Next Article in Journal
Mode of Antibacterial Action of Tomatidine C3-Diastereoisomers
Previous Article in Journal
Novel Dihydrocoumarins Induced by Radiolysis as Potent Tyrosinase Inhibitors
Previous Article in Special Issue
Synthesis of Chalcones: An Improved High-Yield and Substituent-Independent Protocol for an Old Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphine Catalyzed Michael-Type Additions: The Synthesis of Glutamic Acid Derivatives from Arylidene-α-amino Esters †

by
Lesly V. Rodríguez-Flórez
1,
María González-Marcos
1,
Eduardo García-Mingüens
2,
María de Gracia Retamosa
1,
Misa Kawase
1,
Elisabet Selva
2 and
José M. Sansano
1,*
1
Departamento de Química Orgánica, Centro de Innovación en Química Avanzada (ORFEO-CINQA) and Instituto de Síntesis Orgánica, Universidad de Alicante, Ctra. Alicante-San Vicente s/n, 03080 Alicante, Spain
2
Medalchemy, S. L. Ancha de Castelar, 46-48, entlo. A. San Vicente del Raspeig, 03690 Alicante, Spain
*
Author to whom correspondence should be addressed.
Dedicated to the memory of Prof. David A. Evans.
Molecules 2024, 29(2), 342; https://doi.org/10.3390/molecules29020342
Submission received: 12 December 2023 / Revised: 28 December 2023 / Accepted: 7 January 2024 / Published: 10 January 2024
(This article belongs to the Special Issue A Journey of Organic Chemistry in Spain)

Abstract

:
The reaction of arylidene-α-amino esters with electrophilic alkenes to yield Michael-type addition compounds is optimized using several phosphines as organocatalysts. The transformation is very complicated due to the generation of several final compounds, including those derived from the 1,3-dipolar cycloadditions. For this reason, the selection of the reaction conditions is a very complex task and the slow addition of the acrylic system is very important to complete the process. The study of the variation in the structural components of the starting imino ester is performed as well as the expansion of other electron-poor alkenes. The crude products have a purity higher than 90% in most cases without any purification. A plausible mechanism is detailed based on the bibliography and the experimental results. The synthesis of pyroglutamate entities, after the reduction of the imino group and cyclization, is performed in high yields. In addition, the hydrolysis of the imino group, under acidic media, represents a direct access to glutamate surrogates.

Graphical Abstract

1. Introduction

Proteinogenic and non-proteinogenic α-amino acids (AAs) constitute one of the five most important families of essential molecules in many scientific areas. The synthesis of these compounds [1,2,3] obeys several general patterns such as the following [4,5,6,7,8,9,10,11,12,13]: (a) the introduction of the hydrogen atom in the appropriate carbon-carbon or carbon-nitrogen double bond structures; (b) the employment of a methodology able to insert the nitrogen atom at the α-position to the ester group (electrophilic nitrogen source); (c) reactions involving the incorporation of a carboxy group and (d) the coupling of the α-side chain to the AA template. Considering this last approach, α-substituted glutamates have been mainly obtained via the Michael-type additions of glycine derivatives (glycine templates) onto the corresponding α,β-unsaturated reagents. This reliable strategy employs N-arylidene-α-amino acid esters [14,15,16] or tert-butyl N-benzylidieneamino glycinate [17,18,19,20,21,22] (Scheme 1a) and even activated N-arylideneaminomalonates [23,24,25,26,27,28,29,30] (Scheme 1b) as starting materials. In all cases, phase transfer catalysis (PTC) conditions or the employment of organic superbases are the most common trends to complete the reaction. An important drawback detected in the reactions regarding glycine templates is the double alkylation process at the α-position.
Having in mind the natural impact [31] and usefulness of glutamates (and their pyroglutamate surrogates) [32] as synthetic key building blocks and their presence in many biologically active molecules [33,34,35,36,37,38,39,40,41,42,43], we have studied a novel Michael-type approach to their preparation. This methodology consists of the base-free reaction of alkyl N-arylidene-α-amino acid esters with conjugated alkenes [44]. Here, the transformation operates in the presence of a substoichiometric amount of a phosphine, which acts as an organocatalyst (Scheme 1c) [45].

2. Results and Discussion

The reaction between imino esters 1 and Michael acceptors can be controlled to afford the pure conjugated product or the corresponding pyrrolidine via 1,3-dipolar cycloaddition (1,3-DC) [46,47,48,49,50,51,52,53]. These two products are found in the final crude mixture in some cycloadditions. With the aim of selecting the formation of the glutamate structure, we consider the ability of phosphines to catalyze this process. Thus, the reaction of imino ester 1a (1 equiv) with methyl acrylate (1 equiv) was treated with the corresponding phosphine (10 mol% loading), using toluene as a solvent, at 25 °C (Scheme 2 and Table 1). The optimization of this reaction was a very complicated task due to the presence of three identified secondary compounds (3a, 4 and 5, see experimental part and Supplementary Information). Initially, the nature of the phosphine was tested. The nucleophilicity of a triarylphosphine, such as Ph3P, was not enough to promote the desired reaction (Table 1, entry 1). 1,2-Bis(diphenylphophino)ethane (dppe) did not complete the reaction after 15 h, affording 2a impurified with cycloadducts 3a and 5 (Table 1, entry 2). The consumption of the starting material 1a was achieved after 72 h of reaction, but these two impurities were detected as shown in entries 3 and 4 of Table 1. The slow addition (60 min) of imino ester 1a to the reaction mixture avoided the 1,3-DC, although it promoted the generation of the diester 4 as a consequence of the presence of an excess of the alkene (Table 1, entry 5). Byproduct 4 was suppressed with the slow addition of the methyl acrylate (60 min), but imidazolidine 5 was observed instead after 72 h of reaction (Table 1, entry 6). A shorter reaction time avoided the completion of the reaction and obtained significant amounts of starting imino ester 1a (Table 1, entry 7). Both tri-n-butyl and tri-tert-butyl phosphines exclusively afforded compound 5 or the diester 4, even in the process involving a slow addition of the acrylate (Table 1, entries 8–10). Perhaps, the nucleophilicity of trialkyl phosphines is excessive for controlling the desired process. So, the modulation of this property via combining an aryl group with two alkyl substituents bonded to the phosphorous atom, was next attempted. Then, Me2PhP was used as a catalyst, demonstrating a rapid conversion (2 h) but generating large amounts of compounds 4 and 5 (Table 1, entries 11 and 12). Longer addition times (60 min) of the methyl acrylate favored the formation of the cycloaddition products 3a and 5. However, lower addition times contributed to an increase in the presence of compound 4. The optimal addition time of methyl acrylate was 30 min (Table 1, entries 11–16) furnishing a very clean reaction crude product using 1H NMR. In fact, compound 2a did not require any additional purification after the work-up (see experimental section). The lowering of the catalyst loading to 5 mol% did not efficiently promote the reaction (Table 1, entry 17). The effect of the solvent was not significant, obtaining similar results when the reactions were performed in dichloromethane, THF or acetonitrile. Solvents such as MTBE, EtOAc, acetone and water were not suitable. No reaction product 2a was identified in the 1H NMR spectra when dimethylphenylphosphine was substituted with the same loading of triethylamine, DABCO or DBU as catalysts. All the product ratios detailed in Table 1 were accurately analyzed using 1H NMR integrals of these crude materials.
The plausible mechanism of all these processes is described in Scheme 3. The excess of methyl acrylate in the reaction media causes the 1,4-attack of the intermediate I (generated by the Michael-type addition of the phosphine and the acrylate) on another equivalent of methyl acrylate. After a prototropy shift caused by the stabilization of the negative charge in intermediate III, the dimer 4 is formed with the regeneration of the catalyst. The low amounts of the intermediate I, obtained after the slow addition of methyl acrylate, are surrounded by a large excess of imino ester 1a, which can be deprotonated by enolate I, furnishing the stabilized carbanion V. The direct attack of V on IV gives the desired Michael-type adduct 2a with the elimination of the active catalyst. The fine-tuning of the nucleophilicity of the phosphine in the last step is crucial and the overall mechanism is very sensitive to this feature. On the other hand, when the addition of the phosphine is very slow, or simply does not occur, the excess of 1a can give the fleeting azomethine ylide VI after a 1,2-prototropy shift at room temperature. This process is very slow, but the ylide VI is trapped immediately by methyl acrylate (which does not undergo the transformation to the corresponding intermediate I), giving access to cycloadduct 3a. This is a very fast reaction compared with the analogous Mannich type-cyclization to yield product 5. In consequence, entry 15 of Table 1 employs the optimal phosphine. This phosphine is able to generate intermediate I, which has a preference for abstracting the α-H of the imino ester 1a rather than other different reactions. The slow addition of the acrylate inhibits the route to yield dimer 4, but favors the route to generate the expected compound 2a. There is a paramount detail in this last step. The phosphine does not activate the imino group, neither the imino ester 1a nor other different intermediate species, such as V or even VI. The absence of the route IIVV favors the presence of 4, allowing the generation of the ylide VI and giving rise to pyrrolidine 3a (after reaction with methyl acrylate) or imidazolidine 5 (via the self-addition of 1a). Products 3a and 5 are formed when the addition took 60 min and they are minimized when performing the addition in 30 min (Table 1, entries 14 and 15). However, an alternative base-propagation mechanism where the enolate V promotes a Michael-type addition, and not a SN2 onto the phosphonium intermediate, cannot be discarded [54].
With the best reaction conditions established in entry 15 of Table 1, the scope of imino esters 1 and alkyl acrylates was investigated. The results of the crude yields, determined by 1H NMR spectra, using dimethyl terephthalate as an internal standard [55], are depicted in Scheme 3 [56]. The variation in the aromatic moiety in glycinate-derived imino esters was well tolerated (2al, Scheme 4), with even starters containing heterocyclic units as 2-thienyl or 3-pyridyl (2jl, Scheme 4). Methyl, n-butyl and tert-butyl acrylates were randomly employed, giving satisfactory results (Scheme 4) [57]. However, when imino esters with a substituent at the α-position were tested, a 20 mol% of the catalyst loading was required to complete the transformations. Also, different slow addition times and an excess of the acrylate component were modified accordingly in these examples to obtain the best yields and purities (see experimental part). Thus, glutamate derivatives 2mq were obtained (and characterized without purification) in very good yields (Scheme 4). However, with functionalized α-side chain α-amino acid-derived imino esters 1, such as tryptophan, O-benzylserine and methionine, the conversions were good but the crude compounds 2rt were not pure and could not be characterized (grey color in Scheme 4). These three last examples were immediately transformed into the corresponding pyroglutamate surrogates 8 (see below in Scheme 6b). Despite the large quantity of secondary products expected, the final compounds were obtained as crude materials with purities higher than 90% in most cases (see experimental part) without any purification. Chromatographic separation was not possible for those of lower purity due to the formation of amines and aldehydes from the imines on SiO2.
Other acrylic systems like N,N-dimethylacrylamide reacted satisfactorily under these conditions, affording the glutamine derivative 2u with a 90% yield. In the presence of acrylonitrile or phenyl vinyl sulfone (1 equiv), the corresponding molecules 6, originating from a double addition of the alkene, were detected as byproducts. The full conversion of the reaction performed with acrylonitrile was achieved using a 20 mol% of the catalyst and 4 equiv of the Michael-type acceptor. After that, the α,α-disubstituted imino ester 6v was obtained with a 94% yield (by 1H NMR, Scheme 5). Bulkier phenyl vinyl sulfone did not afford pure and clean compound 6w due to the presence of tantamount quantities of monoalkylated substance 2w, even working in the presence of an excess of alkene and using 20 mol% of the catalyst (Scheme 5). This preference for the phenyl vinyl sulfone and acrylonitrile to generate double alkylation products 6, unlike the acrylic esters and acrylamide, is due to the existence of a lower energy LUMO. LUMO’s energies of phenyl vinyl sulfone and acrylonitrile are −1.891 [58] and −2.52 eV [59], respectively, whilst the LUMO’s energy of the methyl acrylate is −0.08 eV [60].
A straightforward access to pyroglutamates, which are key units in biotechnology, in biomedicine and for the treatment of neurodegenerative illnesses [32,61,62,63,64], is easily envisaged. Employing conventional transformations, such as reduction with sodium borohydride followed by mild cyclization conditions using silica gel in refluxing ethyl acetate, substituted pyroglutamates 8 were isolated in moderate to good yields after flash chromatography (Scheme 6a). Non-isolated adducts 2rt, described in Scheme 4, were directly submitted to these sequential reduction-cyclization conditions, obtaining the pyroglutamates 8d, 8e and 8f in 45, 46 and 42% overall yields, respectively (from imino ester 1, Scheme 6b). In addition, the rapid access to glutamic ester derivative 9 was achieved with a 90% yield via treatment with 2M HCl/Et2O (Scheme 7).

3. Materials and Methods

3.1. General

All commercially available reagents and solvents were used without further purification, only aldehydes were also distilled prior to use. Analytical TLC was performed on Schleicher & Schuell F1400/LS 254 (Schleicher & Schuell BioScience GmbH, Dassel, Germany) silica gel plates, and the spots were visualized under UV light (λ = 254 nm). Flash chromatography was carried out on hand-packed columns of Merck silica gel 60 (0.040–0.063 mm). Melting points were determined with a Reichert Thermovar hot plate apparatus and are uncorrected. The structurally most important peaks of the IR spectra (recorded using a Nicolet 510 P-FT (Thermo-Fisher Scientific, Waltham, MA, USA) are listed and wavenumbers are given in cm−1. NMR spectra were obtained using a Bruker AC-300 (Bruker Corporation, Billerica, MA, USA) or AC-400 (Bruker Corporation, Billerica, MA, USA) and were recorded at 300 or 400 MHz for 1H NMR and 75 or 100 MHz for 13C{1H} NMR, using CDCl3 as the solvent and TMS as the internal standard (0.00 ppm) unless otherwise stated. The following abbreviations are used to describe peak patterns where appropriate: s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet or unresolved and br s = broad signal. All coupling constants (J) are given in Hz and chemical shifts in ppm. 13C{1H} NMR spectra were referenced to CDCl3 at 77.16 ppm. Chemical yields and purities of compounds 2 and 6 were calculated by the integration of 1H NMR spectra using dimethyl terephthalate as the internal standard [17]. Low-resolution electron impact (EI) mass spectra were obtained at 70 eV using a Shimadzu QP-5000 via injection or DIP; fragment ions in m/z are given with relative intensities (%) in parentheses. High-resolution mass spectra (HRMS) were measured on an instrument using a quadrupole time-of-flight mass spectrometer (QTOF) and also through the electron impact mode (EI) at 70 eV using a Finnigan VG Platform or a Finnigan MAT 95S.
The synthesis of the starting α-imino esters 1 was performed following the described procedure [65,66]. Thus, the corresponding α-amino acid alkyl ester hydrochloride (3.0 mmol), the corresponding aldehyde (2.3 mmol) and MgSO4 were dissolved in dry dichloromethane (5 mL). Then, triethylamine (3.0 mmol) was slowly added and the mixture was then stirred for 18 h. Later, the reaction mixture was filtered, extracted with dichloromethane (3 × 10 mL), dried over MgSO4 and evaporated under reduced pressure, obtaining 1, which was employed without further purification.
Compounds 3a [66,67], 4 [68] and 5 [69] were obtained using procedures described in the literature just to compare the signals for the determination of the proportions depicted in Table 1 (see Supplementary Materials).

3.2. General Experimental Procedure for the Synthesis of Michael Type Addition Products 2

In a round-button flask and under argon, the corresponding iminoester 1 (0.5 mmol) in toluene (1.5 mL) and the catalyst dimethylphenylphosphine (0.05 mmol) were added. Then a solution of electrophilic alkene (1 mmol) in toluene (1 mL) was added dropwise over a 30 min period using an addition pump. Then, the mixture was stirred at room temperature for 24 h. The solvent was removed under reduced pressure, to afford the corresponding final product 2.
Dimethyl (E)-2-(benzylideneamino)pentanedioate (2a): Pale yellow oil (118.2 mg, 96%, 92% purity). IR (neat) νmax: 1735, 1643, 1438, 1253, 1199, 1164, 755, 694 cm−1. 1H NMR (400 MHz) δ 8.29 (s, 1H, HC=N), 7.82–7.72 (m, 2H, ArH), 7.48–7.37 (m, 3H, ArH), 4.07 (dd, J = 8.1, 5.0 Hz, 1H, NCHCO2Me), 3.74 (s, 3H, CO2CH3), 3.65 (s, 3H, CO2CH3), 2.43–2.38 (m, 2H, CH2), 2.36–2.23 (m, 2H, CH2). 13C NMR (101 MHz) δ 173.3, 171.9 (C=O), 164.3 (C=N), 135.5(CAr), 131.3(CHAr), 128.6 (4x CHAr), 71.8 (CH), 52.2, 51.6 (CH3), 30.2, 28.3 (CH2). MS (EI) m/z: 263 (M+, 11%), 204 (74), 203 (38), 190 (43), 144 (100), 130 (29), 117 (31), 104 (39), 90 (21). HRMS (ESI): m/z calcd for C14H17NO4 [M+] 263.1158; found: 263.1159.
5-Butyl 1-methyl (E)-2-(benzylideneamino)pentanedioate (2b): Pale yellow oil (146 mg, 96%, 91% purity). IR (neat) vmax: 1731, 1643, 1438, 1390, 1168, 1068, 732, 694 cm−1. 1H NMR (400 MHz) δ 8.26 (s, 1H, HC=N), 7.90–7.66 (m, 2H, ArH), 7.51–7.21 (m, 3H, ArH), 4.37–3.92 (m, 3H, NCHCO2Me, CO2CH2CH2CH2CH3), 3.86–3.55 (m, 3H, CO2CH3), 2.34 (d, J = 2.0 Hz, 2H, NCHCH2CH2), 2.33–1.88 (m, 2H, NCHCH2CH2), 1.70–1.41 (m, 2H, CO2CH2CH2CH2CH3), 1.41–1.19 (m, 2H, CO2CH2CH2CH2CH3), 1.07–0.68 (m, 3H, CO2CH2CH2CH2CH3). 13C NMR (101 MHz) δ 172.88 (C=O), 171.90 (C=O), 164.25 (C=N), 135.53 (CAr), 131.26 (CHAr), 128.59 (CHAr x4), 71.81 (CH), 64.34 (CH2), 52.19 (CH3), 30.61 (CH2), 30.41 (CH2), 28.33 (CH2), 19.09 (CH2), 13.66 (CH3). MS (EI) m/z: 305 (M+, 17%), 246 (66), 232 (38), 190 (81), 144 (100). HRMS (ESI): m/z calcd for C17H23NO4 [M+] 305.1627; found: 305.1622.
Dimethyl (E)-2-((4-methylbenzylidene)amino)pentanedioate (2c): Pale yellow oil (119.4 mg, 86%, 95% purity). IR (neat) νmax: 1735, 1643, 1438, 1253, 1199, 1168, 813 cm−1. 1H NMR (300 MHz) δ 8.25 (s, 1H, HC=N), 7.66 (d, J = 8.1 Hz, 2H, ArH), 7.22 (d, J = 7.7 Hz, 2H, ArH), 4.04 (dd, J = 8.0, 4.9 Hz, 1H, NCHCO2Me), 3.74 (s, 3H, CO2CH3), 3.64 (s, 3H, CO2CH3), 2.39 (s, 3H, CH3Ar), 2.38–2.22 (m, 4H, CH2-CH2). 13C NMR (101 MHz) δ 173.5 (C=O), 172.3 (C=O), 164.2 (C=N), 141.7 (CAr), 132.9 (CH3CAr), 129.3 (CHAr), 128.6 (CHAr), 71.8 (CH), 52.2, 51.6 (CH3), 30.1, 28.3 (CH2), 21.5 (CH3CAr). MS (EI) m/z: 277 (M+, 17%), 218 (64), 217 (32) 204 (48), 158 (100), 144 (34), 131 (28), 130 (24), 118 (40). HRMS (ESI): m/z calcd for C15H19NO4 [M+] 277.1314; found: 277.1322.
Dimethyl (E)-2-[(naphth-2-ylmethylene)amino)]pentanedioate (2d): Pale yellow prisms (147.2 mg, 90%, 92% purity). mp: 78–79 °C (n-hexane:AcOEt). IR (neat) νmax: 1727, 1639, 1434, 1176, 1095, 829, 752 cm−1. 1H NMR (300 MHz) δ 8.46 (s, 1H, HC=N), 8.10 (s, 1H, ArH), 8.03 (dd, J = 8.6, 1.4 Hz, 1H, ArH), 7.94–7.81 (m, 3H, ArH), 7.59–7.47 (m, 2H, ArH), 4.15 (dd, J = 7.7, 5.0 Hz, 1H, NCHCO2Me), 3.78 (s, 3H, CO2CH3), 3.66 (s, 3H, CO2CH3), 2.49–2.42 (m, 2H, CH2), 2.41–2.28 (m, 2H, CH2). 13C NMR (101 MHz) δ 173.3 (C=O), 171.9 (C=O), 164.3 (C=N), 134.9 (C=CCH=N), 133.2, 133.0 (PhC), 130.7, 128.7, 128.5, 127.9, 127.6, 126.6, 124.0 (PhCH), 71.8 (NCH), 52.2, 51.6 (CH3), 30.2, 28.4 (CO2CH3). MS (EI) m/z: 313 (M+, 31%), 254 (100), 240 (61), 222 (28), 194 (99), 180 (69), 167 (66), 154 (57), 140 (30), 139 (51), 127 (24). HRMS (ESI): m/z calcd for C18H19NO4 [M+] 313.1314; found: 313.1331.
Dimethyl (E)-2-[(4-methoxybenzylidene)amino]pentanedioate (2e): yellow oil (141.9 mg, 87%, 92% purity). IR (neat) νmax: 1735, 1249, 1164, 1025, 833 cm−1. 1H NMR (400 MHz) δ 8.21 (s, 1H, HC=N), 7.75–7.69 (m, 2H, ArH), 6.95–6.90 (m, 2H, ArH), 4.02 (dd, J = 8.1, 5.0 Hz, 1H, NCHCO2Me), 3.84 (s, 3H, CO2CH3), 3.74 (s, 3H, CO2CH3), 3.64 (s, 3H, CH3OAr), 2.45–2.15 (m, 4H, CH2-CH2). 13C NMR (75 MHz) δ 173.34 (C=O), 172.13 (C=O), 163.54 (C=N), 162.15 (OCH3CAr), 130.26 (2xCHAr), 128.52 (CAr), 114.00 (2xCHAr), 71.75(CH), 55.38 (OCH3CAr), 52.19, 51.58 (CO2CH3), 30.22, 28.39 (CH2). MS (EI) m/z: 293 (M+, 25%), 262 (21), 234 (69), 233 (41), 220 (62), 174 (100), 160 (29), 147 (23), 134 (40). HRMS (ESI): m/z calcd for C15H19NO4 [M+] 293.1263; found: 293.1262.
5-(Tert-butyl) 1-methyl (E)-2-[(2-bromobenzylidene)amino]pentanedioate (2f): Pale yellow oil (196.6 mg, 88%, 91% purity). IR (neat) νmax: 1727, 1149, 752, 686 cm−1. 1H NMR (300 MHz) δ 8.65 (s, 1H, HC=N), 8.08 (dd, J = 7.6, 2.0 Hz, 1H, ArH), 7.60–7.55 (m, 1H, ArH), 7.38–7.28 (m, 2H, ArH), 4.19–4.08 (m, 1H, NCHCO2Me), 3.76 (s, 3H, CH3), 2.37–2.22 (m, 4H, CH2-CH2), 1.44 [s, 9H, (CH3)3]. 13C NMR (101 MHz) δ 172.02 (C=O2tBu), 171.85 (CO2Me), 163.16 (CH=N), 134.01 (CAr), 133.03 (CHAr), 132.37 (CHAr), 129.27 (CHAr), 127.63(CHAr), 125.36 (BrCAr), 80.54 (C(CH3)3), 71.83(NCHCO2Me), 52.29 (CH3), 31.49 (CH2), 28.45 (CH2), 28.11 (C(CH3)3). MS (EI) m/z: 384 (M+, 1%), 329 (47), 327 (48), 312 (38), 310 (38), 276 (93), 269 (44), 268 (100), 224 (41), 222 (37), 184 (41), 89 (50), 57 (66). HRMS (ESI): m/z calc for C17H22BrNO4 [M+] 383.073; found: 385.0694.
5-(Tert-butyl) 1-methyl (E)-2-[(3-bromobenzylidene)amino]pentanedioate (2g): Pale yellow oil (215.3 mg, 93%, 92% purity). IR (neat) νmax: 1727, 1369, 1149, 752, 682 cm−1. 1H NMR (400 MHz) δ 8.22 (s, 1H, HC=N), 7.98 (t, J = 1.8 Hz, 1H, ArH), 7.66–7.62 (m, 1H, ArH), 7.58–7.54 (m, 1H, ArH), 7.28 (t, J = 7.8 Hz, 1H, ArH), 4.09–4.04 (m, 1H, NCHCO2Me), 3.74 (s, 3H, CO2CH3), 2.30–2.15 (m, 4H, CH2-CH2), 1.43 [s, 9H, CO2(CH3)3]. 13C NMR (101 MHz) δ 172.0 (C=O), 171.8 (C=O), 162.5 (CH=N), 137.4 (CAr), 134.1 (CHAr), 131.0 (CHAr), 130.1 (CHAr), 127.5 (CHAr), 122.9 (BrCAr), 80.5 (C(CH3)3), 71.7 (NCHCO2Me), 52.3 (CO2CH3), 31.5 (CH2), 28.4 (CH2), 28.1 [(CH3)3]. MS (EI) m/z: 384 (M+, 7%), 329 (43), 327 (45), 310 (38, 268 (100), 224 (43), 222 (40), 184 (42), 116 (36), 89 (50), 57 (93). HRMS (ESI): m/z calc for C17H22BrNO4 [M+] 383.073; found: 383.0737.
Dimethyl (E)-2-[(4-bromobenzylidene)amino]pentanedioate (2h): Pale yellow oil (151.3 mg, 88%, 93% purity). IR (neat) νmax: 1731, 2116, 1068, 1010, 821, 732 cm−1. 1H NMR (300 MHz) δ 8.24 (s, 1H, HC=N), 7.68–7.61 (m, 2H, ArH), 7.59–7.52 (m, 2H, ArH), 4.15–3.98 (m, 1H, NCHCO2Me), 3.75 (s, 3H, CO2CH3), 3.65 (s, 3H, CO2CH3), 2.47–2.14 (m, 4H, CH2-CH2). 13C NMR (101 MHz) δ 173.24 (C=O), 171.69 (C=O), 163.07 (C=N), 134.39 (BrCAr), 131.9, 129.9 (CHAr), 125.8 (CAr), 71.6 (CH), 52.3, 51.6 (CH3), 30.1, 28.2 (CH2). MS (EI) m/z: 342 (M+, 4%), 282 (87), 268 (55), 222 (100), 184 (54), 143 (42), 116 (40), 89 (88). HRMS (ESI): m/z calcd for C14H16BrNO4 [M+] 341.0263; found: 341.0265.
5-(Tert-butyl) 1-methyl (E)-2-[(4-bromobenzylidene)amino]pentanedioate (2i): Pale yellow oil (139.1 mg, 93%, 93% purity). IR (neat) νmax: 1727, 1369, 1149, 825 cm−1. 1H NMR (300 MHz) δ 8.23 (s, 1H, HC=N), 7.65 (d, J = 8.5 Hz, 2H, ArH), 7.56 (d, J = 8.5 Hz, 2H, ArH), 4.10–4.02 (m, 1H, NCHCO2Me), 3.75 (s, 3H, CO2CH3), 2.34–2.15 (m, 4H, CH2-CH2), 1.43 (s, 9H, C(CH3)3). 13C NMR (101 MHz) δ 172.2 (C=O), 172.0 (C=O), 163.0 (C=N), 134.5 (BrCAr), 132.0 (CHAr), 130.1 (CHAr), 125.9 (CAr), 80.6 [C(CH3)3], 71.9 (CH), 52.4 (CO2CH3), 31.7 (CH2), 28.6 (CH2), 28.2 [CO2C(CH3)3]. MS (EI) m/z: 383(M+, 3%), 327 (49), 312 (42), 268 (100), 224 (48), 184 (40), 89 (63), 57(92). HRMS (ESI): m/z calc for C17H22BrNO4 [M+] 383.0730; found: 383.0720.
Dimethyl (E)-2-[(pyridin-3-ylmethylene)amino]pentanedioate (2j): Pale yellow oil (214.7 mg, 93%, 90% purity). IR (neat) νmax: 1731, 1434, 1172, 806, 709 cm−1. 1H NMR (300 MHz) δ 8.86 (dd, J = 2.1, 0.7 Hz, 1H, ArH), 8.65 (dd, J = 4.8, 1.7 Hz, 1H, ArH), 8.32 (s, 1H, HC=N), 8.15 (dt, J = 7.9, 1.9 Hz, 1H, ArH), 7.34 (dd, J = 7.9, 4.8 Hz, 1H, ArH), 4.10 (dd, J = 7.7, 5.0 Hz, 1H, NHCO2Me), 3.73 (s, 3H, CH3), 3.63 (s, 3H, CH3), 2.42–2.19 (m, 4H, CH2-CH2). 13C NMR (101 MHz) δ 173.2 (C=O), 171.5 (C=O), 161.6 (C=N), 152.1 (CHAr), 150.6 (CHAr), 134.9 (CHAr), 131.1 (CAr), 123.6 (CHAr), 71.7 (CH), 52.3 (CH3), 51.7 (CH3), 30.1 (CH2), 28.2 (CH2). MS (EI) m/z: 264 (M+, 7%), 205 (74), 204 (43), 191 (26), 145 (100), 118 (27), 105 (38). HRMS (ESI): m/z calcd for C13H16N2O4 [M+] 264.111; found: 264.1108.
5-(Tert-butyl) 1-methyl (E)-2-[(pyridin-3-ylmethylene)amino]pentanedioate (2k): Pale yellow oil (241.7 mg, 83%, 89% purity). IR (neat) νmax: 1727, 1369, 1253, 1153, 802, 709 cm−1. 1H NMR (300 MHz) δ 8.88–8.86 (m, 1H, ArH), 8.66 (dd, J = 4.8, 1.7 Hz, 1H, ArH), 8.32 (s, 1H, HC=N), 8.16 (dt, J = 7.9, 1.9 Hz, 1H, ArH), 7.34 (dd, J = 7.7, 4.8 Hz, 1H, ArH), 4.13–4. 06 (m, 1H, NHCO2Me), 3.74 (s, 3H, CH3), 2.38–2.10 (m, 4H, CH2-CH2), 1.42 (s, 9H, [(CH3)3]. 13C NMR (75 MHz) δ 171.9 (C=O), 171.67(C=O), 161.3 (C=N), 152.0 (CHAr), 150.6 (CHAr), 134.9 (CHAr), 131.1 (CAr), 123.6 (CHAr), 80.5 [C(CH3)3], 71.8 (CH), 52.2 (CO2CH3), 31.4 (CH2), 28.3 (CH2), 28.0 [C(CH3)3]. MS (EI) m/z: 306 (M+, 1%), 250 (69), 233 (38), 191 (100), 145 (54), 118 (28), 105 (22), 57 (33). HRMS (ESI): m/z calcd for C16H22N2O4 [M+] 306.158; found: 306.1579.
5-(Tert-butyl) 1-methyl (E)-2-[(thien-2-ylmethylene)amino]pentanedioate (2l): Pale yellow oil (147.2 mg, 90%, 91% purity). IR (neat) νmax: 1727, 1627, 1249, 1211, 1153, 713 cm−1. 1H NMR (400 MHz) δ 8.39 (s, 1H, HC=N), 7.46 (d, J = 5.0 Hz, 1H, Thienyl-H), 7.38 (dd, J = 3.6, 1.0 Hz, 1H, Thienyl-H), 7.10 (dd, J = 4.9, 3.7 Hz, 1H, Thienyl-H), 4.04 (dd, J = 8.5, 4.1 Hz, 1H, NCHCO2Me), 3.75 (s, 3H, CO2CH3), 2.43–2.16 (m, 4H, CH2-CH2), 1.45 [s, 9H, C(CH3)3]. 13C NMR (101 MHz) δ 172.1, 171.9 (C=O), 157.1 (C=N), 141.6 (SCC=N), 131.6, 129.8, 127.4 (ArC), 80.4 [C(CH3)3], 71.4 (NCH), 52.2 (CH3), 31.6, 28.3 (CH2), 28.1 [C(CH3)3]. MS (EI) m/z: 311 (M+, 6%), 255 (93), 238 (64), 195 (100), 194 (95), 150 (71), 123 (35), 112 (44), 110 (29), 96 (44), 57 (51), 43 (36), 41 (23). HRMS (ESI): m/z calcd for C15H21NO4S[M+] 311.1196; found: 311.119.
5-Butyl 1-methyl (E)-2-(benzylideneamino)-2-methylpentanedioate (2m): Pale yellow oil (152 mg, 95%, 86% purity). IR (neat) vmax: 1729, 1643, 1452, 1378, 1174, 1114, 730, 694 cm−1. 1H NMR (400 MHz) δ 8.24 (s, 1H, HC=N), 7.76–7.71 (m, 2H, ArH), 7.40–7.36 (m, 3H ArH), 4.02 (m, J = 6.7, 0.8 Hz, 2H, CO2CH2CH2CH2CH3), 3.71 (s, 3H, CO2CH3), 2.45 (m, J = 9.6, 7.9, 5.7 Hz, 2H, NCH3CH2CH2), 2.33 (m, J = 13.7, 10.1, 5.6 Hz, 1H, NCH3CH2CH2), 2.19–2.10 (m, 1H, NCH3CH2CH2), 1.55 (m, J = 8.2, 7.0, 6.0 Hz, 2H, CO2CH2CH2CH2CH3), 1.49 (s, 3H, NCCH3), 1.39–1.26 (m, 2H, CO2CH2CH2CH2CH3), 0.89 (t, J = 7.4 Hz, 3H, CO2CH2CH2CH2CH3). 13C NMR (101 MHz) δ 174.14 (C=O), 173.53 (C=O), 159.81 (C=N), 136.27 (CAr), 130.97 (CHAr), 128.53 (CHAr x2), 128.35 (CHAr x2), 67.52 (CH), 64.32 (CH2), 52.21 (CH3), 35.06 (CH2), 30.64 (CH2), 29.64 (CH2), 23.40 (CH3), 19.11 (CH2), 13.69(CH3). MS (EI) m/z: 319 (M+, >1%), 260 (100), 158 (38). HRMS (ESI): m/z calcd for C17H22NO4 [M+–CH3] 304.1549; found: 304.1545.
Dimethyl (E)-2-[(4-bromobenzylidene)amino]-2-methylpentanedioate (2n): Pale yellow oil (110.2 mg, 90%, 90% purity). IR (neat) νmax: 1731, 1438, 1245, 1172, 1114, 821 cm−1. 1H NMR (300 MHz) δ 8.22 (s, 1H, HC=N), 7.65–7.59 (m, 2H, ArH), 7.57–7.51 (m, 2H, ArH), 3.74 (s, 3H, CH3), 3.64 (s, 3H, CH3), 2.58–2.26 (m, 4H, CH2-CH2), 1.50 (s, 3H, CH3). 13C NMR (101 MHz) δ 173.9 (C=O), 173.8 (C=O), 158.7 (C=N), 135.1 (CAr), 131.8 (2xCHAr), 129.78 (2xCHAr), 125.4 (BrCAr), 67.5 (NCCO2Me), 52.3 (CH3), 51.6 (CH3), 35.0 (CH2), 29.4 (CH2), 23.3 (CH3CN). MS (EI) m/z: 356 (M+, <1%), 298 (97), 296 (100), 236 (27), 184 (19), 89 (32). HRMS (ESI): m/z calc for C13H15BrNO2 [M+–CO2CH3] 296.0286; found: 296.0286.
5-Butyl 1-methyl (E)-2-[(4-bromobenzylidene)amino]-2-methylpentanedioate (2o): Pale yellow oil (191 mg, 96%, 91% purity). IR (neat) Vmax: 1729, 1643, 1438, 1170, 1114, 1066, 821, 744 cm−1. 1H NMR (400 MHz) δ 8.18 (s, 1H, NH), 7.58 (ddd, J = 8.4, 5.8, 2.7 Hz, 2H, ArH), 7.49 (dt, J = 12.4, 4.7 Hz, 2H, ArH), 4.00 (td, J = 6.7, 2.9 Hz, 2H, CO2CH2CH2CH2CH3), 3.68 (s, 3H CH3), 2.41 (dqt, J = 9.1, 6.5, 3.2 Hz, 2H, CH2CO2nBu), 2.36–2.23 (m, 1H, NCCH2), 2.12 (dddd, J = 12.0, 8.3, 6.0, 2.3 Hz, 1H, NCCH2), 1.53 (dt, J = 12.7, 6.8 Hz, 2H, CO2CH2CH2CH2CH3), 1.45 (s, 3H, CH3), 1.39–1.24 (m, 2H, CO2CH2CH2CH2CH3), 0.87 (tt, J = 6.2, 3.2 Hz, 3H, CH3).13C NMR (101 MHz) δ 173.86 (C=O), 173.38 (C=O), 158.61 (HC=N), 135.16 (CAr), 131.73 (CHAr x2), 129.76 (CHAr x2), 125.37 (CAr), 67.60 (C), 64.32 (CH2), 52.24 (CH3), 35.01 (CH2), 30.62 (CH2), 29.62 (CH2), 23.36 (CH3), 19.10 (CH2), 13.68 (CH3)). MS (EI) m/z: 398 (M+, >1%), 338 (100), 238 (31). HRMS (ESI): m/z calcd for C17H21BrNO4 [M+–CH3] 382.0654; found: 382.0647.
5-(Tert-butyl) 1-methyl (E)-2-[(4-bromobenzylidene)amino]-2-methylpentanedioate (2p): Pale yellow oil (104.5 mg, 77%, 88% purity). IR (neat) νmax: 2981, 1727, 1643, 1589, 1249, 1153, 1114, 821 cm−1. 1H NMR (400 MHz) δ 8.20 (s, 1H, HC=N), 7.64–7.59 (m, 2H, ArH), 7.55–7.50 (m, 2H, ArH), 3.72 (s, 3H, CH3), 2.45–2.34 (m, 2H, CH2), 2.32–2.22 (m, 2H, CH2), 1.48 (s, 3H, CH3), 1.42 [s, 9H, C(CH3)3]. 13C NMR (75 MHz) δ 174.0 (C=O), 172.6 (C=O), 158.5 (C=N), 135.2 (CAr), 131.7 (2xCHAr), 129.8 (2xCHAr), 125.3 (BrCAr), 80.3 (C(CH3)3), 67.6 (NCCO2Me), 52.2 (CO2CH3), 34.9 (CH2), 30.7 (CH2), 28.0 [C(CH3)3], 23.3 (CH3). MS (EI) m/z: 398 (M+, <1%), 284 (97), 282 (100), 268 (100), 238 (20), 236 (20), 184 (20), 160 (48), 89 (35), 57 (31), 43(24). HRMS (ESI): m/z calc for C12H13BrNO2 [M+-C6H11O2] 284.0102; found: 284.0102.
Dimethyl (E)-2-benzyl-2-[(4-bromobenzylidene)amino]pentanedioate (2q): Pale yellow oil (201.2 mg, 90%, 90% purity). IR (neat) νmax: 1731, 1438, 1172, 1083, 821, 740, 701 cm−1. 1H NMR (300 MHz) δ 7.97 (s, 1H, HC=N), 7.60–7.51 (m, 4H, ArH), 7.24–7.14 (m, 5H, ArH), 3.74 (s, 3H, CH3), 3.62 (s, 3H, CH3), 3.35–3.10 (m, 2H, CH2Ar), 2.45–2.23 (m, 4H, CH2-CH2). 13C NMR (75 MHz) δ 173.7 (C=O), 172.8 (C=O), 160.5 (C=N), 135.8 (CAr), 135.1 (CArCH2), 131.8 (2xCHAr), 130.6 (2xCHAr), 129.7 (2xCHAr), 128.1 (2xCHAr), 126.9 (CHAr), 125.4 (BrCAr), 71.6 (NCCO2Me), 52.0 (CH3), 51.6 (CH3), 44.3 (PhCH2), 32.9 (CH2), 29.5 (CH2). MS (EI) m/z: 432 (M+, <1%), 372 (18), 342 (98), 340 (100), 254 (25), 91 (45). HRMS (ESI): m/z calc for C19H19BrNO2 [M+–CO2CH3] 372.0599; found: 372.0605.
Methyl (E)-2-[(4-bromobenzylidene)amino]-5-(dimethylamino)-5-oxopentanoate (2u): Pale yellow oil (153.1 mg 90%, 86% purity). IR (neat) νmax: 1735, 1639, 1203, 1168, 1064, 821 cm−1. 1H NMR (300 MHz) δ 8.24 (s, 1H, HC=N), 7.65–7.59 (m, 2H, ArH), 7.55–7.49 (m, 2H, ArH), 4.19–4.13 (m, 1H, NCHCO2Me), 3.71 (s, 3H, CO2CH3), 2.93 (s, 3H, NCH3), 2.89 (s, 3H, NCH3), 2.40–2.29 (m, 4H, CH2-CH2). 13C NMR (101 MHz) δ 172.1 (C=O), 171.9 (C=O), 162.9 (CH=N), 134.5 (CAr), 131.8 (CHAr), 129.9 (CHAr), 125.6 (BrCAr), 71.6 (NCHCO2Me), 52.2 (CO2CH3), 37.1 (NCH3), 35.9 (NCH3), 28.9 (CH2), 28.7 (CH2). MS (EI) m/z: 354 (M+, 5%), 270 (30), 268 (32), 224 (35), 222 (31), 173 (24), 100 (26), 89 (31), 87 (100), 72 (35), 45 (25). HRMS (ESI): m/z calc for C15H19BrN2O3 [M+] 354.0579; found: 354.0565.
Methyl (E)-2-[(4-bromobenzylidene)amino]-4-cyano-2-(2-cyanoethyl)butanoate (6v): Pale yellow oil (117.0 mg, 90%, 84% purity). IR (neat) νmax: 1727, 1369, 1149, 825 cm−1. 1H NMR (300 MHz) δ 8.24 (s, 1H, HC=N), 7.65–7.59 (m, 2H, ArH), 7.55–7.49 (m, 2H, ArH), 4.19–4.13 (m, 1H, NCHCO2Me), 3.71 (s, 3H, CO2CH3), 2.93 (s, 3H, NCH3), 2.89 (s, 3H, NCH3), 2.40–2.29 (m, 4H, CH2-CH2). 13C NMR (101 MHz) δ 172.1 (C=O), 171.9 (C=O), 162.9 (CH=N), 134.5 (CAr), 131.8 (CHAr), 129.9 (CHAr), 125.7 (BrCAr), 71.6 (NCHCO2Me), 52.2 (CO2CH3), 37.1 (NCH3), 35.3 (NCH3), 28.9 (CH2), 28.7 (CH2). MS (EI) m/z: 354 (M+, 5%), 270 (30), 268 (32), 224 (35), 222 (31), 173 (24), 100 (26), 89 (31), 87 (100), 72 (35), 45 (25). HRMS (ESI): m/z calc for C15H19BrN2O3 [M+] 354.0579; found: 354.0565.

3.3. General Procedure for the Synthesis of Pyroglutamate Derivatives

To a solution of NaBH4 (0.8 mmol, 2 eq) in Methanol (4 mL) at 0 °C, a solution of corresponding adduct 2 in Methanol (2mL) was added and the reaction was refluxed at 80 °C for 2 h. After that, the solvent was removed and the crude product was redissolved in AcEOt (4 mL) followed by the addition of SiO2 and refluxed again for 24 h. Then, the mixture was filtered and removed from the solvent under vacuum. Finally, the crude product was purified by flash column chromatography on silica gel (Hexane/AcEOt, 3:1) to afford the corresponding cycloadducts 8.
Methyl 5-oxo-1-(pyridin-3-ylmethyl)pyrrolidine-2-carboxylate (8a): Pale yellow oil (93 mg, 60%). IR (neat) νmax: 1689, 1411, 1211, 1025, 794, 721 cm−1. 1H NMR (300 MHz) δ 8.48 (dd, J = 4.7, 1.2 Hz, 1H, ArH), 8.41 (d, J = 1.2 Hz, 1H, ArH), 7.54 (dt, J = 7.8, 2.0 Hz, 1H, ArH), 7.27–7.18 (m, 1H, ArH), 4.90 (d, J = 15.1 Hz, 1H, CH2NC=O), 4.04 (d, J =15.1 Hz, 1H, CH2NC=O), 3.94 (dd, J = 9.0, 3.2 Hz, 1H, NCHCO2Me), 3.62 (s, 3H, CH3), 2.56–2.45 (m, 1H, CH2), 2.43–2.34 (m, 1H, CH2), 2.27–2.17 (m, 1H, CH2), 2.11–1.99 (m, 1H, CH2). 13C NMR (101 MHz) δ 175.2 (NC=O), 171.9 (C=O), 149.6 (NCHAr), 149.3 (NCHAr), 136.3 (CHAr), 131.6 (CAr), 123.7 (CHAr), 58.8 (NCHCO2), 52.6 (CH3), 43.2 (CH2N), 29.3 (CH2), 22.8 (CH2). MS (EI) m/z: 234 (M+, 14%), 175 (98), 92 (100), 65(17). HRMS (ESI): m/z calc for C12H14N2O3 [M+] 234.1004; found: 234.0998.
Methyl 1-(4-bromobenzyl)-2-methyl-5-oxopyrrolidine-2-carboxylate (8b): Pale yellow oil (49.1 mg, 42%). IR (neat) νmax: 2360, 1735, 1689, 1392, 1168 cm−1. 1H NMR (400 MHz) δ 7.44–7.38 (m, 2H, ArH), 7.19–7.11 (m, 2H, ArH), 4.52 (d, J = 15.6 Hz, 1H, ArCH2N), 4.29 (d, J = 15.5 Hz, 1H, ArCH2N), 3.51 (s, 3H, CO2CH3), 2.57 (dt, J = 17.0, 9.8 Hz, 1H, NC=OCH2), 2.46 (ddd, J = 17.0, 9.7, 2.7 Hz, 1H, NC=OCH2), 2.34 (ddd, J = 12.9, 9.3, 2.7 Hz, 1H, NCCH2), 1.90 (dt, J = 13.2, 9.9 Hz, 1H, NCCH2), 1.41 (s, 3H, CH3). 13C NMR (101 MHz) δ 175.8 (C=O), 173.7 (C=O), 136.6 (CAr), 131.8 (2xCHAr), 129.9 (2xCHAr), 121.2 (BrCAr), 66.0 (NCCH3), 52.4 (CO2CH3), 43.9 (ArCH2N), 32.1 (C=OCH2), 29.5 (CCH2), 23.2 (NCCH3). MS (EI) m/z: 325 (M+, 7%), 268 (72), 266 (73), 171 (95), 169 (100), 90 (28), 89 (22). HRMS (ESI): m/z calc for C14H16BrNO3 [M+] 325.0314; found: 325.0306.
Methyl 2-benzyl-1-(4-bromobenzyl)-5-oxopyrrolidine-2-carboxylate (8c): Colorless needles (77 mg, 63%). Mp: 78–79 ºC (n-hexane/AcOEt). IR (neat) νmax: 1689, 1392, 1261, 1184, 1068, 806, 732, 701 cm−1. 1H NMR (300 MHz) δ 7.46–7.40 (m, 2H, ArH), 7.30–7.24 (m, 3H, ArH), 7.23–7.17 (m, 2H, ArH), 7.10–7.04 (m, 2H, ArH), 4.81 (d, J = 15.5 Hz, 1H, HCN), 4.40 (d, J = 15.5 Hz, 1H, HCN), 3.38 (s, 3H, CH3), 3.27 (d, J = 14.0 Hz, 1H, NCCHPh), 3.00 (d, J = 14.0 Hz, 1H, NCCHPh), 2.46–2.23 (m, 2H, NCOCH2), 2.10–1.84 (m, 2H, NCCH2). 13C NMR (75 MHz) δ 176.0 (C=O), 172.2 (C=O), 136.5 (CAr), 134.4 (CAr), 131.4 (2xCHAr), 130.0 (2xCHAr), 129.9 (2xCHAr), 128.6 (2xCHAr), 127.3 (CHAr), 121.2 (BrCAr), 69.2 (NCCO2Me), 52.2 (CH3), 44.0 (PhCH2N), 40.4 (CCH2Ph), 29.2 (CH2C=O), 27.5 (NCCH2). MS (EI) m/z: 402 (M+, 1%), 312 (67), 310 (68), 171 (97), 169 (100), 91 (22), 90 (27). HRMS (ESI): m/z calc for C20H20BrNO3 [M+] 401.0627; found: 401.0619.
Methyl 2-[(1H-indol-3-yl)methyl]-1-(4-bromobenzyl)-5-oxopyrrolidine-2-carboxylate (8d): Purple oil (105.6 mg, 45%). IR (neat) vmax: 1735, 1673, 1434, 1394, 1257, 1070, 736 cm−1. 1H NMR (400 MHz) δ 8.29 (s, 1H, NH), 7.56 (dt, J = 7.7, 1.0 Hz, 1H, ArH), 7.46–7.41 (m, 2H, ArH), 7.35 (dt, J = 8.1, 1.0 Hz, 1H, ArH), 7.27–7.11 (m, 4H, ArH), 6.77 (s, 1H, CHNH), 4.85 (d, J = 15.3 Hz, 1H, H2CN), 4.36 (d, J = 15.2 Hz, 1H, H2CN), 3.52 (d, J = 15.2 Hz, 1H, CH2CCHNH), 3.37 (s, 3H, CO2CH3), 3.18 (d, J = 15.3 Hz, 1H, CH2CCHNH), 2.44 (ddd, J = 16.2, 9.8, 5.5 Hz, 1H, NCOCH2), 2.30–2.20 (m, 1H, NCOCH2CH2), 2.13–1.97 (m, 2H, NCOCH2, NCOCH2CH2). 13C NMR (101 MHz) δ 176.7 (C=O), 173.1 (C=O), 136.4 (CAr), 135.9 (CAr), 131.6 (CHAr), 130.4 (CHAr), 128.4 (CAr), 123.2 (CHAr), 122.4 (CHAr), 121.5 (CAr), 119.9 (CHAr), 118.9 (CHAr), 111.5 (CHAr), 108.5 (CAr), 69.7 (C), 52.3 (CH3), 44.2 (CH2), 29.7 (CH2), 29.6 (CH2), 27.9 (CH2). MS (EI) m/z: 441 (M+, >1%), 310 (21), 252 (26), 168 (100), 126 (20), 90 (26). HRMS (ESI): m/z calcd for C22H21BrN2O3 [M+] 440.0736; found: 440.0712.
Methyl 2-[(benzyloxy)methyl]-1-(4-bromobenzyl)-5-oxopyrrolidine-2-carboxylate (8e): Colorless liquid (107.2 mg, 46%) IR (neat) vmax: 1739, 1693, 1432, 1392, 1166, 1070, 736, 698 cm−1. 1H NMR (400 MHz) δ 7.57–7.23 (m, 5H, ArH), 7.23–7.01 (m, 4H, ArH), 4.59 (d, J = 15.5 Hz, 1H, H2CN), 4.34 (d, J = 15.5 Hz, 1H, H2CN), 4.23 (q, J = 12.0 Hz, 2H, CCH2OCH2Ph), 3.68 (d, J = 10.0 Hz, 1H, CCH2OCH2Ph), 3.58 (s, 3H, CO2CH3), 3.46 (d, J = 10.0 Hz, 1H, CCH2OCH2Ph), 2.55–2.45 (m, 2H, CH2C=O), 2.20 (m, J = 13.2, 7.3, 5.8 Hz, 1H, CH2CCO2Me), 2.10 (m, J = 13.1, 9.5 Hz, 1H CH2CCO2Me). 13C NMR (101 MHz) δ 176.3 (C=O), 172.3 (C=O), 137.3 (CAr), 137.1 (CAr), 131.3 (CHAr), 129.9 (CHAr), 128.5 (CHAr), 128.0 (CHAr), 127.7 (CHAr), 121.0 (C), 73.4 (CH2), 71.2 (CH2), 69.6 (C), 52.6 (CH3), 44.7 (CH2), 29.3 (CH2), 27.1 (CH2). MS (EI) m/z: 432 (M+, 2%), 310 (66), 169 (100), 91 (73). HRMS (ESI): m/z calcd for C21H22BrNO4 [M+] 431.0732; found: 433.0723.
Methyl 1-(4-bromobenzyl)-2-[2-(methylthio)ethyl]-5-oxopyrrolidine-2-carboxylate (8f): Pale yellow oil (65.8 mg, 42%). IR (neat) vmax: 1733, 1689, 1432, 1390, 1160, 1070, 732 cm−1. 1H NMR (400 MHz) δ 7.41 (d, J = 8.4 Hz, 2H, ArH), 7.16 (d, J = 8.4 Hz, 2H ArH), 4.51–4.29 (m, 2H, H2CN), 3.45 (s, 3H, CO2CH3), 2.61 (dt, J = 17.0, 9.6 Hz, 1H, CCH2CH2SCH3), 2.53–2.36 (m, 2H, CCH2CH2SCH3, CCH2CH2SCH3), 2.35–2.25 (m, 2H, CCH2CH2SCH3, NCCH2CH2CO), 2.23–2.10 (m, 1H, NCCH2CH2CO), 2.01 (s, 3H, SCH3), 1.98–1.87 (m, 2H, NCCH2CH2CO, NCCH2CH2CO). 13C NMR (101 MHz) δ 176.1 (C=O), 172.8 (C=O), 136.2 (CAr), 131.6 (CHAr x 2), 130.3 (CHAr x 2), 121.6 (CAr), 68.5 (C), 52.5 (CH3), 44.1 (CH2), 35.0 (CH2), 29.6 (CH2), 28.0 (CH2), 27.9 (CH2), 15.7 (CH3). MS (EI) m/z: 386 (M+, 13%), 328 (64), 169 (100), 90 (23), 61 (30). HRMS (ESI): m/z calcd for C16H20BrNO3S [M+] 385.0347; found: 387.0328.

3.4. Procedure to Obtain the Glutamic Acid 1,5-Dimethyl Ester Hydrochloride 9 [70]

To a solution of 2a (1 eq, 0.33mmol) in Et2O (0.7 mL), 2M HCl/Et2O (0.35 mL) was added and stirred until a precipitate was observed, then the solvent was removed under vacuum. The crude product was purified, washed with Et2O (3 times), and the supernatant was removed. The remaining solid was characterized (62 mg, 90%). Mp 91–92 °C (89–91 °C) [70]. 1H NMR (300 MHz, Methanol-d4) δ 4.12 (t, J = 6.8 Hz, 1H), 3.84 (s, 3H), 3.70 (s, 3H), 2.58 (td, J = 7.3, 2.1 Hz, 2H), 2.30–2.07 (m, 2H). 13C NMR (101 MHz, Methanol-d4) δ 172.6, 169.2, 52.3, 51.8, 51.0, 28.7, 25.1.

4. Conclusions

The 1,4-addition of imino esters, derived from the condensation of amino esters and aldehydes, was a complicated task due to the formation of three secondary byproducts. The use of dimethylphenylphosphine as an organocatalyst was crucial for the control of the desired reaction, circumventing all of the undesired molecules and favoring the α-alkylated compound, which did not undergo the retro-Michael process. To the best of our knowledge, this is the first occasion that phosphines are involved in this particular transformation with the advantage of the minimization of the over-alkylation of the glycine template with alkyl acrylates. The reaction is very versatile because it tolerates many aromatic and heteroaromatic units bonded to the imino group, as well as substituents at the α-position of the imino ester. This methodology does not employ benzene as a solvent and the crude reaction materials were not impurified with variable amounts (10–25%) of the corresponding pyrrolidines (formed by 1,3-dipolar cycloadditions), such as occurred in a previous communication [16]. In addition, no sophisticated halogenated compounds were employed as electrophiles [15] and the presence of strong bases was avoided [17,18,19,20,21]. The access to different synthetic glutamates was ensured and the family of the corresponding pyroglutamate derivatives was successfully obtained by simple organic transformations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29020342/s1, The NMR and representative FTIR copies.

Author Contributions

Conceptualization, J.M.S.; methodology, L.V.R.-F., E.G.-M., E.S. and M.K.; validation, E.G.-M. and M.d.G.R.; formal analysis, L.V.R.-F., M.G.-M., E.G.-M. and M.d.G.R.; investigation, L.V.R.-F., E.G.-M., E.G.-M., E.S. and M.K.; resources, M.d.G.R. and J.M.S.; writing—original draft preparation, J.M.S.; writing—review and editing, M.d.G.R. and J.M.S.; supervision, M.d.G.R. and J.M.S.; project administration, M.d.G.R. and J.M.S.; funding acquisition, M.d.G.R. and J.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge financial support from the Spanish Ministerio de Ciencia, Innovación y Universidades (RED2022-134287-T ORFEO CINQA and RED2022-134331-T CASI) the Spanish Ministerio de Economía, Industria y Competitividad, Agencia Estatal de Investigación (AEI) and Fondo Europeo de Desarrollo Regional (FEDER, EU) (projects CTQ2017-82935-P and PID2019-107268GB-I00), the Generalitat Valenciana (IDIFEDER/2021/013, GVA-COVID19/2021/079 and CIDEGENT/2020/058), Medalchemy S. L. (Medalchemy-22T) and the University of Alicante (VIGROB-068, UAUSTI21-05). LV.R.-F.-F thanks Generalitat Valenciana for Grisolía’s fellowship (GRISOLIAP/2020/111).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References and Notes

  1. Eder, I.; Haider, V.; Zebrowski, P.; Waser, M. Recent Progress in the Asymmetric Syntheses of α-Heterofunctionalized (Masked) α- and β-Amino Acid Derivatives. Eur. J. Org. Chem. 2021, 2021, 202–219. [Google Scholar] [CrossRef]
  2. Yuan, Z.; Liu, X.; Liu, C.; Zhang, Y.; Rao, Y. Recent Advances in Rapid Synthesis of Non-proteinogenic Amino Acids from Proteinogenic Amino Acids Derivatives via Direct Photo-Mediated C–H Functionalization. Molecules 2020, 25, 5270. [Google Scholar] [CrossRef] [PubMed]
  3. Masamba, W. Petasis vs. Strecker Amino Acid Synthesis: Convergence, Divergence and Opportunities in Organic Synthesis. Molecules 2021, 26, 1707. [Google Scholar] [CrossRef] [PubMed]
  4. Zhou, M.; Feng, Z.; Zhang, X. Recent advances in the synthesis of fluorinated amino acids and peptides. Chem. Commun. 2023, 59, 1434–1448. [Google Scholar] [CrossRef] [PubMed]
  5. Shatskiy, A.; Kaerkaes, M.D. Photoredox-enabled decarboxylative synthesis of unnatural α-amino acids. Synlett 2022, 33, 109–115. [Google Scholar] [CrossRef]
  6. Forro, E.; Fulop, F. Enzymatic Strategies for the Preparation of Pharmaceutically Important Amino Acids through Hydrolysis of Amino Carboxylic Esters and Lactams. Curr. Med. Chem. 2022, 29, 6218–6227. [Google Scholar] [CrossRef]
  7. Larionov, V.A.; Stoletova, N.V.; Maleev, V.I. Advances in asymmetric amino acid synthesis enabled by radical chemistry. Adv. Synth. Cat. 2020, 362, 4325–4367. [Google Scholar] [CrossRef]
  8. Saghyan, A.S.; Langer, P. Asymmetric Synthesis of Non-Proteinogenic Amino Acid; Wiley-VCH: Weinheim, Germany, 2016. [Google Scholar]
  9. O’Donnell, M.J. Benzophenone Schiff bases of glycine derivatives: Versatile starting materials for the synthesis of amino acids and their derivatives. Tetrahedron 2019, 75, 3667–3696. [Google Scholar] [CrossRef]
  10. Xue, Y.-P.; Cao, C.-H.; Zheng, Y.-G. Enzymatic asymmetric synthesis of chiral amino acids. Chem. Soc. Rev. 2018, 47, 1516–1561. [Google Scholar] [CrossRef]
  11. Wang, Y.; Song, X.; Wang, J.; Moriwaki, H.; Soloshonok, V.A.; Liu, H. Recent approaches for asymmetric synthesis of α-amino acids via homologation of Ni(II) complexes. Amino Acids 2017, 49, 1487–1520. [Google Scholar] [CrossRef]
  12. He, G.; Wang, B.; Nack, W.A.; Chen, G. Syntheses and Transformations of α-Amino Acids via Palladium-Catalyzed Auxiliary-Directed sp3 C-H Functionalization. Acc. Chem. Res. 2016, 49, 635–645. [Google Scholar] [CrossRef] [PubMed]
  13. Nájera, C.; Sansano, J.M. Catalytic Asymmetric Synthesis of α-Amino Acids. Chem. Rev. 2007, 107, 4584–4671. [Google Scholar] [CrossRef] [PubMed]
  14. Yin, G.-W.; Wu, S.-L.; Yan, J.-H.; Zhang, P.-F.; Yang, M.-M.; Li, L.; Xu, Z.; Yang, K.-F.; Xu, L.-W. Swollen-induced in-situ encapsulation of chiral silver catalysts in cross-linked polysiloxane elastomers: Homogeneous reaction and heterogeneous separation. Mol. Catal. 2021, 515, 111901. [Google Scholar] [CrossRef]
  15. Seibel, Z.M.; Bandar, J.S.; Lambert, T.H. Enantioenriched α-substituted glutamates/pyroglutamates via enantioselective cyclopropenimine-catalyzed Michael addition of amino ester imines. Beilst. J. Org. Chem. 2021, 17, 2077–2084. [Google Scholar] [CrossRef]
  16. Grigg, R.; Kemp, J.; Malone, J.P.; Rajviroongit, S.; Tangthonkum, A. X=Y-ZH Systems as potential 1,3-dipoles. Part 17. Sequential Michael addition-5-endo-trig cyclisation of arylideneimines of α-amino acid esters. Tetrahedron 1988, 44, 5361–5374. [Google Scholar] [CrossRef]
  17. Bai, Y.-J.; Cheng, M.L.; Zheng, X.-H.; Zhang, S.-Y.; Wang, P.A. Chiral Cyclopropenimine-catalyzed Asymmetric Michael Addition of Bulky Glycine Imine to α,β-Unsaturated Isoxazoles. chiral organosuperbase catalyst. Chem. Asian J. 2022, 17, e202200131. [Google Scholar] [CrossRef] [PubMed]
  18. Leonardi, C.; Brandolese, A.; Preti, L.; Bortolini, O.; Polo, E.; Dambruoso, P.; Ragno, D.; Di Carmine, G.; Massi, A. Expanding the Toolbox of Heterogeneous Asymmetric Organocatalysts: Bifunctional Cyclopropenimine Superbases for Enantioselective Catalysis in Batch and Continuous Flow. Adv. Synth. Catal. 2021, 363, 5473–5485. [Google Scholar] [CrossRef]
  19. Tyszka-Gumkowska, A.; Jurczak, J. Divergent synthesis of pyrrolidine and glutamic acid derivatives using a macrocyclic phase-transfer catalyst under high-pressure conditions. Org. Chem. Front. 2021, 8, 5888–5894. [Google Scholar] [CrossRef]
  20. Bai, Y.-J.; Hu, X.-M.; Bai, Y.-J.; Zheng, X.-H.; Zhang, S.-Y.; Ping-An, W. DBU-catalyzed Michael addition of bulky glycine imine to α,β-unsaturated isoxazoles and pyrazolamides. Tetrahedron 2021, 101, 132511. [Google Scholar] [CrossRef]
  21. Meninno, S.; Carratu, M.; Overgaard, J.; Lattanzi, A. Diastereoselective Synthesis of Functionalized 5-Amino-3,4-Dihydro-2H-Pyrrole-2-Carboxylic Acid Esters: One-Pot Approach Using Commercially Available Compounds and Benign Solvents. Chem. Eur. J. 2021, 27, 4573–4577. [Google Scholar] [CrossRef]
  22. O’Donnell, M.J.; Boniece, J.M.; Earp, S.E. The synthesis of amino acids by phase-transfer reaction. Tetrahedron Lett. 1978, 19, 2641–2644. [Google Scholar] [CrossRef]
  23. Duan, X.-Y.; Tian, Z.; Liu, B.; He, T.; Zhao, L.-L.; Dong, M.; Zhang, P.; Qi, J. Highly Enantioselective Synthesis of Pyrroloindolones and Pyrroloquinolinones via an N-Heterocyclic Carbene-Catalyzed Cascade Reaction. Org. Lett. 2021, 23, 3777–3781. [Google Scholar] [CrossRef] [PubMed]
  24. Ren, N.; Zhang, L.; Hu, Y.; Wang, X.; Deng, Z.; Chen, J.; Deng, H.; Zhang, H.; Tang, X.-J.; Cao, W. Perfluoroalkyl-Promoted Synthesis of Perfluoroalkylated Pyrrolidine-Fused Coumarins with Methyl β-Perfluoroalkylpropionates. J. Org. Chem. 2021, 86, 15717–15725. [Google Scholar] [CrossRef] [PubMed]
  25. Xiao, F.; Xu, S.M.; Dong, X.-Q.; Wang, C.-J. Allylations Ir-Catalyzed Asymmetric Tandem Allylation/Iso-Pictet−Spengler Cyclization Reaction for the Enantioselective Construction of Tetrahydro-γ-carbolines. Org. Lett. 2021, 23, 706–710. [Google Scholar] [CrossRef] [PubMed]
  26. Barlow, S.R.; Callaghan, L.J.; Franckevicius, V. Investigation of the palladium-catalysed cyclisation of α-amido malonates with propargylic compounds. Tetrahedron 2021, 80, 131866. [Google Scholar] [CrossRef]
  27. Cui, Z.; Zhang, K.; Gu, L.; Bu, Z.; Zhao, J.; Wang, Q. Diastereoselective trifunctionalization of pyridinium salts to access structurally crowded azaheteropolycycles. Chem. Commun. 2021, 57, 9402–9405. [Google Scholar] [CrossRef]
  28. Ha, M.W.; Lee, M.; Choi, S.; Kim, S.; Hong, S.; Park, Y.; Kim, M.; Kim, T.-S.; Lee, J.; Lee, J.K.; et al. Construction of Chiral α-Amino Quaternary Stereogenic Centers via Phase-Transfer Catalyzed Enantioselective α-Alkylation of α-Amidomalonates. J. Org. Chem. 2015, 80, 3270–3279. [Google Scholar] [CrossRef]
  29. Li, J.-T.; Luo, J.-N.; Wang, J.-L.; Wang, D.-K.; Yu, Y.-Z.; Zhuo, C.-X. Stereoselective intermolecular radical cascade reactions of tryptophans or ɤ-alkenyl-α-amino acids with acrylamides via photoredox catalysis. Nat. Commun. 2022, 13, 1778–1785. [Google Scholar] [CrossRef]
  30. Banerjee, I.; Ghosh, K.C.; Oheix, E.; Jean, M.; Naubron, J.V.; Réglier, M.; Iranzo, O.; Sinha, S. Synthesis of Protected 3,4- and 2,3-Dimercaptophenylalanines as Building Blocks for Fmoc-Peptide Synthesis and Incorporation of the 3,4-Analogue in a Decapeptide Using Solid-Phase Synthesis. J. Org. Chem. 2021, 86, 2210–2223. [Google Scholar] [CrossRef]
  31. Liao, H.-S.; Chung, Y.-H.; Hsieh, M.-H. Glutamate: A multifunctional amino acid in plants. Plant Sci. 2022, 318, 111238. [Google Scholar] [CrossRef]
  32. Bayer, T.A. Pyroglutamate Aβ cascade as drug target in Alzheimer’s disease. Mol. Psychiatry 2022, 27, 1880–1885. [Google Scholar] [CrossRef]
  33. Lin, T.-Y.; Lu, C.-W.; Hsieh, P.-W.; Chiu, K.-M.; Lee, M.-Y.; Wang, S.-J. Natural Product Isoliquiritigenin Activates GABAB Receptors to Decrease Voltage-Gate Ca2+ Channels and Glutamate Release in Rat Cerebrocortical Nerve Terminals. Biomolecules 2021, 11, 1537. [Google Scholar] [CrossRef]
  34. Gulder, T.A.M.; Moore, B.S. Salinosporamide Natural Products: Potent 20 S Proteasome Inhibitors as Promising Cancer Chemotherapeutics. Angew. Chem. Int. Ed. 2010, 49, 9346–9367. [Google Scholar] [CrossRef] [PubMed]
  35. Panday, S.K.; Prasad, J.; Dikshit, D.K. Pyroglutamic acid: A unique chiral synthon. Tetrahedron Asymmetry 2009, 20, 1581–1632. [Google Scholar] [CrossRef]
  36. Zhang, J.; Flippen-Anderson, J.L.; Kozikowski, A.P. A Tandem Michael Addition Ring-Closure Route to the Metabotropic Receptor Ligand α-(Hydroxymethyl)glutamic Acid and Its γ-Alkylated Derivatives. J. Org. Chem. 2001, 66, 7555–7559. [Google Scholar] [CrossRef] [PubMed]
  37. Tekkam, S.; Alam, M.A.; Just, M.J.; Berry, S.M.; Johnson, J.L.; Jonnalagadda, S.C.; Mereddy, V.R. Stereoselective Synthesis of Pyroglutamate Natural Product Analogs from α-Aminoacids and their Anti-Cancer Evaluation. Anti-Cancer Agents Med. Chem. 2013, 13, 1514–1530. [Google Scholar] [CrossRef] [PubMed]
  38. Katoh, M.; Hisa, C.; Honda, T. Enantioselective synthesis of (R)-deoxydysibetaine and (−)-4-epi-dysibetaine. Tetrahedron Lett. 2007, 48, 4691–4694. [Google Scholar] [CrossRef]
  39. Isaacson, J.; Loo, M.; Kobayashi, Y. Total Synthesis of (±)-Dysibetaine. Org. Lett. 2008, 10, 1461–1463. [Google Scholar] [CrossRef]
  40. Ma, G.; Nguyen, H.; Romo, D. Concise Total Synthesis of (±)-Salinosporamide A, (±)-Cinnabaramide A, and Derivatives via a Bis-cyclization Process:  Implications for a Biosynthetic Pathway? Org. Lett. 2007, 9, 2143–2146. [Google Scholar] [CrossRef]
  41. Ling, T.; Macherla, V.R.; Manam, R.R.; McArthur, K.A.; Potts, B.C.M. Enantioselective Total Synthesis of (−)-Salinosporamide a (NPI-0052). Org. Lett. 2007, 9, 2289–2292. [Google Scholar] [CrossRef]
  42. Shibasaki, M.; Kanai, M.; Fukuda, N. Total Synthesis of Lactacystin and Salinosporamide A. Chem. Asian J. 2007, 2, 20–38. [Google Scholar] [CrossRef]
  43. Masse, C.E.; Morgan, A.J.; Adams, J.; Panek, J.S. Syntheses and Biological Evaluation of (+)-Lactacystin and Analogs. Eur. J. Org. Chem. 2000, 2000, 2513–2528. [Google Scholar] [CrossRef]
  44. Bai, X.-F.; Li, L.; Xu, Z.; Zheng, Z.-J.; Xia, C.G.; Cui, Y.M.; Xu, L.-W. Asymmetric Michael-addition of aldimino esters with chalcones catalyzed by Silver/Xing-Phos: Mechanism-oriented divergent synthesis of chiral pyrrolines. Chem. Eur. J. 2016, 22, 10399–10404. [Google Scholar] [CrossRef]
  45. Guo, H.; Fan, Y.C.; Sun, Z.; Wu, Y.; Kwon, O. Phosphine Organocatalysis. Chem. Rev. 2018, 118, 10049–10293. [Google Scholar] [CrossRef] [PubMed]
  46. Molteni, G.; Silvani, A. Spiro-2-oxindoles via 1,3-dipolar cycloadditions. A decade update. Eur. J. Org. Chem. 2021, 2021, 1653–1675. [Google Scholar] [CrossRef]
  47. Nájera, C.; Sansano, J.M. Synthesis of pyrrolizidines and indolizidines by multicomponent 1,3-dipolar cycloaddition of azomethine ylides. Pure Appl. Chem. 2019, 91, 575–596. [Google Scholar] [CrossRef]
  48. Dondas, H.A.; Retamosa, M.G.; Sansano, J.M. Current trends towards the synthesis of bioactive heterocycles and natural products using 1, 3-dipolar cycloadditions (1, 3-DC) with azomethine ylides. Synthesis 2017, 49, 2819–2851. [Google Scholar] [CrossRef]
  49. Bdiri, B.; Zhao, B.-J.; Zhou, Z.-M. Recent advances in the enantioselective 1,3-dipolar cycloaddition of azomethine ylides and dipolarophiles. Tetrahedron Asymmetry 2017, 28, 876–899. [Google Scholar] [CrossRef]
  50. Wei, L.; Chang, X.; Wang, C.-J. Catalytic Asymmetric Reactions with N-Metallated Azomethine Ylides. Acc. Chem. Res. 2020, 53, 1084–1100. [Google Scholar] [CrossRef]
  51. Adrio, J.; Carretero, J.C. Stereochemical diversity in pyrrolidine synthesis by catalytic asymmetric 1,3-dipolar cycloaddition of azomethine ylides. Chem. Commun. 2019, 55, 11979–11991. [Google Scholar] [CrossRef]
  52. Fang, X.; Wang, C.-J. Catalytic asymmetric construction of spiropyrrolidines via 1,3-dipolar cycloaddition of azomethine ylides. Org. Biomol. Chem. 2018, 16, 2591–2601. [Google Scholar] [CrossRef]
  53. De, N.; Yoo, E.J. Recent Advances in the Catalytic Cycloaddition of 1,n-Dipoles. ACS Catal. 2018, 8, 48–58. [Google Scholar] [CrossRef]
  54. Stewart, I.C.; Bergman, R.G.; Toste, F.D. Phosphine-Catalyzed Hydration and Hydroalkoxylation of Activated Olefins:  Use of a Strong Nucleophile to Generate a Strong Base. J. Am. Chem. Soc. 2003, 125, 8696–8697. [Google Scholar] [CrossRef] [PubMed]
  55. Rundlöf, T.; Mathiasson, M.; Bekiroglu, S.; Hakkarainen, B.; Bowden, T.; Arvidsson, T. Survey and qualification of internal standards for quantification by 1 H NMR Spectroscopy. J. Pharm. Biomed. Anal. 2010, 52, 645–651. [Google Scholar] [CrossRef] [PubMed]
  56. Purities of compounds were determined accordingly. See experimental section of this article.
  57. The addition of a substituent at the β-position in the acrylate did not give the expected reaction.
  58. Zhang, X.; Zhang, J.; Jia, M.; Peng, L.; Zhang, N.; Qi, S.; Zhang, L. Bifunctional additive phenyl vinyl sulfone for boosting cyclability of lithium metal batteries. Green. Chem. Engin. 2023, 4, 49–56. [Google Scholar] [CrossRef]
  59. Fahim, A.M.; Ghabbour, H.A.; Kabil, M.M.; Al-Rashood, S.T.; Abdel-Aziz, H.A. Synthesis, X-ray crystal structure, Hirshfeld analysis and computational investigation of bis(methylthio)acrylonitrile with antimicrobial and docking evaluation. J. Mol. Struct. 2022, 1260, 132793. [Google Scholar] [CrossRef]
  60. Ishihara, M.; Fujisawa, S. A structure-activity relationship study on the mechanisms of methacrylate-induced toxicity using NMR chemical shift of β-carbon, RP-HPLC log P and semiempirical molecular descriptor. Dental Mat. J. 2009, 28, 113–120. [Google Scholar] [CrossRef]
  61. Pérez-Garmendia, R.; Gevorkian, G. Pyroglutamate-Modified Amyloid Beta Peptides: Emerging Targets for Alzheimer’s Disease Immunotherapy. Curr. Neuropharm. 2013, 11, 491–498. [Google Scholar] [CrossRef]
  62. Cynis, H.; Frost, J.L.; Crehan, H.; Lemere, C.A. Immunotherapy targeting pyroglutamate-3 Aβ: Prospects and challenges. Mol. Neurodegener. 2016, 11, 48/1–48/11. [Google Scholar] [CrossRef]
  63. Piechotta, A.; Parthier, C.; Kleinschmidt, M.; Gnoth, K.; Pillot, T.; Lues, I.; Demuth, H.-U.; Schilling, S.; Rahfeld, J.-U.; Stubbs, M.T. Structural and functional analyses of pyroglutamate-amyloid-β-specific antibodies as a basis for Alzheimer immunotherapy. J. Biol. Chem. 2017, 292, 12713–12724. [Google Scholar] [CrossRef]
  64. Xu, C.; Wang, Y.-n.; Wu, H. Glutaminyl cyclase, diseases, and development of glutaminyl cyclase inhibitors. J. Med. Chem. 2021, 64, 6549–6565. [Google Scholar] [CrossRef]
  65. Casas, J.; Grigg, R.; Nájera, C.; Sansano, J.M. The Effect of Phase-Transfer Catalysis in the 1,3-Dipolar Cycloaddition Reactions of Azomethine Ylides− Synthesis of Substituted Prolines Using AgOAc and Inorganic Base in Substoichiometric Amounts. Eur. J. Org. Chem. 2001, 2001, 1971–1982. [Google Scholar] [CrossRef]
  66. Caleffi, G.; Larrañaga, O.; Ferrándiz-Saperas, M.; Costa, P.; Nájera, C.; de Cózar, A.; Cossío, F.; Sansano, J.M. Switching Diastereoselectivity in Catalytic Enantioselective (3+2) Cycloadditions of Azomethine Ylides Promoted by Metal Salts and Privileged Segphos-Derived Ligands. J. Org. Chem. 2019, 84, 10593–10605. [Google Scholar] [CrossRef]
  67. Beksultanova, N.; Doğan, Ö. Asymmetric synthesis of aryl-substituted pyrrolidines by using CFAM ligand–AgOAc chiral system via 1,3-dipolar cycloaddition reaction. Chirality 2023, 35, 435–448. [Google Scholar] [CrossRef] [PubMed]
  68. Moins, S.; Coulembier, O. Dimerization of Methyl Acrylate through CO2-pressurized DBU Mediated Process. Asian J. Org. Chem. 2022, 11, e202100734. [Google Scholar] [CrossRef]
  69. Yu, B.; Bai, X.-F.; Ly, J.-Y.; Yuan, Y.; Cao, J.; Zheng, Z.-J.; Xu, Z.; Cui, Y.-M.; Yang, K.F.; Xu, L.-W. Enantioselective Synthesis of Chiral Imidazolidine Derivatives by Asymmetric Silver/Xing-Phos-Catalyzed Homo-1,3-Dipolar [3+2] Cycloaddition of Azomethine Ylides. Adv. Synth. Cat. 2017, 359, 3577–3584. [Google Scholar] [CrossRef]
  70. Weigl, M.; Wunsch, B. Synthesis of chiral non-racemic 3-(dioxopiperazin-2-yl)propionic acid derivatives. Tetrahedron 2002, 58, 1173–1183. [Google Scholar] [CrossRef]
Scheme 1. (ac) Strategies for the preparation of glutamate derivatives.
Scheme 1. (ac) Strategies for the preparation of glutamate derivatives.
Molecules 29 00342 sch001
Scheme 2. Optimization of the Michael-type addition of imino ester 1a onto methyl acrylate (MA).
Scheme 2. Optimization of the Michael-type addition of imino ester 1a onto methyl acrylate (MA).
Molecules 29 00342 sch002
Scheme 3. Mechanism of the formation of products 2a, 3a, 4 and 5.
Scheme 3. Mechanism of the formation of products 2a, 3a, 4 and 5.
Molecules 29 00342 sch003
Scheme 4. Scope of the Michael-type addition of imino esters 1 onto alkyl acrylates.
Scheme 4. Scope of the Michael-type addition of imino esters 1 onto alkyl acrylates.
Molecules 29 00342 sch004
Scheme 5. Michael-type addition of imino ester 1g onto AN and PVS.
Scheme 5. Michael-type addition of imino ester 1g onto AN and PVS.
Molecules 29 00342 sch005
Scheme 6. Synthesis of pyroglutamate derivatives 8. (a) From compounds 2. (b) From compounds 1 without isolating intermediate molecules 2.
Scheme 6. Synthesis of pyroglutamate derivatives 8. (a) From compounds 2. (b) From compounds 1 without isolating intermediate molecules 2.
Molecules 29 00342 sch006
Scheme 7. Synthesis of glutamate derivative 9.
Scheme 7. Synthesis of glutamate derivative 9.
Molecules 29 00342 sch007
Table 1. Optimization conditions of the reaction shown in Scheme 2.
Table 1. Optimization conditions of the reaction shown in Scheme 2.
Slow Addition Products’ Ratio a
EntryPR3 (10 mol%)(min)t (h)2a3a451a
1PPh3---15090091
2dppe---1537220437
3dppe---48223203016
4dppe---725639050
5dppe1a (60)728002000
6dppeMA (60)728800120
7dppeMA (60)127100920
8Bu3P---720001000
9Bu3PMA (60)720010000
10But3P---720010000
11Me2PhP---1223038380
12Me2PhP---270015150
13Me2PhPMA (60)27050520
14Me2PhPMA (60)248010055
15Me2PhPMA (30)24970003
16Me2PhPMA (5)246402907
17Me2PhP bMA (5)245000050
a The ratio was estimated according to integral values in crude product. 1H NMR and also using dimethyl terephthalate as the internal standard. b A 5 mol% of Me2PhP was added.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rodríguez-Flórez, L.V.; González-Marcos, M.; García-Mingüens, E.; Retamosa, M.d.G.; Kawase, M.; Selva, E.; Sansano, J.M. Phosphine Catalyzed Michael-Type Additions: The Synthesis of Glutamic Acid Derivatives from Arylidene-α-amino Esters. Molecules 2024, 29, 342. https://doi.org/10.3390/molecules29020342

AMA Style

Rodríguez-Flórez LV, González-Marcos M, García-Mingüens E, Retamosa MdG, Kawase M, Selva E, Sansano JM. Phosphine Catalyzed Michael-Type Additions: The Synthesis of Glutamic Acid Derivatives from Arylidene-α-amino Esters. Molecules. 2024; 29(2):342. https://doi.org/10.3390/molecules29020342

Chicago/Turabian Style

Rodríguez-Flórez, Lesly V., María González-Marcos, Eduardo García-Mingüens, María de Gracia Retamosa, Misa Kawase, Elisabet Selva, and José M. Sansano. 2024. "Phosphine Catalyzed Michael-Type Additions: The Synthesis of Glutamic Acid Derivatives from Arylidene-α-amino Esters" Molecules 29, no. 2: 342. https://doi.org/10.3390/molecules29020342

Article Metrics

Back to TopTop