Tandem Alkylation–Second-Order Asymmetric Transformation

Aug 22, 2018 - ... for the Preparation of Phenylalanine-Type Tailor-Made α-Amino Acids ... open access article published under an ACS AuthorChoice Li...
0 downloads 0 Views 937KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 9729−9737

http://pubs.acs.org/journal/acsodf

Tandem Alkylation−Second-Order Asymmetric Transformation Protocol for the Preparation of Phenylalanine-Type Tailor-Made α‑Amino Acids Ryosuke Takeda,*,†,‡ Aki Kawashima,† Junya Yamamoto,† Tatsunori Sato,† Hiroki Moriwaki,† Kunisuke Izawa,† Hidenori Abe,† and Vadim A. Soloshonok*,‡,§ †

Hamari Chemicals Ltd., 1-4-29 Kunijima, Higashi-Yodogawa-ku, Osaka 533-0024, Japan Department of Organic Chemistry I, Faculty of Chemistry, University of the Basque Country (UPV/EHU), Paseo Manuel Lardizábal 3, 20018 San Sebastián, Spain § IKERBASQUE, Basque Foundation for Science, María Díaz de Haro 3, Plaza Bizkaia, 48013 Bilbao, Spain ACS Omega 2018.3:9729-9737. Downloaded from pubs.acs.org by 185.46.87.169 on 08/24/18. For personal use only.



S Supporting Information *

ABSTRACT: In this work, we disclose an advanced general process for the synthesis of tailor-made α-amino acids (α-AAs) via tandem alkylation−second-order asymmetric transformation. The first step is the alkylation of the chiral Ni(II) complex of glycine Schiff base, which is conducted under mild phasetransfer conditions allowing the structural construction of target α-AAs. The second step is based on the methodologically rare second-order asymmetric transformation, resulting in nearly complete precipitation of the corresponding (SC,RN,RC)configured diastereomer, which can be collected by a simple filtration. The operational convenience and potential scalability of all experimental procedures, coupled with excellent stereochemical outcome, render this method of high synthetic value for the preparation of various tailor-made α-AAs.



INTRODUCTION Amino acids (AAs) are among a handful of paramount molecules involved in virtually all aspects of the phenomenon of life. Quite naturally, they played a significant role in organic, bioorganic, medicinal, and pharmaceutical chemistry from the earliest days of modern healthcare and life sciences.1 In particular, the use of tailor-made2 AAs in the design of peptides/peptidomimetics with restricted number of conformations and precise positioning of the side-chains in the χ-space3 is a well-established paradigm in the modern pharmaceutical industry.1,4,5 Quite interestingly, one of the emerging areas of tailor-made AA application is the biological containment strategy of the produced genetically modified organisms.6 Accordingly, the current interest in the asymmetric synthesis of tailor-made AAs7 and their analogues, such as amino sulfonic,8 phosphonic,9 and fluorine-containing10 acids, is at an all-time high, targeting novel structural motifs, functions, and properties. Consistent with our continuous interest in asymmetric synthesis of various tailor-made AAs, for example, fluorine-11 and phosphorus-containing,12 sterically constrained,13 and their nonlinear chiroptical properties, such as self-disproportionation of enantiomers,14 we were actively developing the Ni(II) complex chemistry of AA Schiff bases4d,7d,e as a general approach for the preparation of tailor-made AAs (Scheme 1). © 2018 American Chemical Society

As shown in Scheme 1, trichloro-substituted ligand (S)- or (R)-1, currently the most frequently used source of recyclable chirality in this method,15 is reacted with glycine and a source of Ni(II) ions to produce the Schiff base complex 2.16 The square-planar Ni(II) complex 2 serves as a versatile chiral nucleophilic glycine equivalent to afford products 3 by the reactions with various electrophilic reagents. Most commonly used reaction types include, for example, alkyl halide alkylations,17 Michael,18 aldol,19 and Mannich20 addition reactions. Intermediate products 3 can be conveniently disassembled to release target AAs 4−10 and to recycle chiral tridentate ligand 1. This method can be used for the preparation of general type α-AAs 4, β-functionalized/ substituted derivatives 5, sym-6,21 or chiral quaternary 722 αAAs, in particular, (1R,2S)-1-amino-2-vinylcyclopropane-1carboxylic acid 823 and bis-α-AAs of type 9.24 Moreover, the major advantage of this method over other well-developed general approaches7b,c,25 is its exceptionally successful application for kinetic resolution of unprotected α- and βAAs of types 426 and 10,27 respectively. This latter application was shown to be competitive with traditionally economically Received: June 22, 2018 Accepted: August 8, 2018 Published: August 22, 2018 9729

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

advancement of this methodology, we now report a tandem alkylation−SOAT protocol, which includes synthesis of glycine Schiff base Ni(II) complex 14 and its alkylation under phasetransfer catalysis (PTC) conditions followed by SOAT via equilibration/precipitation of major products 15.

Scheme 1. General Approach for the Asymmetric Synthesis of AAs via Ni(II) Complexes of Schiff Bases



RESULTS AND DISCUSSION Unlike the asymmetric synthesis, where the major product is a result of preferential stereochemical interactions, which can be comprehended, predicted, and rationally designed, the SOAT is essentially a serendipitous approach resulting from rather complex and multifactorial interplay of numerous chemical/ steric/physical properties.33 Nevertheless, success in developing a SOAT process using ligand 11 was predicated on the idea of “extreme” lipophilic properties32 of the corresponding rimantadine-containing Ni(II) complexes.31 Drawing from the previous results, we decided to extend the synthetic versatility of this approach by preparing the required diastereomeric complexes not via the reactions of ligand 11 with racemates of the target α-AAs but via homologation of the intermediate glycine derivative. To this end, we focused on the synthesis of required glycine complexes 14 (Scheme 3). After some extensive experimentation, we found that the application of Na2CO3 as a base, in combination with catalytic amounts of n-Bu4NI, provides for the most practical preparation of the target glycine complexes (SC,RN)-14 and (SC,SN)-14, in about 3:1 ratio. Under these conditions, using only 10% excess of glycine and the Ni(II) source allowed for the complete conversion of the starting ligand (S)-11 to products 14, isolated with quantitative chemical yield. Clean preparation of glycine complexes 14 was very important part of the project allowing us to avoid any unnecessary purification of the intermediate derivatives. Having achieved virtually byproduct-less synthesis of glycine complexes 14, we focused on the next alkylation step. On the basis of the previous research,4d we selected PTC conditions34 allowing for very mild homologation of the glycine moiety with usually excellent chemical yields. The results obtained are summarized in Scheme 4. As a model experiment, we selected the alkylation reaction using m-bromobenzyl bromide 16 (Scheme 4). In this case, the expected product 15a was previously obtained using the reaction of racemic AA with ligand (S)-11,31 thus facilitating

attractive enzymatic approaches.28 Most recently, using the modular approach29 for the design of chiral tridentate ligands, we developed a series of ligands featuring free N−H functionality and multiple stereogenic centers.30 In particular, the adamantyl-containing derivative 11 (Scheme 2) was found to possess some remarkable physicochemical properties for the development of second-order asymmetric transformation (SOAT) approach for the general synthesis of α-AAs.31 As presented in Scheme 2, in the previous work, we demonstrated that the newly designed ligand 11 readily reacts with racemic α-AAs 12 to produce a mixture of four theoretically possible diastereomeric products.31 However, because of the extreme lipophilicity of the adamantyl moiety,32 one of the diastereomers, (SC,RN,RC)-configured 13, precipitates from the methanol or aqueous methanol solution, thus driving the equilibrium of all other diastereomers to the one major reaction product. As the next logical step in the

Scheme 2. Applications of (S)-Rimantadine-Derived Ligand 11; Previous Work on SOAT of Racemic α-AAs and This Work Dealing with Tandem Alkylation−SOAT, for General Asymmetric Synthesis of α-AAs

9730

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

Scheme 3. Optimized Conditions for the Preparation of the Requisite Ni(II) Complexes of Glycine (SC,RN)-14 and (SC,SN)-14

attempts to use this tandem procedure for the preparation of α-AA-containing alkyl side-chains were unsuccessful. The major problem we encountered was low reactivity of alkyl halides under the PTC conditions. Diastereomerically pure Ni(II) complexes 15 can be readily disassembled to release the target AAs or their protected derivatives.4d For example, diastereomer (SC,RN,RC)-15a was treated with 3 N HCl in methanol to produce m-bromophenylalanine (R)-20 and chiral ligand (S)-11 (Scheme 5). The latter, configurationally uncompromised, was almost quantitatively recovered and reused for continuous synthesis of other AAs. AA (R)-20 was not isolated in analytically pure form but transformed in situ to the corresponding Fmoc derivative (R)21. The standard protection protocol with application of N-(9fluorenylmethoxycarbonyloxy)succinimide (Fmoc-OSu) was used, with an exception that ethylenediaminetetraacetic acid (EDTA) was added to chelate the Ni(II) ions. In another example, presented in Scheme 6, we demonstrate the isolation of free p-nitro-phenylalanine (R)-23. Thus, after the disassembly of diastereomerically pure Ni(II) complex (SC,RN,RC)-15f in methanol under the action of 3 N HCl, AA (R)-23 was isolated by precipitation at the isoelectric point (pH 5.5) in the presence of EDTA to chelate the Ni(II) ions. For the further use of this tailor-made AA in our peptiderelated projects, we also prepared Fmoc-protected derivative (R)-24, similar to the preparation of m-bromo-phenylalanine (R)-21 (Scheme 6).

the identification and stereochemical assignments. Starting glycine complexes 14 were used as obtained, without any purification, as a 3:1 mixture of (SC,RN)-14 and (SC,SN)-14 diastereomers. We found that using 1,2-dichloroethane as an organic solvent, 30% aqueous NaOH as a base, and 25 mol % of n-Bu4NI as a phase-transfer catalyst, virtually complete consumption of starting glycine complexes 14 can be achieved in about 10 min of the reaction time. Analysis of the reaction mixture by thin-layer chromatography (TLC) showed expected alkylation products 15a as a major (>50%) product and 17a−19a as minor products. The stereogenic nitrogen and AA α-carbon atoms in compounds 15a and 17a−19a are configurationally unstable. Therefore, under the basic conditions, all four diastereomers undergo rapid epimerization and interconversion via stepwise inversion of either nitrogen or αcarbon or both atoms. Nonetheless, diastereomeric products 15a and 17a−19a are compounds of different chemical properties, in particular, the solubility in a given solvent, providing the opportunity for a SOAT process. Indeed, the major diastereomer (SC,RN,RC)-15a was found to gradually precipitate from the reaction mixture, leading to a progressive transformation of more soluble 17a−19a to the crystalline precipitate of compound 15a. Thus, after about 1 h of stirring at room temperature, the reaction mixture was filtered affording diastereomerically pure (SC,RN,RC)-15a with 92.4% yield. This tandem alkylation−SOAT protocol was found to be of general application for the preparation of various phenylalanine-type tailor-made α-AAs. For example, the derivatives of p-bromo- (15b), p-chloro- (15c), p-trifluoromethyl- (15d), piodo- (15e), and p-nitro-substituted (15f) phenylalanines were readily prepared in diastereomerically pure form under the same conditions using filtration as the final isolation procedure. Of particular interest is the preparation of naphthylalanine derivative (SC,RN,RC)-15g, which was isolated with respectable chemical yield. In the case of compounds 15c,e, and f, the application of the standard conditions gave relatively lower isolated yields of ∼50−56%. In this regard, we would like to emphasize that relatively modest isolated yields are the result of incomplete precipitation of the major diastereomer from the reaction mixture. Furthermore, it should be noted that under the SOAT conditions, the isolated yield depends on a physicochemical property which can be conditioned by optimizing the solvent system for each particular case.33 For instance, a gradual addition of water to the reaction mixture leads to a lower solubility of products allowing increasing precipitation of the target products. This possibility was demonstrated on the example of product 15f, yield of which was improved from 56.5 to 86.5% by adding about 20% of water to the reaction mixture (see Supporting Information). It should be noted, however, that the diastereomeric purity of thus prepared complex 15f was 93.2% de. On the other hand,



CONCLUSIONS We demonstrate that Ni(II) complexes (SC,RN)- and (SC,SN)14, derived from glycine Schiff base and (S)-rimantadine, can be regioselectivity alkylated on the glycine residue under the PTC conditions affording a mixture of up to four diastereomeric products. Without purification, the alkylated products can be submitted to the SOAT conditions allowing a transformation of the diastereomeric mixture to single precipitated stereoisomer 15 of (SC,RN,RC) absolute configuration. The operational convenience and chromatographyfree experimental procedures render this method of potentially high synthetic value for the preparation of phenylalanine-type tailor-made α-AAs.



EXPERIMENTAL SECTION General Information. All reagents and solvents were used as received. The reaction mixture was magnetically stirred and monitored with the aid of TLC on precoated silica gel plates, and visualization was carried out using UV light and ninhydrin. Flash chromatography was performed with the indicated solvents on silica gel (particle size 0.040−0.063 mm). Yields reported are for isolated, spectroscopically pure compounds. 9731

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

Scheme 4. Alkylation of Ni(II) Complexes of Glycine (SC,RN)-14 and (SC,SN)-14 under PTC Conditions; Equilibration of Diastereomeric Complexes 17−19 and 15 and Precipitation of the Final Major Product (SC,RN,RC)-15

1

H and 13C spectra were recorded on a 300 MHz Brüker instrument. Chemical shifts are given in ppm (δ), referenced to the residual proton resonances of the solvents. Coupling constants (J) are given in Hz. The letters s, d, t, q, m, and br stand for singlet, doublet, triplet, quartet, multiplet, and broad, respectively. High-resolution mass spectra (HRMS) were recorded using an ultraperformance liquid chromatography/ quadrupole time-of-flight mass spectrometry system in the ESI mode.

Procedure for the Preparation of Ni(II) Complexes by Reaction of Ligand (S)-11 with Glycine. To a mixture of ligand (S)-11 (1.0 g, 2.22 mmol), nickel(II) nitrate hexahydrate (0.71 g, 2.44 mmol), glycine (0.18 g, 2.44 mmol), tetrabutylammonium iodide (0.08 g, 10 mol %), and methanol (40 mL) was added sodium carbonate (1.18 g, 11.1 mmol), and the reaction mixture was stirred at reflux for 2 h. After the ligand (S)-11 was consumed, the reaction was quenched by pouring icy 5% aqueous acetic acid (200 mL) to give a precipitate. The precipitate was filtrated, washed with 9732

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

Ni(II)/(S)-11/(R)-2-Amino-3-(4-bromophenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15b. 1 Yield: 79.3%. [α]25 D −2511 (c 0.050, CHCl3). H NMR (300 MHz, CDCl3): δ 1.38 (d, J = 7.1 Hz, 3H), 1.34−1.46 (m, 6H), 1.58−1.80 (m, 6H), 2.04 (s, 3H), 2.39 (br, 1H), 2.48 (q, J = 7.1 Hz, 1H), 2.73 (dd, J = 5.1, 13.5 Hz, 1H), 3.07 (dd, J = 2.5, 13.5 Hz, 1H), 3.12−3.20 (m, 2H), 4.34 (dd, J = 3.0, 4.0 Hz, 1H), 6.76 (d, J = 2.1 Hz, 1H), 7.08−7.15 (m, 1H), 7.26−7.34 (m, 2H), 7.44 (d, J = 8.1 Hz, 2H), 7.57−7.67 (m, 3H), 7.78 (d, J = 8.1 Hz, 2H), 8.49 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 10.2, 28.5, 36.8, 36.9, 39.1, 39.8, 50.9, 63.0, 72.9, 122.1, 125.7, 126.2, 127.7, 127.7, 128.7, 129.8, 130.0, 130.9, 132.0, 132.7, 132.9, 133.3, 133.9, 136.0, 142.2, 169.8, 177.8, 179.6. HRMS: calcd for C36H38BrClN3NiO3 [M + H]+, 732.1139; found, 732.1152. Ni(II)/(S)-11/(R)-2-Amino-3-(4-chlorophenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15c. 1 Yield: 50.8%. [α]25 D −2886 (c 0.045, CHCl3). H NMR (300 MHz, CDCl3): δ 1.28 (d, J = 7.1 Hz, 3H), 1.23−1.35 (m, 6H), 1.46−1.69 (m, 6H), 1.93 (s, 3H), 2.26−2.34 (m, 1H), 2.39 (q, J = 7.1 Hz, 1H), 2.64 (dd, J = 5.1, 13.5 Hz, 1H), 2.94−3.07 (m, 3H), 4.23 (dd, J = 3.2, 5.0 Hz, 1H), 6.66 (d, J = 2.5 Hz, 1H), 6.98−7.03 (m, 1H), 7.16−7.23 (m, 2H), 7.37 (d, J = 8.4 Hz, 2H), 7.48−7.55 (m, 5H), 8.38 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 9.8, 28.0, 36.3, 36.5, 38.7, 39.3, 50.4, 62.7, 72.4, 125.2, 125.8, 127.2, 127.3, 128.2, 128.6, 129.4, 129.6, 130.4, 132.3, 132.5, 132.9, 133.1, 133.6, 135.0, 141.7, 169.4, 177.5, 179.1. HRMS: calcd for C36H38Cl2N3NiO3 [M + H]+, 688.1644; found, 688.1639. Ni(II)/(S)-11/(R)-2-Amino-3-(4-trifluorophenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15d. 1 Yield: 73.4%. [α]25 D −2019 (c 0.054, CHCl3). H NMR (300 MHz, CDCl3): δ 1.27 (d, J = 7.1 Hz, 3H), 1.16−1.33 (m, 6H), 1.42−1.68 (m, 6H), 1.91 (s, 3H), 2.25 (d, J = 6.2 Hz, 1H), 2.44 (q, J = 7.0 Hz, 1H), 2.75 (dd, J = 5.1, 13.5 Hz, 1H), 2.85− 3.11 (m, 3H), 4.21−4.31 (m, 1H), 6.67 (d, J = 2.3 Hz, 1H), 6.96−7.05 (m, 1H), 7.14−7.27 (m, 2H), 7.44−7.58 (m, 5H), 7.78 (d, J = 7.9 Hz, 2H), 8.42 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 9.6, 28.0, 36.3, 36.4, 38.6, 39.8, 50.4, 63.0, 72.1, 124.2 (q, J = 272.4 Hz), 125.1, 125.3 (q, J = 3.7 Hz), 125.8, 127.3, 127.4, 128.2, 129.4, 129.6, 130.1 (q, J = 32.3 Hz), 130.5, 131.9, 132.4, 132.7, 132.9, 140.7 (q, J = 1.4 Hz), 141.8, 169.8, 177.5, 179.0. HRMS: calcd for C37H38ClF3N3NiO3 [M + H]+, 722.1907; found, 722.1897.

Scheme 5. Disassembly of Diastereomerically Pure (SC,RN,RC)-15a; Isolation of Target Fmoc-Derivative of AA (R)-21 and Recycling of Chiral Ligand (S)-11

5% aqueous acetic acid, and dried in vacuo at 50 °C overnight to afford the diastereomixture of Ni(II) complex (SC,RN)-14 and (SC,SN)-14. A 3:1 ratio mixture of two diastereomers was obtained. The spectral data were reported in our previous literature. General Procedure for the Preparation of Ni(II) Complexes via Tandem Alkylation of Ni(II) Complexes of Glycine (SC,RN)-14 and (SC,SN)-14. To a solution of Ni(II) complexes of glycine (SC,RN)-14 and (SC,SN)-14 (1 equiv) in 1,2-dichloroethane (8 volumes), 30% aqueous sodium hydroxide solution (40 equiv) was added followed by tetrabutylammonium iodide (25 mol %) and the corresponding alkyl halide (1.1 equiv). After stirring at room temperature (rt) for 10−30 min, the reaction mixture was diluted with water and extracted with dichloromethane; then, the organic layer was evaporated to dryness. The obtained residue and potassium carbonate (4 equiv) were stirred in methanol (40 volumes) at reflux for 1 h. After the reaction, several amounts of water were dropped to the mixture to give a precipitate. The precipitate was filtrated, washed with aqueous methanol, and dried in vacuo at 60 °C overnight to afford the corresponding Ni(II) complex. Ni(II)/(S)-11/(R)-2-Amino-3-(3-bromophenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15a. Yield: 92.4%. The spectral data were reported in our previous literature.

Scheme 6. Disassembly of Diastereomerically Pure (SC,RN,RC)-15f; Isolation of Target-Free AA (R)-23 and Its Fmoc Derivative (R)-24, as Well as the Recycling of Chiral Ligand (S)-11

9733

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

recrystallized from THF/IPE (1:1) and dried at 50 °C to afford (R)-21 (466 mg, 99.1%, 97.1% ee). The spectral data were found to be identical with the commercial enantiopure sample. 1H NMR (200 MHz, DMSO-d6): δ 2.85 (dd, J = 10.4, 13.7 Hz, 1H), 3.10 (dd, J = 4.4, 13.7 Hz, 1H), 4.02−4.33 (m, 4H), 7.14−7.56 (m, 8H), 7.57−7.77 (m, 3H), 7.88 (d, J = 7.3 Hz, 2H). 13C NMR (50 MHz, DMSO-d6): δ 36.1, 46.6, 55.4, 65.6, 120.1, 121.4, 125.1, 125.2, 127.0, 127.6, 128.2, 129.2, 130.2, 131.9, 140.6, 140.6, 141.0, 143.7, 143.7, 155.8, 172.9. Synthesis of Fmoc-(R)-p-nitro-phenylalanine, (R)-24. To a suspension of (SC,RN,RC)-15f (5 g, 7.14 mmol) in methanol (60 mL) was added 3 N HCl (11.9 mL, 35.7 mmol), and the reaction mixture was heated at 50 °C for 1 h. Upon disappearance of the red color of the starting complex, the reaction mixture was concentrated to give a residue. To the residue were added water (25 mL) and ethyl acetate (50 mL) and then the phases were separated. The organic phase was washed with brine (10 mL), dried over sodium sulfate, and evaporated under vacuum to afford the recycle ligand (S)-4 (2.93 g, 84.1%). To the aqueous phase was added EDTA disodium salt dihydrate (2.66 g, 7.14 mmol), and the reaction mixture was basified with 48% sodium hydroxide to pH 13. After stirring at rt for 10 min, the solution was acidified with 12 N HCl to pH 5.5 to give a precipitate. The precipitate was filtered and washed with water (20 mL) to afford AA (R)-23 (1.30 g, 86.6%). The AA (R)-23 thus obtained was dissolved in water (50 mL) and THF (25 mL), and the reaction mixture was basified with sodium carbonate (1.51 g, 14.3 mmol) and sodium bicarbonate (1.20 g, 14.3 mmol). To this solution was added a solution of Fmoc-OSu (2.41 g, 7.14 mmol) in THF (25 mL) and stirred at rt for 16 h. After evaporation of the organic solvents, the aqueous residue was acidified with 3 N HCl to pH 2 to give a precipitate. The precipitate was filtered and washed with water (50 mL) to afford crude (R)-24. The crude solid thus obtained was recrystallized from THF/IPE (1:1) and dried at 50 °C to afford (R)-24 (2.59 g, 97.1%, 98.6% ee). The spectral data were found to be identical with the commercial enantiopure sample. 1H NMR (200 MHz, DMSO-d6): δ 3.00 (dd, J = 10.9, 13.7 Hz, 1H), 3.24 (dd, J = 4.4, 13.7 Hz, 1H), 4.02−4.37 (m, 4H), 7.28 (dddd, J = 1.0, 3.2, 7.4, 7.4 Hz, 2H), 7.40 (ddd, J = 1.0, 7.4, 7.4 Hz, 2H), 7.54 (d, J = 8.6 Hz, 2H), 7.61 (d, J = 7.4 Hz, 2H), 7.77 (d, J = 8.5 Hz, 1H), 7.88 (d, J = 7.4 Hz, 2H), 8.14 (d, J = 8.6 Hz, 2H). 13C NMR (50 MHz, DMSO-d6): δ 36.2, 46.6, 54.9, 65.6, 120.2, 123.3, 125.2, 127.1, 127.7, 130.5, 140.7, 140.7, 143.7, 143.8, 146.3, 146.5, 156.0, 172.9.

Ni(II)/(S)-11/(R)-2-Amino-3-(4-iodephenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15e. Yield: 53.5%. 1 [α]25 D −2558 (c 0.050, CHCl3). H NMR (300 MHz, CDCl3): δ 1.28 (d, J = 7.3 Hz, 3H), 1.25−1.36 (m, 6H), 1.48−1.69 (m, 6H), 1.94 (s, 3H), 2.21−2.30 (m, 1H), 2.35 (q, J = 7.1 Hz, 1H), 2.60 (dd, J = 4.9, 13.5 Hz, 1H), 2.94 (dd, J = 3.3, 13.5 Hz, 1H), 3.00−3.15 (m, 2H), 4.23 (dd, J = 3.3, 4.9 Hz, 1H), 6.65 (d, J = 2.5 Hz, 1H), 6.96−7.05 (m, 1H), 7.15−7.26 (m, 4H), 7.46−7.55 (m, 3H), 7.88 (d, J = 8.2 Hz, 2H), 8.39 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 9.8, 28.0, 36.3, 36.4, 38.6, 39.4, 50.6, 62.5, 72.5, 93.0, 125.2, 125.7, 127.2, 127.2, 128.2, 129.3, 129.5, 130.4, 132.2, 132.4, 132.8, 133.8, 136.2, 137.5, 141.7, 169.3, 177.3, 179.2. HRMS: calcd for C36H38ClIN3NiO3 [M + H]+, 780.1000; found, 780.0981. Ni(II)/(S)-11/(R)-2-Amino-3-(4-nitrophenyl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15f. Yield: 56.5%. 1 [α]25 D −2237 (c 0.058, CHCl3). H NMR (300 MHz, CDCl3): δ 1.31 (d, J = 7.0 Hz, 3H), 1.20−1.39 (m, 6H), 1.42−1.69 (m, 6H), 1.92 (s, 3H), 2.49−2.66 (m, 2H), 2.88−3.15 (m, 4H), 4.16−4.28 (m, 1H), 6.68 (s, 1H), 6.99 (d, J = 6.7 Hz, 1H), 7.16−7.28 (m, 2H), 7.34 (d, J = 8.1 Hz, 2H), 7.46−7.62 (m, 3H), 8.23 (d, J = 8.1 Hz, 2H), 8.45 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 10.0, 28.4, 36.8, 39.2, 40.5, 51.0, 63.8, 71.9, 123.8, 125.4, 126.3, 127.7, 127.8, 128.3, 129.9, 130.1, 131.0, 132.1, 132.8, 133.3, 142.2, 144.2, 148.1, 170.6, 177.8, 179.2. HRMS: calcd for C36H38ClN4NiO5 [M + H]+, 699.1884; found, 699.1882. Ni(II)/(S)-11/(R)-2-Amino-3-(naphthalen-2-yl)propanoic Acid Schiff Base Complex (SC,RN,RC)-15g. 1 Yield: 88.1%. [α]25 D −2412 (c 0.046, CHCl3). H NMR (300 MHz, CDCl3): δ 0.96 (s, 6H), 1.10 (d, J = 7.1 Hz, 3H), 1.34− 1.63 (m, 8H), 1.82 (s, 3H), 2.16 (q, J = 7.1 Hz, 1H), 2.45 (d, J = 15.5 Hz, 1H), 2.80 (dd, J = 5.3, 13.5 Hz, 1H), 3.20 (dd, J = 3.0, 13.5 Hz, 1H), 4.31 (dd, J = 2.9, 5.2 Hz, 1H), 6.68 (d, J = 2.5 Hz, 1H), 7.10−7.26 (m, 4H), 7.43−7.61 (m, 5H), 7.86− 8.10 (m, 4H), 8.33 (d, J = 9.2 Hz, 1H). 13C NMR (75 MHz, CDCl3): δ 9.9, 28.3, 36.3, 36.8, 38.7, 40.2, 49.9, 62.6, 73.3, 125.6, 126.0, 127.1, 127.2, 127.7, 127.8, 128.0, 128.6, 129.7, 129.9, 130.6, 130.7, 130.8, 132.6, 132.7, 133.2, 133.3, 133.9, 134.5, 142.2, 169.4, 178.2, 179.6. HRMS: calcd for C40H41ClN3NiO3 [M + H]+, 704.2190; found, 704.2193. Synthesis of Fmoc-(R)-m-bromo-phenylalanine, (R)21. To a suspension of (SC,RN,RC)-15a (740 mg, 1.01 mmol) in methanol (10 mL) was added 3 N HCl (1.68 mL, 5.04 mmol), and the reaction mixture was heated at 50 °C for 1 h. Upon disappearance of the red color of the starting complex, the reaction mixture was concentrated to give a residue. To the residue were added water (10 mL) and ethyl acetate (10 mL) and then the phases were separated. The organic phase was washed with brine (10 mL), dried over sodium sulfate, and evaporated under vacuum to afford the recycle ligand (S)-4 (488 mg, 99.3%). The aqueous phase was concentrated to give crude AA (R)-20. The residue was dissolved in water (14 mL) and tetrahydrofuran (THF) (7 mL) to which EDTA disodium salt dihydrate (375 mg, 1.01 mmol) was added, and the reaction mixture was basified with sodium carbonate (214 mg, 2.02 mmol) and sodium bicarbonate (169 mg, 2.02 mmol). To this solution was added a solution of Fmoc-OSu (374 mg, 1.11 mmol) in THF (7 mL) and stirred at rt for 16 h. After evaporation of the organic solvents, the aqueous residue was acidified with 3 N HCl to pH 2 to give a precipitate. The precipitate was filtered and washed with water (20 mL) to afford crude (R)-21. The crude solid thus obtained was



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b01424. Chiral high-performance liquid chromatography analysis data, copies of 1H and 13C NMR spectra, and modified precipitation of complex 15f (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (R.T.). *E-mail: [email protected] (V.A.S.). ORCID

Ryosuke Takeda: 0000-0003-4502-8368 9734

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

(7) (a) So, S. M.; Kim, H.; Mui, L.; Chin, J. Mimicking Nature to Make Unnatural Amino Acids and Chiral Diamines. Eur. J. Org. Chem. 2012, 229−241. (b) Metz, A. E.; Kozlowski, M. C. Recent Advances in Asymmetric Catalytic Methods for the Formation of Acyclic α,αDisubstituted α-Amino Acids. J. Org. Chem. 2015, 80, 1−7. (c) He, G.; Wang, B.; Nack, W. A.; Chen, G. Syntheses and Transformations of α-Amino Acids via Palladium-Catalyzed Auxiliary-Directed sp3 CH Functionalization. Acc. Chem. Res. 2016, 49, 635−645. (d) Sorochinsky, A. E.; Aceña, J. L.; Moriwaki, H.; Sato, T.; Soloshonok, V. Asymmetric synthesis of α-amino acids via homologation of Ni(II) complexes of glycine Schiff bases. Part 2: Aldol, Mannich addition reactions, deracemization and (S) to (R) interconversion of α-amino acids. Amino Acids 2013, 45, 1017−1033. (e) Aceñ a, J. L.; Sorochinsky, A. E.; Soloshonok, V. Asymmetric synthesis of αamino acids via homologation of Ni(II) complexes of glycine Schiff bases. Part 3: Michael addition reactions and miscellaneous transformations. Amino Acids 2014, 46, 2047−2073. (f) Soloshonok, V.; Sorochinsky, A. Practical Methods for the Synthesis of Symmetrically α,α-Disubstituted α-Amino Acids. Synthesis 2010, 2319−2344. (8) Grygorenko, O. O.; Biitseva, A. V.; Zhersh, S. Amino Sulfonic Acids, Peptidosulfonamides and Other Related Compounds. Tetrahedron 2018, 74, 1355−1421. (9) Turcheniuk, K. V.; Kukhar, V. P.; Röschenthaler, G.-V.; Aceña, J. L.; Soloshonok, V. A.; Sorochinsky, A. E. Recent Advances in the Synthesis of Fluorinated Aminophosphonates and Aminophosphonic Acids. RSC Adv. 2013, 3, 6693−6716. (10) (a) Kukhar, V. P.; Sorochinsky, A. E.; Soloshonok, V. A. Practical synthesis of fluorine-containing α- and β-amino acids: recipes from Kiev, Ukraine. Future Med. Chem. 2009, 1, 793−819. (b) Aceña, J. L.; Sorochinsky, A. E.; Moriwaki, H.; Sato, T.; Soloshonok, V. A. Synthesis of fluorine-containing α-amino acids in enantiomerically pure form via homologation of Ni(II) complexes of glycine and alanine Schiff bases. J. Fluorine Chem. 2013, 155, 21−38. (c) Sorochinsky, A.; Mikami, K.; Fustero, S.; Sánchez-Roselló, M.; Aceña, J.; Soloshonok, V. Synthesis of Fluorinated β-Amino Acids. Synthesis 2011, 3045−3079. (11) (a) Bravo, P.; Farina, A.; Kukhar, V. P.; Markovsky, A. L.; Meille, S. V.; Soloshonok, V. A.; Sorochinsky, A. E.; Viani, F.; Zanda, M.; Zappalà, C. Stereoselective Additions of α-Lithiated Alkyl-ptolylsulfoxides toN-PMP(fluoroalkyl)aldimines. An Efficient Approach to Enantiomerically Pure Fluoro Amino Compounds. J. Org. Chem. 1997, 62, 3424−3425. (b) Soloshonok, V. A.; Kirilenko, A. G.; Kukhar’, V. P.; Resnati, G. Transamination of fluorinated β-keto carboxylic esters. A biomimetic approach to β-polyfluoroalkyl-βamino acids. Tetrahedron Lett. 1993, 34, 3621−3624. (c) Shibata, N.; Nishimine, T.; Shibata, N.; Tokunaga, E.; Kawada, K.; Kagawa, T.; Sorochinsky, A. E.; Soloshonok, V. A. Organic Base-catalyzed Stereodivergent Synthesis of (R)- and (S)-3-Amino-4,4,4-trifluorobutanoic Acids. Chem. Commun. 2012, 48, 4124−4126. (12) (a) Röschenthaler, G.-V.; Kukhar, V. P.; Kulik, I. B.; Belik, M. Y.; Sorochinsky, A. E.; Rusanov, E. B.; Soloshonok, V. A. Asymmetric Synthesis of Phosphonotrifluoroalanine and Its Derivatives using Ntert-Butanesulfinyl Imine Derived from Fluoral. Tetrahedron Lett. 2012, 53, 539−542. (b) Turcheniuk, K. V.; Poliashko, K. O.; Kukhar, V. P.; Rozhenko, A. B.; Soloshonok, V. A.; Sorochinsky, A. E. Efficient asymmetric synthesis of trifluoromethylated β-aminophosphonates and their incorporation into dipeptides. Chem. Commun. 2012, 48, 11519−11521. (c) Soloshonok, V. A.; Belokon, Y. N.; Kuzmina, N. A.; Maleev, V. I.; Svistunova, N. Y.; Solodenko, V. A.; Kukhar, V. P. Asymmetric synthesis of phosphorus analogues of dicarboxylic αamino acids. J. Chem. Soc., Perkin Trans. 1 1992, 1525−1529. (13) (a) Soloshonok, V. A.; Kirilenko, A. G.; Fokina, N. A.; Kukhar, V. P.; Galushko, S. V.; Š vedas, V. K.; Resnati, G. Chemo-enzymatic approach to the synthesis of each of the four isomers of α-alkyl-βfluoroalkyl-substituted β-amino acids. Tetrahedron: Asymmetry 1994, 5, 1225−1228. (b) Soloshonok, V. A.; Tang, X.; Hruby, V. J.; Meervelt, L. V. Asymmetric Synthesis of α,β-Dialkyl-α-phenylalanines via Direct Alkylation of a Chiral Alanine Derivative with Racemic α-

Vadim A. Soloshonok: 0000-0003-0681-4526 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully acknowledge the financial support from the IKERBASQUE, Basque Foundation for Science, and the Basque Government. We are also grateful to Servicios Generales de Investigación (SGIker, UPV/EHU) for HRMS analyses.



REFERENCES

(1) (a) Henninot, A.; Collins, J. C.; Nuss, J. M. The Current State of Peptide Drug Discovery: Back to the Future? J. Med. Chem. 2018, 61, 1382−1414. (b) Blaskovich, M. A. T. Unusual Amino Acids in Medicinal Chemistry. J. Med. Chem. 2016, 59, 10807−10836. (c) Soloshonok, V. A.; Izawa, K. Asymmetric Synthesis and Application of α-Amino Acids; ACS Symposium Series #1009; Oxford University Press, 2009. (2) For the definition of tailor-made amino acids, see: Soloshonok, V. A.; Cai, C.; Hruby, V. J.; Van Meervelt, L. Asymmetric Synthesis of Novel Highly Sterically Constrained (2S,3S)-3-Methyl-3-trifluoromethyl- and (2S,3S,4R)-3-Trifluoromethyl-4-methylpyroglutamic Acids. Tetrahedron 1999, 55, 12045−12058. (3) (a) Gante, J. Peptidomimetics-Tailored Enzyme Inhibitors. Angew. Chem., Int. Ed. Engl. 1994, 33, 1699−1720. (b) Hruby, V. J.; Li, G.; Haskell-Luevano, C.; Shenderovich, M. Design of Peptides, Proteins, and Peptidomimetics in Chi Space. Biopolymers 1997, 43, 219−266. (c) Hruby, V.; Balse, P. Conformational and Topographical Considerations in Designing Agonist Peptidomimetics from Peptide Leads. Curr. Med. Chem. 2000, 7, 945−970. (d) Vagner, J.; Qu, H.; Hruby, V. J. Peptidomimetics, a Synthetic Tool of Drug Discovery. Curr. Opin. Chem. Biol. 2008, 12, 292−296. (e) Cai, M.; Cai, C.; Mayorov, A. V.; Xiong, C.; Cabello, C. M.; Soloshonok, V. A.; Swift, J. R.; Trivedi, D.; Hruby, V. J. Biological and conformational study of βsubstituted prolines in MT-II template: steric effects leading to human MC5 receptor selectivity*. J. Pept. Res. 2004, 63, 116−131. (4) (a) Ma, J. S. Unnatural Amino Acids in Drug Discovery. Chim. Oggi 2003, 21, 65−68. (b) Hodgson, D. R. W.; Sanderson, J. M. The Synthesis of Peptides and Proteins Containing Non-Natural Amino Acids. Chem. Soc. Rev. 2004, 33, 422−430. (c) Sato, T.; Izawa, K.; Aceña, J. L.; Liu, H.; Soloshonok, V. A. Tailor-Made α-Amino Acids in the Pharmaceutical Industry: Synthetic Approaches to (1R,2S)-1Amino-2-vinylcyclopropane-1-carboxylic Acid (Vinyl-ACCA). Eur. J. Org. Chem. 2016, 2757−2774. (d) Sorochinsky, A. E.; Aceña, J. L.; Moriwaki, H.; Sato, T.; Soloshonok, V. A. Asymmetric synthesis of αamino acids via homologation of Ni(II) complexes of glycine Schiff bases; Part 1: alkyl halide alkylations. Amino Acids 2013, 45, 691−718. (5) (a) Lovering, F. Escape from Flatland 2: Complexity and Promiscuity. MedChemComm 2013, 4, 515−519. (b) Fosgerau, K.; Hoffmann, T. Peptide Therapeutics: Current Status and Future Directions. Drug Discovery Today 2015, 20, 122−128. (c) Craik, D. J.; Fairlie, D. P.; Liras, S.; Price, D. The Future of Peptide-Based Drugs. Chem. Biol. Drug Des. 2013, 81, 136−147. (d) Wang, S.; Wang, Y.; Wang, J.; Sato, T.; Izawa, K.; Soloshonok, V. A.; Liu, H. The Secondgeneration of Highly Potent Hepatitis C Virus (HCV) NS3/4A Protease Inhibitors: Evolutionary Design Based on Tailor-made Amino Acids, Synthesis and Major Features of Bio-activity. Curr. Pharm. Des. 2017, 23, 4493−4554. (6) (a) Mandell, D. J.; Lajoie, M. J.; Mee, M. T.; Takeuchi, R.; Kuznetsov, G.; Norville, J. E.; Gregg, C. J.; Stoddard, B. L.; Church, G. M. Biocontainment of Genetically Modified Organisms by Synthetic Protein Design. Nature 2015, 518, 55−60. (b) Rovner, A. J.; Haimovich, A. D.; Katz, S. R.; Li, Z.; Grome, M. W.; Gassaway, B. M.; Amiram, M.; Patel, J. R.; Gallagher, R. R.; Rinehart, J.; Isaacs, F. J. Recoded Organisms Engineered to Depend on Synthetic Amino Acids. Nature 2015, 518, 89−93. 9735

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

Alkylbenzyl Bromides. A Case of High Enantiomer Differentiation at Room Temperature. Org. Lett. 2001, 3, 341−343. (c) Qiu, W.; Gu, X.; Soloshonok, V. A.; Carducci, M. D.; Hruby, V. J. Stereoselective Synthesis of Conformationally Constrained Reverse Turn Dipeptide Mimetics. Tetrahedron Lett. 2001, 42, 145−148. (14) (a) Ueki, H.; Yasumoto, M.; Soloshonok, V. A. Rational application of self-disproportionation of enantiomers via sublimationa novel methodological dimension for enantiomeric purifications. Tetrahedron: Asymmetry 2010, 21, 1396−1400. (b) Sorochinsky, A. E.; Katagiri, T.; Ono, T.; Wzorek, A.; Aceña, J. L.; Soloshonok, V. A. Optical Purifications via Self-Disproportionation of Enantiomers by Achiral Chromatography: Case Study of a Series of α-CF3-containing Secondary Alcohols. Chirality 2013, 25, 365−368. (c) Suzuki, Y.; Han, J.; Kitagawa, O.; Aceña, J. L.; Klika, K. D.; Soloshonok, V. A. A Comprehensive Examination of the Self-Disproportionation of Enantiomers (SDE) of Chiral Amides via Achiral, LaboratoryRoutine, Gravity-Driven Column Chromatography. RSC Adv. 2015, 5, 2988−2993. (15) For other types of ligands possessing axial chirality, see: (a) Takeda, R.; Kawamura, A.; Kawashima, A.; Sato, T.; Moriwaki, H.; Izawa, K.; Akaji, K.; Wang, S.; Liu, H.; Aceña, J. L.; Soloshonok, V. A. Chemical Dynamic Kinetic Resolution andS/R Interconversion of Unprotected α-Amino Acids. Angew. Chem., Int. Ed. 2014, 53, 12214−12217. (b) Jörres, M.; Aceña, J. L.; Soloshonok, V. A.; Bolm, C. Asymmetric Carbon-Carbon Bond Formation under Solventless Conditions in Ball Mills. ChemCatChem 2015, 7, 1265−1269. (16) (a) Ueki, H.; Ellis, T. K.; Martin, C. H.; Soloshonok, V. A. Efficient Large-Scale Synthesis of Picolinic Acid-Derived Nickel(II) Complexes of Glycine. Eur. J. Org. Chem. 2003, 1954−1957. (b) Ueki, H.; Ellis, T. K.; Martin, C. H.; Boettiger, T. U.; Bolene, S. B.; Soloshonok, V. A. Improved Synthesis of Proline-Derived Ni(II) Complexes of Glycine: Versatile Chiral Equivalents of Nucleophilic Glycine for General Asymmetric Synthesis of α-Amino Acids. J. Org. Chem. 2003, 68, 7104−7107. (17) (a) Ellis, T. K.; Martin, C. H.; Ueki, H.; Soloshonok, V. A. Efficient, practical synthesis of symmetrically α,α-disubstituted αamino acids. Tetrahedron Lett. 2003, 44, 1063−1066. (b) Soloshonok, V. A.; Tang, X.; Hruby, V. J. Large-scale asymmetric synthesis of novel sterically constrained 2′,6′-dimethyl- and α,2′,6′-trimethyltyrosine and -phenylalanine derivatives via alkylation of chiral equivalents of nucleophilic glycine and alanine. Tetrahedron 2001, 57, 6375−6382. (c) Wang, J.; Lin, D.; Zhou, S.; Ding, X.; Soloshonok, V. A. Asymmetric Synthesis of Sterically and Electronically Demanding Linear ω-Trifluoromethyl Containing Amino Acids via Alkylation of Chiral Equivalents of Nucleophilic Glycine and Alanine. J. Org. Chem. 2011, 76, 684−687. (18) (a) Yamada, T.; Okada, T.; Sakaguchi, K.; Ohfune, Y.; Ueki, H.; Soloshonok, V. A. Efficient Asymmetric Synthesis of Novel 4Substituted and Configurationally Stable Analogues of Thalidomide. Org. Lett. 2006, 8, 5625−5628. (b) Soloshonok, V. A.; Cai, C.; Hruby, V. J. (S)- or (R)-3-(E-Enoyl)-4-phenyl-1,3- oxazolidin-2-ones: Ideal Michael Acceptors To Afford a Virtually Complete Control of Simple and Face Diastereoselectivity in Addition Reactions with Glycine Derivatives. Org. Lett. 2000, 2, 747−750. (c) Yamada, T.; Sakaguchi, K.; Shinada, T.; Ohfune, Y.; Soloshonok, V. A. Efficient asymmetric synthesis of the functionalized pyroglutamate core unit common to oxazolomycin and neooxazolomycin using Michael reaction of nucleophilic glycine Schiff base with α,β-disubstituted acrylate. Tetrahedron: Asymmetry 2008, 19, 2789−2795. (19) (a) Soloshonok, V. A.; Kukhar, V. P.; Galushko, S. V.; Svistunova, N. Y.; Avilov, D. V.; Kuz’mina, N. A.; Raevski, N. I.; Struchkov, Y. T.; Pysarevsky, A. P.; Belokon, Y. N. General method for the synthesis of enantiomerically pure β-hydroxy-α-amino acids, containing fluorine atoms in the side chains. Case of stereochemical distinction between methyl and trifluoromethyl groups. X-Ray crystal and molecular structure of the nickel(II) complex of (2S,3S)2(trifluoromethyl)threonine. J. Chem. Soc., Perkin Trans. 1 1993, 3143−3155. (b) Soloshonok, V. A.; Avilov, D. V.; Kukhar’, V. P. Highly Diastereoselective Asymmetric Aldol Reactions of Chiral

Ni(II)-Complex of Glycine with Alkyl Trifluoromethyl Ketones. Tetrahedron: Asymmetry 1996, 7, 1547−1550. (c) Soloshonok, V. A.; Avilov, D. V.; Kukhar’, V. P.; Tararov, V. I.; Savel’eva, T. F.; Churkina, T. D.; Ikonnikov, N. S.; Kochetkov, K. A.; Orlova, S. A.; Pysarevsky, A. P.; Struchkov, Y. T.; Raevsky, N. I.; Belokon’, Y. N. Asymmetric aldol reactions of chiral Ni(II)-complex of glycine with aliphatic aldehydes. Stereodivergent synthesis of syn-(2S)- and syn-(2R)-βalkylserines. Tetrahedron: Asymmetry 1995, 6, 1741−1756. (20) (a) Soloshonok, V. A.; Avilov, D. V.; Kukhar, V. P.; Meervelt, L. V.; Mischenko, N. Highly Diastereoselective Aza-Aldol Reactions of a Chiral Ni(II) Complex of Glycine with Imines. An Efficient Asymmetric Approach to 3-perfluoroalkyl-2,3-diamino Acids. Tetrahedron Lett. 1997, 38, 4671−4674. (b) Kawamura, A.; Moriwaki, H.; Röschenthaler, G.-V.; Kawada, K.; Aceña, J. L.; Soloshonok, V. A. Synthesis of (2S,3S)-β-(trifluoromethyl)-α,β-diamino acid by Mannich addition of glycine Schiff base Ni(II) complexes to N-tertbutylsulfinyl-3,3,3-trifluoroacetaldimine. J. Fluorine Chem. 2015, 171, 67−72. (21) (a) Ellis, T. K.; Hochla, V. M.; Soloshonok, V. A. Efficient Synthesis of 2-Aminoindane-2-carboxylic Acid via Dialkylation of Nucleophilic Glycine Equivalent. J. Org. Chem. 2003, 68, 4973−4976. (b) Ellis, T. K.; Martin, C. H.; Ueki, H.; Soloshonok, V. A. Efficient, practical synthesis of symmetrically α,α-disubstituted α-amino acids. Tetrahedron Lett. 2003, 44, 1063−1066. (22) (a) Soloshonok, V.; Boettiger, T.; Bolene, S. Asymmetric Synthesis of (2S,3S)- and (2R,3R)-α,β-Dialkyl-α-amino Acids via Alkylation of Chiral Nickel(II) Complexes of Aliphatic α-Amino Acids with Racemic α-Alkylbenzyl Bromides. Synthesis 2008, 2594− 2602. (b) Yamamoto, J.; Kawashima, A.; Kawamura, A.; Abe, H.; Moriwaki, H.; Shibata, N.; Soloshonok, V. A. Operationally Convenient and Scalable Asymmetric Synthesis of (2S )- and (2R )-α-(Methyl)cysteine Derivatives through Alkylation of Chiral Alanine Schiff Base NiII Complexes. Eur. J. Org. Chem. 2017, 1931−1939. (23) (a) Kawashima, A.; Xie, C.; Mei, H.; Takeda, R.; Kawamura, A.; Sato, T.; Moriwaki, H.; Izawa, K.; Han, J.; Aceña, J. L.; Soloshonok, V. A. Asymmetric synthesis of (1R,2S)-1-amino-2-vinylcyclopropanecarboxylic acid by sequential SN2-SN2′ dialkylation of (R)-N-(benzyl)proline-derived glycine Schiff base Ni(ii) complex. RSC Adv. 2015, 5, 1051−1058. (b) Kawashima, A.; Shu, S.; Takeda, R.; Kawamura, A.; Sato, T.; Moriwaki, H.; Wang, J.; Izawa, K.; Aceña, J. L.; Soloshonok, V. A.; Liu, H. Advanced Asymmetric Synthesis of (1R,2S)-1-Amino-2vinylcyclopropanecarboxylic Acid by Alkylation/Cyclization of Newly Designed Axially Chiral Ni(II) Complex of Glycine Schiff Base. Amino Acids 2016, 48, 973−986. (24) (a) Taylor, S. M.; Yamada, T.; Ueki, H.; Soloshonok, V. A. Asymmetric Synthesis of Enantiomerically Pure 4-Aminoglutamic Acids via Methylenedimerization of Chiral Glycine Equivalents with Dichloromethane under Operationally Convenient Conditions. Tetrahedron Lett. 2004, 45, 9159−9162. (b) Wang, J.; Liu, H.; Aceña, J. L.; Houck, D.; Takeda, R.; Moriwaki, H.; Sato, T. Synthesis of bis-α,α′-amino acids through diastereoselective bis-alkylations of chiral Ni(ii)-complexes of glycine. Org. Biomol. Chem. 2013, 11, 4508−4515. (25) (a) Metrano, A. J.; Miller, S. J. Peptide-Catalyzed Conversion of Racemic Oxazol-5(4H)-ones into Enantiomerically Enriched αAmino Acid Derivatives. J. Org. Chem. 2014, 79, 1542−1554. (b) So, S. M.; Moozeh, K.; Lough, A. J.; Chin, J. Highly Stereoselective Recognition and Deracemization of Amino Acids by Supramolecular Self-Assembly. Angew. Chem., Int. Ed. 2014, 53, 829− 832. (c) Moozeh, K.; So, S. M.; Chin, J. Catalytic Stereoinversion ofLAlanine to DeuteratedD-Alanine. Angew. Chem., Int. Ed. 2015, 54, 9381−9385. (d) Fanelli, R.; Jeanne-Julien, L.; René, A.; Martinez, J.; Cavelier, F. Stereoselective synthesis of unsaturated α-amino acids. Amino Acids 2015, 47, 1107−1115. (e) Schwieter, K. E.; Johnston, J. N. Enantioselective synthesis of d-α-amino amides from aliphatic aldehydes. Chem. Sci. 2015, 6, 2590−2595. (f) Bull, S.; Hutchby, M.; Sedgwick, A. Orthogonally Protected Schöllkopf’s Bis-lactim Ethers for the Asymmetric Synthesis of α-Amino Acid Derivatives and 9736

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737

ACS Omega

Article

Dipeptide Esters. Synthesis 2016, 48, 2036−2049. (g) Wangweerawong, A.; Hummel, J. R.; Bergman, R. G.; Ellman, J. A. Preparation of Enantiomerically Pure Perfluorobutanesulfinamide and Its Application to the Asymmetric Synthesis of α-Amino Acids. J. Org. Chem. 2016, 81, 1547−1557. (h) Schettini, R.; Nardone, B.; De Riccardis, F.; Della Sala, G.; Izzo, I. Cyclopeptoids as Phase-Transfer Catalysts for the Enantioselective Synthesis of α-Amino Acids. Eur. J. Org. Chem. 2014, 7793−7797. (i) Morrill, L. C.; Lebl, T.; Slawin, A. M. Z.; Smith, A. D. Catalytic asymmetric α-amination of carboxylic acids using isothioureas. Chem. Sci. 2012, 3, 2088−2093. (j) Etxabe, J.; Izquierdo, J.; Landa, A.; Oiarbide, M.; Palomo, C. Catalytic Enantioselective Synthesis of N,Cα,Cα-Trisubstituted α-Amino Acid Derivatives Using 1H-Imidazol-4(5H)-ones as Key Templates. Angew. Chem., Int. Ed. 2015, 54, 6883−6886. (26) (a) Soloshonok, V. A.; Ellis, T. K.; Ueki, H.; Ono, T. Resolution/Deracemization of Chiral α-Amino Acids Using Resolving Reagents with Flexible Stereogenic Centers. J. Am. Chem. Soc. 2009, 131, 7208−7209. (b) Sorochinsky, A. E.; Ueki, H.; Aceña, J. L.; Ellis, T. K.; Moriwaki, H.; Sato, T. Chemical approach for interconversion of (S)- and (R)-α-amino acids. Org. Biomol. Chem. 2013, 11, 4503− 4507. (c) Soloshonok, A. E.; Ueki, H.; Aceña, J. L.; Ellis, T. K.; Moriwaki, H.; Sato, T.; Soloshonok, V. A. Chemical deracemization and (S) to (R) interconversion of some fluorine-containing α-amino acids. J. Fluorine Chem. 2013, 152, 114−118. (27) (a) Zhou, S.; Wang, J.; Chen, X.; Aceña, J. L.; Soloshonok, V. A.; Liu, H. Chemical Kinetic Resolution of Unprotected β-Substituted β-Amino Acids Using Recyclable Chiral Ligands. Angew. Chem., Int. Ed. 2014, 53, 7883−7886. (b) Zhou, S.; Wang, S.; Wang, J.; Nian, Y.; Peng, P.; Soloshonok, V. A.; Liu, H. Configurationally Stable (S )- and (R )-α-Methylproline-Derived Ligands for the Direct Chemical Resolution of Free Unprotected β3 -Amino Acids. Eur. J. Org. Chem. 2018, 1821−1832. (28) (a) Tessaro, D.; Cerioli, L.; Servi, S.; Viani, F.; D’Arrigo, P. lAmino Acid Amides via Dynamic Kinetic Resolution. Adv. Synth. Catal. 2011, 353, 2333−2338. (b) D’Arrigo, P.; Cerioli, L.; Fiorati, A.; Servi, S.; Viani, F.; Tessaro, D. Naphthyl-l-α-amino Acids via ChemoEnzymatic Dynamic Kinetic Resolution. Tetrahedron: Asymmetry 2012, 23, 938−944. (c) D’Arrigo, P.; Cerioli, L.; Servi, S.; Viani, F.; Tessaro, D. Synergy between Catalysts: Enzymes and Bases. DKR of Non-Natural Amino Acids Derivatives. Catal. Sci. Technol. 2012, 2, 1606−1616. (d) Baxter, S.; Royer, S.; Grogan, G.; Brown, F.; HoltTiffin, K. E.; Taylor, I. N.; Fotheringham, I. G.; Campopiano, D. J. An Improved Racemase/Acylase Biotransformation for the Preparation of Enantiomerically Pure Amino Acids. J. Am. Chem. Soc. 2012, 134, 19310−19313. (e) Yasukawa, K.; Asano, Y. Enzymatic Synthesis of Chiral Phenylalanine Derivatives by a Dynamic Kinetic Resolution of Corresponding Amide and Nitrile Substrates with a Multi-Enzyme System. Adv. Synth. Catal. 2012, 354, 3327−3332. (29) (a) Ellis, T. K.; Ueki, H.; Yamada, T.; Ohfune, Y.; Soloshonok, V. A. Design, Synthesis, and Evaluation of a New Generation of Modular Nucleophilic Glycine Equivalents for the Efficient Synthesis of Sterically Constrained α-Amino Acids. J. Org. Chem. 2006, 71, 8572−8578. (b) Soloshonok, V. A.; Ueki, H.; Ellis, T. K.; Yamada, T.; Ohfune, Y. Application of modular nucleophilic glycine equivalents for truly practical asymmetric synthesis of β-substituted pyroglutamic acids. Tetrahedron Lett. 2005, 46, 1107−1110. (30) (a) Bergagnini, M.; Fukushi, K.; Han, J.; Shibata, N.; Roussel, C.; Ellis, T. K.; Aceña, J. L.; Soloshonok, V. A. NH-type of Chiral Ni(II) Complexes of Glycine Schiff Base: Design, Structural Evaluation, Reactivity and synthetic applications. Org. Biomol. Chem. 2014, 12, 1278−1291. (b) Soloshonok, V. A.; Ellis, T. K. Design and Synthesis of a New Generation of ‘NH’-Ni(II) Complexes of Glycine Schiff Bases and their Unprecedented C-H vs. N-H Chemoselectivity in Alkyl Halide Alkylations and Michael Addition Reactions. Synlett 2006, 0533−0538. (c) Takeda, R.; Kawamura, A.; Kawashima, A.; Moriwaki, H.; Sato, T.; Aceña, J. L.; Soloshonok, V. A. Design and synthesis of (S)- and (R)-α-(phenyl)ethylamine-derived NH-type ligands and their application for the chemical resolution of α-amino acids. Org. Biomol. Chem. 2014, 12, 6239−6249.

(31) Takeda, R.; Kawamura, A.; Kawashima, A.; Sato, T.; Moriwaki, H.; Izawa, K.; Abe, H.; Soloshonok, V. A. Second-order asymmetric transformation and its application for the practical synthesis of αamino acids. Org. Biomol. Chem. 2018, 16, 4968−4972. (32) Wanka, L.; Iqbal, K.; Schreiner, P. R. The Lipophilic Bullet Hits the Targets: Medicinal Chemistry of Adamantane Derivatives. Chem. Rev. 2013, 113, 3516−3604. (33) (a) Jamison, M. M.; Turner, E. E. 84. The Inter-Relation of First- and Second-Order Asymmetric Transformations. J. Chem. Soc. 1942, 437−440. (b) Busch, D. H. High Yield Resolution of Tris(ethylenediamine)-cobalt(III) by a Second-order Asymmetric Process. J. Am. Chem. Soc. 1955, 77, 2747−2748. (c) Faigl, F.; Mátravölgyi, B.; Holczbauer, T.; Czugler, M.; Madarász, J. Resolution of 1-[2Carboxy-6-(trifluoromethyl)phenyl]-1H-pyrrole-2-carboxylic Acid with Methyl (R)-2-Phenylglycinate, Reciprocal Resolution and Second Order Asymmetric Transformation. Tetrahedron: Asymmetry 2011, 22, 1879−1884. (d) Tsubaki, K.; Miura, M.; Morikawa, H.; Tanaka, H.; Kawabata, T.; Furuta, T.; Tanaka, K.; Fuji, K. Synthesis of Optically Active Oligonaphthalenes via Second-Order Asymmetric Transformation. J. Am. Chem. Soc. 2003, 125, 16200−16201. (e) Yamada, K.-i.; Ishii, R.; Nakagawa, H.; Kawazura, H. A SecondOrder Asymmetric Transformation of Racemic 2-Hydroxymethyl[5]thiaheterohelicene into a Single Enantiomer upon Uptake by Bovine Serum Albumin. Tetrahedron: Asymmetry 1996, 7, 737−746. (f) Donaldson, W. A.; Shang, L.; Rogers, R. D. Reactivity of Tricarbonyl(pentadienyl)iron(1+) Cations: Preparation of an Optically Pure Tricarbonyl(diene)iron Complex via Second-Order Asymmetric Transformation. Organometallics 1994, 13, 6−7. (g) Smrcina, M.; Lorenc, M.; Hanus, V.; Sedmera, P.; Kocovsky, P. Synthesis of Enantiomerically Pure 2,2′-Dihydroxy-1,1′-binaphthyl, 2,2′-Diamino-1,1′-binaphthyl, and 2-Amino-2′-hydroxy-1,1′-binaphthyl. Comparison of Processes Operating as Diastereoselective Crystallization and as Second Order Asymmetric Transformation. J. Org. Chem. 1992, 57, 1917−1920. (34) Houck, D.; Luis Aceña, J.; Soloshonok, V. A. Alkylations of Chiral Nickel(II) Complexes of Glycine under Phase-Transfer Conditions. Helv. Chim. Acta 2012, 95, 2672−2679.

9737

DOI: 10.1021/acsomega.8b01424 ACS Omega 2018, 3, 9729−9737