantioselectivity, it was wise to slow down the reaction. When the reaction temperature decreased from 45 °C to 25 °C, both the reaction rate and enantio-selectivity for the hydrolysis of 3a clearly decreased (30% conv. in 10 min, E = 48 vs. 48% conv. in 10 min, E = 108, Table 1, entry 1). To our surprise, the fastest reaction was achieved at 3 °C with the highest degree of conversion (50% in 10 min) and an E value of 134. In order to collect more information, we decided to carry out the reaction with different enzyme concentrations at 3 °C. As shown in Table 3, there was no significant difference in the reaction rates, when the enzyme concentration decreased from 10 to 5 or 2 mg mL–1 (~50% conv. in 10 min reaction time, entries 1–3). In contrast, E dropped significantly when the reaction was performed with a 2-mg mL–1 enzyme (entry 3). Since both high E and satisfactory reaction rate were attained at 45 °C, we decided to use this optimal reaction temperature for preparative-scale reactions. Additionally, a set of preliminary experiments was performed in order to determine the influence of enzyme concentration on the reaction rate (Table 4). The reaction rate for the hydrolysis of (±)-3a clearly increased as the concentration of enzymes was increased. The highest reaction rate was observed with a 40-mg mL–1 enzyme (entry 5). However, for a satisfactory reaction time (the time needed to reach 50% conversion), the use of a 30-mg mL–1 enzyme (Table 1, entry 1) was selected for preparative-scale resolutions.

2.2.2. Preparative-Scale Resolutions of (±)-3ae

Preparative-scale resolution of (±)-3ae were performed under the optimized conditions (30 mg mL–1 lipase PSIM, 0.5 equiv Et3N, 0.5 equiv H2O, iPr2O, 45 °C) to yield the unreacted β-amino carboxylic ester enantiomers (R)-4a–e and product β-amino acids (S)-5ae with excellent ee (≥99%) and good yields (>48%) (Table 5).

2.2.3. Determination of Absolute Configurations

The absolute configurations in the cases of (S)-5a {[α] = −3.1 (c 0.28, H2O), lit. [36] [α] = −3.9 (c 1.0, H2O)}, (S)-5b {[α] = −5.0 (c 0.26, H2O), lit. [36] [α] = −4.5 (c 0.50, H2O)}, (S)-5c {[α] = −3.0 (c 0.28, H2O), lit. [37] [α] = +3.9 (c 0.40, H2O for the antipode (R))}, (R)-4a {[α] = +17.9 (c 0.44, CHCl3), lit. [36] [α] = −21.5 (c 1.0, CHCl3 for antipode (S))}, (R)-4b {[α] = +9.0 (c 0.29, CHCl3), lit. [36] [α] = −12.6 (c 0.50, CHCl3 for antipode (S))} and (R)-4c {[α] = +18.9 (c 0.41, CHCl3), lit. [18] [α] = +18.5 (c 1.0, CHCl3)} were assigned by comparing the [α] values with literature data. Taking into consideration the most stable conformation of the 3-amino-3-phenylpropanoate core matching nicely, the (S)-configuration of the products [18] and the GC chromatograms analyzed, the same enantio-preference for the (S)-selective hydrolysis by Lipase PSIM for 5d and 5e was indicated.

3. Experimental Section

3.1. General Methods

Lipase PSIM and lipase AK were from Amano Pharmaceuticals and lipase AY was from Fluka. PPL and CAL-B immobilized on acrylic resin were purchased from Sigma (Budapest, Hungary). Substituted benzaldehydes were from Sigma-Aldrich. Triethylamine was from Merck. Solvents of the highest analytical grade were from Sigma-Aldrich. Optical rotations were measured with a Perkin-Elmer 341 Polarimeter. 1H-NMR spectra were recorded on a Bruker Avance DRX 500 spectrometer. Melting points were determined on a Kofler apparatus (see the Supplementary Materials). The enantiomeric excess ee values for the unreacted β-amino carboxylic ester and the β-amino acid enantiomers produced were determined by GC equipped with a Chirasil-L-Val column after double derivatization [34] with (i) diazomethane [Caution: the derivatization with diazomethane should be performed under a well-working hood] and (ii) acetic anhydride in the presence of 4-dimethylaminopyridine and pyridine [90 °C for 10 min → 170 °C (temperature rise 20 °C min−1), 10 psi]. Retention times (min) for 4a: 36.308 (antipode: 36.535), for 5a: 32.365 (antipode: 31.515), for 4b: 32.137 (antipode: 32.550), for 5b: 29.031 (antipode: 28.282), for 4c: 33.305 (antipode: 33.860), for 5c: 29.905 (antipode: 29.528), for 4d: 26.064 (antipode: 26.187), for 5d: 23.766 (antipode: 23.463), for 4e: 35.421 (antipode: 36.018), for 5e: 30.946 (antipode: 30.541)].

3.2. General Procedure for the Syntheses of Racemic β-Amino Acids 2ae

Compounds 2ae were synthesized based on the modified Rodionov synthesis [31,32] through condensation of the corresponding aldehydes 1ae (2 g) with malonic acid (1 equiv) in the presence of NH4OAc (2 equiv) in EtOH under reflux for 8 h [19]. The resulting white crystals were filtered off and washed with acetone and then they were recrystallized from H2O and acetone.

3.2.1. (±)-3-Amino-3-(3,4-Difluorophenyl) Propionic Acid 2a

Yield: (1.2 g, 42%), mp 233–235 °C. 1H-NMR (D2O, 500 MHz): δ = 7.30–7.26 (m, 1H, Ar), 7.25–7.21 (m, 1H, Ar), 7.15–7.13 (m, 1H, Ar), 4.54 (t, J = 7.33 Hz, 1H, CH), 2.78 (dd, JAB = 16.31 Hz, JAX = 7.89 Hz, 1H, C(2)HA), 2.70 (dd, JBA = 16.27 Hz, JBX = 6.78 Hz, 1H, C(2)HB).

3.2.2. (±)-3-Amino-3-(3,5-Difluorophenyl) Propionic Acid 2b

Yield: (1.36 g, 48%), mp 237–239 °C. 1H-NMR (D2O, 500 MHz): δ = 6.98–6.97 (m, 2H, Ar), 6.93–6.89 (m, 1H, Ar), 4.56 (t, J = 7.23 Hz, 1H, CH), 2.78 (dd, JAB = 16.34 Hz, JAX = 7.84 Hz, 1H, C(2)HA), 2.71 (dd, JBA = 16.34 Hz, JBX = 6.61 Hz, 1H, C(2)HB).

3.2.3. (±)-3-Amino-3-(4-Fluorophenyl) Propionic Acid 2c

Yield: (1.50 g, 51%), mp 244–246 °C. 1H-NMR (D2O, 500 MHz): δ = 7.37–7.34 (m, 2H, Ar), 7.11–7.07 (m, 2H, Ar), 4.54 (t, J = 7.31 Hz, 1H, CH), 2.81 (dd, JAB = 16.10 Hz, JAX = 8.05 Hz, 1H, C(2)HA), 2.72 (dd, JBA = 16.10 Hz, JBX = 6.73 Hz, 1H, C(2)HB).

3.2.4. (±)-3-Amino-3-(2-Fluoro-4-Triflouromethylphenyl) Propionic Acid 2d

Yield: (0.786 g, 30%), mp 251–253 °C. 1H-NMR (D2O, 500 MHz): δ = 7.75–7.72 (m, 1H, Ar), 7.62–7.57 (m, 2H, Ar), 4.89 (t, J = 7 Hz, 1H, CH), 2.82 (dd, JAB = 16.63 Hz, JAX = 8.63 Hz, 1H, C(2)HA), 2.77 (dd, JBA = 16.86 Hz, JBX = 5.31 Hz, 1H, C(2)HB).

3.2.5. (±)-3-Amino-3-(2-Fluoro-4-Methylphenyl) Propionic Acid 2e

Yield: (0.4275 g, 15%), mp 239–241 °C. 1H-NMR (D2O, 500 MHz): δ = 7.39–7.36 (m, 1H, Ar), 7.11–7.03 (m, 2H, Ar), 4.84 (t, J = 7.78 Hz, 1H, CH), 2.80 (dd, JAB = 16.81 Hz, JAX = 9.82 Hz, 1H, C(2)HA), 2.69 (dd, JBA = 16.75 Hz, JBX = 4.24 Hz, 1H, C(2)HB), 2.38 (s, 3H, CH3).

3.3. General Procedure for the Syntheses of Racemic β-Amino Carboxylic Ester Hydrochloride Salts 3ae

SOCl2 (1.05 equiv) was added to 30 mL of EtOH at a temperature kept under –15 °C with saline ice. To this solution, 2ae (1 g) were added at once. The mixture was stirred at 0 °C for 30 min, then at room temperature for 3 h, and, finally, heated under reflux for 1 h. The solvent was evaporated off under reduced pressure and the resulting white 3ae. HCl salts were recrystallized from EtOH and Et2O.

3.3.1. Hydrochloride Salt of Ethyl (±)-3-Amino-3-(3,4-Difluorophenyl) Propanoate 3a. HCl

Yield: (1 g, 76%), mp 142–144 °C. 1H-NMR (D2O, 500 MHz): δ = 7.32–7.28 (m, 1H, Ar), 7.27–7.23 (m, 1H, Ar), 7.18–7.16 (m, 1H, Ar), 4.69 (t, J = 8.73 Hz, 1H, CH), 4.05–4.01 (m, 2H, CH2), 3.08 (dd, JAB = 16.77 Hz, JAX = 7.20 Hz, 1H, C(2)HA), 3.00 (dd, JBA = 16.77 Hz, JBX = 7.50 Hz, 1H, C(2)HB), 1.05 (t, J = 7.09 Hz, 3H, CH3). 13C-NMR (D2O, 126 MHz): δ = 13.2, 37.9, 50.7, 62.5, 116.6 (d, 2JC–F = 18.42 Hz), 118.4 (d, 2JC–F = 17.65 Hz), 124.1 (dd, 3JC–F = 7.10 Hz, 4JC–F = 3.64 Hz), 132.2 (d, 3JC–F = 3.87 Hz), 150.2 (dd, 1JC–F = 250.46 Hz, 2JC–F = 16.46 Hz), 150.6 (dd, 1JC–F = 241.81 Hz, 2JC–F = 5.74 Hz), 171.3. 19F-NMR (D2O, 471 MHz): δ = −136.6 Hz, −136.9 Hz.

3.3.2. Hydrochloride Salt of Ethyl (±)-3-Amino-3-(3,5-Difluorophenyl) Propanoate 3b. HCl

Yield: (1.15 g, 87%), mp 182–184 °C. 1H-NMR (D2O, 500 MHz): δ = 7.02–7.01 (m, 2H, Ar), 6.97–6.93 (m, 1H, Ar), 4.74 (t, J = 7.20 Hz, 1H, CH), 4.07–4.02 (m, 2H, CH2), 3.11 (dd, JAB = 16.98 Hz, JAX = 7.24 Hz, 1H, C(2)HA), 2.81 (dd, JBA = 16.96 Hz, JBX = 7.23 Hz, 1H, C(2)HB), 1.08 (t, J = 6.98 Hz, 3H, CH3). 13C-NMR (D2O, 126 MHz): δ = 15.6, 40.2, 53.2, 65.0, 107.6 (t, 2JC–F = 25.43 Hz), 113.0 (dd, 2JC–F = 20.28 Hz, 4JC–F = 6.70 Hz), 141.0 (t, 3JC–F = 9.47 Hz), 165.6 (dd, 1JC–F = 248.20 Hz, 3JC–F = 13.04 Hz), 173.7. 19F-NMR (D2O, 471 MHz): δ = −108.3 Hz.

3.3.3. Hydrochloride Salt of Ethyl (±)-3-Amino-3-(4-Flourophenyl) Propanoate 3c. HCl

Yield: (1.11 g, 82%), mp 165–167 °C. 1H-NMR (D2O, 500 MHz): δ = 7.39–7.37 (m, 2H, Ar), 7.13–7.09 (m, 2H, Ar), 4.70 (t, J = 7.82 Hz, 1H, CH), 4.04–3.99 (m, 2H, CH2), 3.09 (dd, JAB = 16.61 Hz, JAX = 7.17 Hz, 1H, C(2)HA), 3.00 (dd, JBA = 16.66 Hz, JBX = 7.65 Hz, 1H, C(2)HB), 1.05 (t, J = 7.21 Hz, 3H, CH3). 13C-NMR (D2O, 126 MHz): δ = 13.3, 38.3, 51.1, 62.6, 116.3 (d, 2JC–F = 22.08 Hz), 129.5 (d, 3JC–F = 8.94 Hz), 131.3 (d, 4JC–F = 2.99 Hz), 163.2 (d, 1JC–F = 246.30 Hz),171.6. 19F-NMR (D2O, 471 MHz): δ = −112.3 Hz.

3.3.4. Hydrochloride Salt of Ethyl (±)-3-Amino-3-(2-Fluoro-4-Trifluoromethylphenyl) Propanoate 3d. HCl

Yield: (1.2 g, 95%), mp 133–135 °C. 1H-NMR (D2O, 500 MHz): δ = 7.61–7.58 (m, 1H, Ar), 7.56–7.53 (m, 2H, Ar), 5.03 (t, J = 7.31 Hz, 1H, CH), 4.07–4.02 (m, 2H, CH2), 3.20 (dd, JAB = 16.90 Hz, JAX = 7.27 Hz, 1H, C(2)HA), 3.12 (dd, JBA = 16.94 Hz, JBX = 7.35 Hz, 1H, C(2)HB), 1.07 (t, J = 7.51 Hz, 3H,CH3). 13C-NMR (D2O, 126 MHz): δ = 13.1, 36.9, 45.8 (d, 3JC–F = 2.69 Hz), 62.6, 113.9 (dq, 2JC–F = 25.50 Hz, 3JC–F = 3.80 Hz), 123 (q, 1JC–F = 274.12 Hz), 122.2 (m), 126.3 (d, 2JC–F = 12.87 Hz), 129.8 (d, 3JC–F = 3.44 Hz), 133.2 (m), 160.0 (d, 1JC–F = 248.95 Hz), 171.1. 19F-NMR (D2O, 471 MHz): δ = −62.6 Hz, −114.9 Hz.

3.3.5. Hydrochloride Salt of Ethyl (±)-3-Amino-3-(2-Fluoro-4-Methylphenyl) Propanoate 3e. HCl

Yield: (1.30 g, 98%), mp 172–174 °C. 1H-NMR (D2O, 500 MHz): δ = 7.26–7.23 (m, 1H, Ar), 7.03–6.99 (m, 2H, Ar), 4.90 (t, J = 7.51 Hz, 1H, CH), 4.05–4.01 (m, 2H, CH2), 3.15 (dd, JAB = 16.55 Hz, JAX = 7.32 Hz, 1H, C(2)HA), 3.06 (dd, JBA = 16.57 Hz, JBX = 7.73 Hz, 1H, C(2)HB), 2.25 (s, 3H, CH3), 1.06 (t, J = 7.12 Hz, 3H, CH3,). 13C-NMR (D2O, 126 MHz): δ = 13.1, 20.3, 37.2, 46.1 (d, 3JC–F = 2.77 Hz), 62.5, 116.6 (d, 2JC–F = 21.22 Hz), 118.8 (d, 2JC–F = 13.22 Hz,),125.8 (d, 4JC–F = 2.85 Hz), 128.3 (d, 3JC–F = 3.71 Hz), 143.3 (d, 3JC–F = 8.42 Hz), 160.1 (d, 1JC–F = 245.57 Hz), 171.5. 19F-NMR (D2O, 471 MHz): δ = −118.5 Hz.

3.4. General Procedure for the Preparative-Scale Resolutions of (±) 3ae

Racemic hydrochloride salts 3ae (200 mg) were dissolved in iPr2O (10 mL). Lipase PSIM (30 mg mL−1), Et3N (5 equiv), and H2O (0.5 equiv) were added and the mixture was shaken in an incubator shaker at 45 °C for Rt: 8 h 3a, 72 h 3b, 18 h 3c, 26 h 3d, 23 h 3e (Table 5). Reactions were stopped by filtering off the enzyme at close to 50% conversion. The filtered enzyme was washed with Et2O (3 × 15 mL). The solvents were dried by using Na2SO4, and then evaporated off to yield unreacted β-amino carboxylic esters (R)-4ae. The filtered enzyme was washed with distilled H2O (3 × 15 mL). Then evaporation of the filtrate yielded the crystalline (S)-5ae products, which where recrystallized from EtOH and H2O. All of the enantiomers formed in enzymatic reactions were isolated in basic form due to a relatively slow in situ liberation of basic amino ester from its hydrochloric salt, which is followed by enzymatic hydrolysis.

3.4.1. (R)-Ethyl 3-Amino-3-(3,4-Difluorophenyl) Propanoate 4a

Yield: (84 mg, 48.7%), [α] = +17.9 (c 0.44, CHCl3), lit. [36] [α] = –21.5 (c 1.0, CHCl3 for antipode (S)), [α] = +4.1 (c 0.33, EtOH). The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 3a.

3.4.2. (R)-Ethyl 3-Amino-3-(3,5-Difluorophenyl) Propanoate 4b

Yield: (66 mg, 38.26%), [α] = +9.0 (c 0.29, CHCl3), lit. [36] [α] = –12.6 (c 0.50, CHCl3 for antipode (S)), [α] = –5.0 (c 0.20, EtOH). The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 3b.

3.4.3. (R)-Ethyl 3-Amino-3-(4-Fluorophenyl) Propanoate 4c

Yield: (83 mg, 49%), [α] = +18.9 (c 0.41, CHCl3), lit. [18] [α] = +18.5 (c 1.0, CHCl3), [α] = +12.8 (c 0.32, EtOH). The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 3c.

3.4.4. (R)-Ethyl 3-Amino-3-(2-Fluoro-4-Triflouromethylphenyl) Propanoate 4d

Yield: (84.6 mg, 47.83%), [α] = +20.3 (c 0.53, CHCl3), [α] = +13.7 (c 0.30, EtOH). The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 3d.

3.4.5. (R)-Ethyl 3-Amino-3-(2-Fluoro-4-Methylphenyl) Propanoate 4e

Yield: (94.8 mg, 47.4%), [α] = +16.0 (c 0.13, CHCl3), [α] = +6.0 (c 0.21, EtOH). The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 3e.

3.4.6. (S)-3-Amino-3-(3,4-Difluorophenyl) Propionic Acid 5a

Yield: (72.8 mg, 48.08%), [α] = −3.1 (c 0.28, H2O), lit. [36] [α] = −3.9 (c 1.0, H2O), mp 229–231 °C, lit. [38] mp 226–230 °C. The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 2a.

3.4.7. (S)-3-Amino-3-(3,5-Difluorophenyl) Propionic Acid 5b

Yield: (73 mg, 48.22%), [α] = −5.0 (c 0.26, H2O), lit. [36] [α] = −4.5 (c 0.50, H2O), mp 256–258 °C. The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 2b.

3.4.8. (S)-3-Amino-3-(4-Fluorophenyl) Propionic Acid 5c

Yield: (71.7 mg, 48.5%), [α] = −3.0 (c 0.28, H2O), lit. [37] [α] = +3.9 (c 0.40, H2O for antipode (R)), mp 242–244 °C, lit. [37] mp 245–247 °C. The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 2c.

3.4.9. (S)-3-Amino-3-(2-Fluoro-4-Trifluoromethylphenyl) Propionic Acid 5d

Yield: (77.5 mg, 48.7%), [α] = −11.0 (c 0.19, MeOH), mp 255–257 °C. The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 2d.

3.4.10. (S)-3-Amino-3-(2-Fluoro-4-Methylphenyl) Propionic Acid 5e

Yield: (84.8 mg, 48.42%), [α] = −13.0 (c 0.21, MeOH), mp 258–260 °C. The 1H-NMR (D2O, 500 MHz) spectroscopic data were similar to those for 2e.

4. Conclusions

Novel fluorine-containing amino acid enantiomers have been prepared through the hydrolysis of racemic β-amino carboxylic ester hydrochloride salts 3ae catalyzed by lipase PSIM. Excellent enantioselectivities (E > 200) were obtained when the reactions were performed with H2O as a nucleophile in iPr2O at 45 °C, in the presence of Et3N. Both unreacted amino carboxylic esters (R)-4ae and product amino acids (S)-5ae were isolated with excellent ee (usually ≥99%) and good yields (>48%). Suitable analytical methods were devised to follow the enzymatic reactions and calculate the enantiomeric excess, conversions, and enantio-selectivities.

Supplementary Materials

The following are available online, Figures S1–S15: 1H-NMR spectra for (±)-2ae, (±)-3ae, (R)-4ae, and (S)-5ae, Figures S16–S25: GC chromatograms for (S)-5ae and (R)-4ae, Figures S26–S30: 13C-NMR spectra for (±)-3ae, Figures S31–S35: 19F-NMR spectra for (±)-3ae.

Author Contributions

F.F. and E.F. planned and designed the project. S.S. performed the syntheses and characterized the synthesized compounds. E.F. and S.S. prepared the manuscript for publication. All authors discussed the results and commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the Hungarian Scientific Research Council (OTKA, K129049) and the Ministry of National Economy, National Research, Development and Innovation Office (GINOP, 2.3.2-15-2016-00014) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wasserman, H.H.; Berger, G.D. The use of β-lactams in the synthesis of spermine and spermidine alkaloids. Tetrahedron 1983, 39, 2459–2464. [Google Scholar] [CrossRef]
  2. Renault, O.; Guillon, J.; Dallemagne, P.; Rault, S. Efficient synthesis of 2-aryl-6-methyl-2,3-dihydro-1H-pyridin-4-ones. Tetrahedron Lett. 2000, 41, 681–683. [Google Scholar] [CrossRef]
  3. Juaristi, E.; Soloshonok, V.A. Enantioselective Synthesis of β-Amino Acids, 2nd ed.; Wiley: Hoboken, NJ, USA, 2005. [Google Scholar]
  4. Wani, M.C.; Taylor, H.L.; Wall, M.E.; Coggon, P.; McPhail, A.T. Plant antitumore agents. VI. Isolation and structure of taxol, a novel antileukemic and antitumor agent from taxus brevifolia. J. Am. Chem. Soc. 1971, 93, 2325–2327. [Google Scholar] [CrossRef] [PubMed]
  5. Jin, M.; Fischbach, M.A.; Clardy, J. A biosynthetic gene cluster for the acetyl-CoA carboxylase inhibitor andrimid. J. Am. Chem. Soc. 2006, 128, 10660–10661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Cardillo, G.; Gentilucci, L.; Melchiorre, P.; Spampinato, S. Synthesis and binding activity of endomorphin-1-analogues containing β-amino acids. Bioorg. Med. Chem. Lett. 2000, 10, 2755–2758. [Google Scholar] [CrossRef]
  7. Ojima, I. Fluorine in Medicinal Chemistry and Chemical Biology; Wiley-Blackwell: Chichester, UK, 2009. [Google Scholar]
  8. Kirsch, P. Modern Fluoroorganic Chemistry Synthesis, Reactivity, Applications; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  9. Thornberry, N.; Weber, A. Discovery of JANUVIA™ (Sitagiliptin) a selective dipeptidyl peptidase IV inhibitor for the treatment of type2 diabetes. Curr. Top. Med. Chem. 2007, 7, 557–568. [Google Scholar] [CrossRef]
  10. Pepin, J.; Guern, C.; Milord, F.; Schechter, P.J. Difluoromethylornithine for arseno-resistant trypanosome brucei gambiense sleeping sickness. Lancet 1987, 330, 1431–1433. [Google Scholar] [CrossRef]
  11. Wolf, J.E.; Shander, D.; Huber, F.; Jackson, J.; Lin, C.-S.; Mathes, B.M.; Schrode, K. Randomized, double-blind clinical evaluation of the efficacy and safety of topical eflornithine HCl 13.9% cream in the treatment of women with facial hair: Eflornithine treatment for unwanted facial hair. Int. J. Dermatol. 2007, 46, 94–98. [Google Scholar] [CrossRef]
  12. Cimarelli, C.; Palmieri, G.; Volpini, E. An improved synthesis of enantiopure β-amino acids. Synth. Commun. 2001, 31, 2943–2953. [Google Scholar] [CrossRef]
  13. Sivakumar, A.V.; Babu, G.S.; Bhat, S.V. Asymmetric synthesis of β-amino acids through application of chiral sulfoxide. Tetrahedron Asymmetry 2001, 12, 1095–1099. [Google Scholar] [CrossRef]
  14. Wenzel, A.G.; Jacobsen, E.N. Asymmetric catalytic mannich reactions catalyzed by urea derivatives: Enantioselective synthesis of β-aryl-β-amino acids. J. Am. Chem. Soc. 2002, 124, 12964–12965. [Google Scholar] [CrossRef]
  15. Vaidyanathan, R.; Hesmondhalgh, L.; Hu, S. A chemoenzymatic synthesis of an androgen receptor antagonist. Org. Process Res. Dev. 2007, 11, 903–906. [Google Scholar] [CrossRef]
  16. Allwein, S.P.; Roemmele, R.C.; Haley, J.J.; Mowrey, D.R.; Petrillo, D.E.; Reif, J.J.; Gingrich, D.E.; Bakale, R.P. Development and scale-up of an optimized route to the ALK inhibitor CEP-28122. Org. Process Res. Dev. 2012, 16, 148–155. [Google Scholar] [CrossRef]
  17. Forró, E.; Fülöp, F. Recent lipase-catalyzed hydrolytic approaches to pharmacologically important β-and γ-amino acids. Curr. Med. Chem. 2012, 19, 6178–6187. [Google Scholar] [PubMed]
  18. Rodríguez-Mata, M.; García-Urdiales, E.; Gotor-Fernández, V.; Gotor, V. Stereoselective chemoenzymatic preparation of β-amino esters: Molecular modelling considerations in lipase-mediated processes and application to the synthesis of (S)-dapoxetine. Adv. Synth. Catal. 2010, 352, 395–406. [Google Scholar] [CrossRef]
  19. Tasnádi, G.; Forró, E.; Fülöp, F. An efficient new enzymatic method for the preparation of β-aryl-β-amino acid enantiomers. Tetrahedron Asymmetry 2008, 19, 2072–2077. [Google Scholar] [CrossRef]
  20. Tasnádi, G.; Forró, E.; Fülöp, F. Burkholderia cepacia lipase is an excellent enzyme for the enantioselective hydrolysis of β-heteroaryl-β-amino esters. Tetrahedron Asymmetry 2009, 20, 1771–1777. [Google Scholar] [CrossRef]
  21. Tasnádi, G.; Forró, E.; Fülöp, F. Improved enzymatic syntheses of valuable β-arylalkyl-β-amino acid enantiomers. Org. Biomol. Chem. 2010, 8, 793–799. [Google Scholar] [CrossRef]
  22. Forró, E.; Megyesi, R.; Paál, T.A.; Fülöp, F. Efficient dynamic kinetic resolution method for the synthesis of enantiopure 6-hydroxy- and 6-methoxy-1,2,3,4-tetrahydroisoquinoline-1-carboxylic acid. Tetrahedron Asymmetry 2016, 27, 1213–1216. [Google Scholar] [CrossRef]
  23. Megyesi, R.; Mándi, A.; Kurtán, T.; Forró, E.; Fülöp, F. Dynamic kinetic resolution of ethyl 1,2,3,4-tetrahydro-β-carboline-1-carboxylate: Use of different hydrolases for stereocomplementary processes. Eur. J. Org. Chem. 2017, 32, 4713–4718. [Google Scholar] [CrossRef]
  24. Paál, T.A.; Forró, E.; Fülöp, F.; Liljeblad, A.; Kanerva, L.T. Lipase-catalyzed kinetic resolution of 1,2,3,4-tetrahydroisoquinoline-1-acetic acid esters. Tetrahedron Asymmetry 2008, 19, 2784–2788. [Google Scholar] [CrossRef]
  25. Paál, T.A.; Forró, E.; Liljeblad, A.; Kanerva, L.T.; Fülöp, F. Lipase-Catalyzed kinetic and dynamic kinetic resolution of 1,2,3,4-tetrahydroisoquinoline-1-carboxylic acid. Tetrahedron Asymmetry 2007, 18, 1428–1433. [Google Scholar] [CrossRef]
  26. Rangel, H.; Carrillo-Morales, M.; Galindo, J.M.; Castillo, E.; Obregón-Zúñiga, A.; Juaristi, E.; Escalante, J. Structural features of N-benzylated-β-amino acid methyl esters essential for enantiodifferentiation by lipase B from candida antarctica in hydrolytic reactions. Tetrahedron Asymmetry 2015, 26, 325–332. [Google Scholar] [CrossRef]
  27. Pérez-Venegas, M.; Reyes-Rangel, G.; Neri, A.; Escalante, J.; Juaristi, E. Mechanochemical enzymatic resolution of N-benzylated-β3-amino esters. Beilstein J. Org. Chem. 2017, 13, 1728–1734. [Google Scholar] [CrossRef] [Green Version]
  28. Nagy, B.; Galla, Z.; Bencze, L.C.; Toșa, M.I.; Paizs, C.; Forró, E.; Fülöp, F. Covalently immobilized lipases are efficient stereoselective catalysts for the kinetic resolution of rac-(5-phenylfuran-2-yl)-β-alanine ethyl ester hydrochlorides. Eur. J. Org. Chem. 2017, 20, 2878–2882. [Google Scholar] [CrossRef]
  29. Patel, R.N. Green Biocatalysis; Wiley: Hoboken, NJ, USA, 2016. [Google Scholar]
  30. Zhang, X.-X.; Gao, Y.; Hu, X.-S.; Ji, C.-B.; Liu, Y.-L.; Yu, J.-S. Recent advances in catalytic enantioselective synthesis of fluorinated α- and β-amino acids. Adv. Synth. Catal. 2020, 362, 4763–4793. [Google Scholar] [CrossRef]
  31. Zablocki, J.A.; Tjoeng, F.S.; Bovy, P.R.; Miyano, M.; Garland, R.B.; Williams, K.; Schretzman, L.; Zupec, M.E.; Rico, J.G.; Lindmark, R.J.; et al. A novel series of orally active antiplatelet agents. Bioorg. Med. Chem. 1995, 3, 539–551. [Google Scholar] [CrossRef]
  32. Johnson, T.B.; Livak, J.E. Researches on pyrimidines. CXLIX. The synthesis of aryl substituted dihydrouracils and their conversion to uracil derivatives. J. Am. Chem. Soc. 1936, 58, 299–303. [Google Scholar] [CrossRef]
  33. Forró, E.; Fülöp, F. New enzymatic two-step cascade reaction for the preparation of a key intermediate for the taxol side-chain. Eur. J. Org. Chem. 2010, 16, 3074–3079. [Google Scholar] [CrossRef]
  34. Forró, E. New gas chromatographic method for the enantioseparation of β-amino acids by a rapid double derivatization technique. J. Chromatogr. A 2009, 1216, 1025–1029. [Google Scholar] [CrossRef]
  35. Straathof, A.J.J.; Rekels, J.L.L.; Heijnen, J.J. Mass balancing in kinetic resolution: Calculating yield and enantiomeric excess using chiral balance. Biotechnol. Bioeng. 1995, 45, 536–538. [Google Scholar] [CrossRef] [PubMed]
  36. Davies, S.G.; Fletcher, A.M.; Lv, L.; Roberts, P.M.; Thomson, J.E. Asymmetric synthesis of β-fluoroaryl-β-amino acids. Tetrahedron Asymmetry 2012, 23, 910–925. [Google Scholar] [CrossRef]
  37. Forró, E.; Paál, T.; Tasnádi, G.; Fülöp, F. A new route to enantiopure β-aryl-substituted β-amino acids and 4-aryl-substituted β-lactams through lipase-catalyzed enantioselective ring cleavage of β-lactams. Adv. Synth. Catal. 2006, 348, 917–923. [Google Scholar] [CrossRef]
  38. Bull, S.D.; Davies, S.G.; Delgado-Ballester, S.; Kelly, P.M.; Kotchie, L.J.; Gianotti, M.; Laderas, M.; Smith, A.D. Asymmetric synthesis of β-haloaryl-β-amino acid derivatives. J. Chem. Soc. Perkin Trans 1 2001, 23, 3112–3121. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of (±)-3ae.
Scheme 1. Synthesis of (±)-3ae.
Molecules 25 05990 sch001
Scheme 2. Enzymatic kinetic resolution of (±)-3ae through a hydrolytic procedure.
Scheme 2. Enzymatic kinetic resolution of (±)-3ae through a hydrolytic procedure.
Molecules 25 05990 sch002
Table 1. Enzyme screening in the hydrolysis of (±)-3a a.
Table 1. Enzyme screening in the hydrolysis of (±)-3a a.
EntryEnzymeees (%) beep (%) cConv. (%) dEe
1Lipase PSIM889548108
2Lipase AY29181
3Lipase AK1875198
4PPL22952
5CAL-B25301
a 0.025 M substrate, 30 mg mL–1 lipase, 1 mL iPr2O, 5 equiv. Et3N, 0.5 equiv. H2O, at 45 °C after 10 min. b according to GC after derivatization. c according to GC after double derivatization [34]. d c = ees/(ees + eep) [35]. e E = {ln [(1 − c) × (1 − eep)]/ln [(1 − c) × (1 + eep)]}.
Table 2. Solvent screening in the hydrolysis of (±)-3a a.
Table 2. Solvent screening in the hydrolysis of (±)-3a a.
EntrySolvent (1 mL)ees (%) beep (%) cConv. (%) dEe
1TBME95885259
22-Me-THF979351113
3EtOAc652113
4Propylene carbonate92795427
5no solvent90924974
a 0.025 M substrate, 30 mg mL–1 lipase PSIM, 5 equiv. Et3N, 0.5 equiv. H2O, at 45 °C after 10 min. b according to GC after derivatization. c according to GC after double derivatization [34]. d c = ees/(ees + eep) [35]. e E = {ln[(1 − c) × (1 − eep)]/ln[(1 − c) × (1 + eep)]}.
Table 3. Effect of enzyme concentration in the hydrolysis of (±)-3a a.
Table 3. Effect of enzyme concentration in the hydrolysis of (±)-3a a.
EntryLipase PSIM (mg mL–1)ees (%) beep (%) cConv. (%) dEe
110979750>200
25959849>200
3285924863
a 0.025 M substrate, 1 mL iPr2O, 5 equiv. Et3N, 0.5 equiv. H2O, at 3 °C after 10 min. b according to GC after derivatization. c according to GC after double derivatization [34]. d c = ees/(ees + eep) [35]. e E = {ln[(1 − c) × (1 − eep)]/ln[(1 − c) × (1 + eep)]}.
Table 4. Effect of enzyme concentration at 45 °C in the hydrolysis of (±)-3a a.
Table 4. Effect of enzyme concentration at 45 °C in the hydrolysis of (±)-3a a.
EntryEnzyme Conc. (mg mL–1)ees (%) beep (%) cConv. (%) dEe
12481510
2515851514
31021862017
42046923338
54097895274
a 0.025 M substrate, lipase PSIM, 5 equiv. Et3N, 0.5 equiv. H2O, after 10 min. b according to GC after derivatization. c according to GC after double derivatization [34]. d c = ees/(ees + eep) [35]. e E = {ln[(1 − c) × (1 − eep)]/ln[(1 − c) × (1 + eep)]}.
Table 5. Lipase PSIM-catalyzed hydrolysis of (±)-3ae a.
Table 5. Lipase PSIM-catalyzed hydrolysis of (±)-3ae a.
SubstrateRt (h)Conv. (%)Eβ-Amino Acid (5a–e)β-Amino Ester (4a–e)
Yield (%)Isomereeb (%) [ α ] D 25 (H2O)Yield (%)Isomereec (%) [ α ] D 25 (CHCl3)
3a850>20048(S)>99–3.1 d49(R)97+17.9 e
3b7249>20048(S)>99–5 f38(R)94+9 g
3c1850>20049(S)>99–3 h49(R)>99+18.9 i
3d2649>20049(S)>99–11 j48(R)>99+20.3 k
3e2350>20048(S)>99–13 l47(R)>99+16 m
a 30 mg mL–1 enzyme in iPr2O, 5 equiv. Et3N, 0.5 equiv. H2O, at 45 °C. b according to GC after derivatization. c according to GC after double derivatization [34]. d c = 0.28. e c = 0.44. f c = 0.26. g c = 0.29. h c = 0.28. i c = 0.41. j c = 0.19 (MeOH). k c = 0.53. l c = 0.21 (MeOH). m c = 0.13.
Sample Availability: Samples of the compounds (±)-2ae, (±)-3ae, (R)-4ae and (S)-5ae are not available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shahmohammadi, S.; Fülöp, F.; Forró, E. Efficient Synthesis of New Fluorinated β-Amino Acid Enantiomers through Lipase-Catalyzed Hydrolysis. Molecules 2020, 25, 5990. https://doi.org/10.3390/molecules25245990

AMA Style

Shahmohammadi S, Fülöp F, Forró E. Efficient Synthesis of New Fluorinated β-Amino Acid Enantiomers through Lipase-Catalyzed Hydrolysis. Molecules. 2020; 25(24):5990. https://doi.org/10.3390/molecules25245990

Chicago/Turabian Style

Shahmohammadi, Sayeh, Ferenc Fülöp, and Enikő Forró. 2020. "Efficient Synthesis of New Fluorinated β-Amino Acid Enantiomers through Lipase-Catalyzed Hydrolysis" Molecules 25, no. 24: 5990. https://doi.org/10.3390/molecules25245990

Article Metrics

Back to TopTop