Next Article in Journal
Energetic and Geometric Characteristics of the Substituents: Part 2: The Case of NO2, Cl, and NH2 Groups in Their Mono-Substituted Derivatives of Simple Nitrogen Heterocycles
Next Article in Special Issue
Enantioseparation of P-Stereogenic 1-Adamantyl Arylthiophosphonates and Their Stereospecific Transformation to 1-Adamantyl Aryl-H-phosphinates
Previous Article in Journal
Unravelling the Anticancer Mechanisms of Traditional Herbal Medicines with Metabolomics
Previous Article in Special Issue
Asymmetric Synthesis of Stereogenic Phosphorus P(V) Centers Using Chiral Nucleophilic Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Novel Crown Ether-Squaramides and Their Application as Phase-Transfer Catalysts

1
Department of Organic Chemistry and Technology, Budapest University of Technology and Economics, Szent Gellért tér 4, H-1111 Budapest, Hungary
2
Department of Inorganic and Analytical Chemistry, Budapest University of Technology and Economics, Szent Gellért tér 4, H-1111 Budapest, Hungary
3
MS Proteomics Research Group, Research Centre for Natural Sciences, Hungarian Academy of Sciences, Magyar Tudósok körútja 2, H-1117 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(21), 6542; https://doi.org/10.3390/molecules26216542
Submission received: 15 September 2021 / Revised: 18 October 2021 / Accepted: 27 October 2021 / Published: 29 October 2021
(This article belongs to the Special Issue New Insights into the Preparation and Separation of Enantiomers)

Abstract

:
This work presents the synthesis of six new phase-transfer organocatalysts in which the squaramide unit is directly linked to the nitrogen atom of an aza-crown ether. Four chiral skeletons, namely hydroquinine, quinine, cinchonine (cinchonas), and α-d-glucopyranoside were responsible for the asymmetric construction of an all-carbon quaternary stereogenic center in α-alkylation and Michael addition reactions of malonic esters. We investigated the effects of these different chiral units and that of crown ethers with different sizes on catalytic activity and enantioselectivity. During extensive parameter investigations, both conventional and emerging green solvents were screened, providing valuable α,α-disubstituted malonic ester derivatives with excellent yields (up to 98%).

Graphical Abstract

1. Introduction

Investigating the creation of quaternary stereogenic centers is essential in organic synthesis as it poses a challenge to organic chemists due to the possible steric repulsion between the groups around the stereocenter. In the last decade, the synthesis of chiral α,α-disubstituted malonates has gained considerable attention [1,2]. These compounds can be applied in the preparation of α,α-disubstituted amino acids, as they can undergo chemoselective transformations if their carboxylic acid moieties are protected with different groups, e.g., through Curtius rearrangement [3]. A decade ago, a new method involving enantioselective phase-transfer catalytic double α-alkylation of malonates was developed for the construction of chiral quaternary carbon centers, applying tert-butyl diphenylmethyl α-alkylmalonates as starting materials [1]. This method was later extended in numerous publications by examining a broad substrate scope [3,4,5,6]. Moreover, similar phase-transfer catalytic reactions, such as the α-benzoyloxylation of tert-butyl methyl α-alkylmalonates [7] and the Michael addition reaction of tert-butyl methyl α-benzylmalonate to acrylates, have also been elaborated [2]. In all of these papers, quaternary ammonium salts were applied as phase-transfer catalysts (PTCs) in the enantioselective α-alkylation of malonates, while chiral macrocycle PTCs remain a rather neglected tool for these asymmetric reactions.
Phase-transfer catalysis is one of the most efficient asymmetric synthetic methods, which is also inexpensive and sustainable as it involves simple procedures and mild reaction conditions. Thanks to these advantages, phase-transfer catalysis is particularly suitable for industrial applications [8].
The most common types of asymmetric phase-transfer catalysts are chiral quaternary ammonium and phosphonium salts, but chiral crown ether derivatives and other macrocycles emerge as alternatives despite their cumbersome and costly synthesis [9,10,11]. Crown ethers are neutral ligands that can complex and transport alkali metal cations into the organic phase. This crown ether–metal cation complex plays the same role as the quaternary onium cation, but crown ethers work with a different mechanism, called cation-binding catalysis, during which the entire reacting ion pair is transported into the organic phase, not just the anion [12]. The most conspicuous benefits that help crown ether derivatives to stand out from other catalysts are as follows: they are usually more resistant to strong bases, and the cation is usually more accessible than the positively charged nitrogen or phosphorus atom in ammonium or phosphonium salts; consequently, a stronger interaction can take place with the reactive anion. Furthermore, crown ethers are more effective in extracting inorganic salts from their solid form [13].
In asymmetric phase-transfer catalysis, introducing hydrogen bond donor units into catalysts has recently attained broad application [14]. The most common of those units include the hydroxyl group, amide, (thio)urea and squaramide. In the case of quaternary onium salts, there are many examples for the installation of these units into catalyst scaffold [15,16,17,18,19,20]. However, in crown ether derivatives, only hydroxyl groups are frequently used as ancillary components capable of forming hydrogen bonds.
Squaramides are highly effective double hydrogen bond donor units due to their rigid, aromatic four-membered ring [21]. There are a few examples in which a squaramide unit was connected indirectly to crown ether derivatives [22,23,24,25,26,27]. However, these ligands have only been applied for ion pair transport so far.
Herein, we present the synthesis of novel chiral crown ether-squaramide phase-transfer organocatalysts, and the catalytic performance of these new derivatives were tested in the synthesis of α,α-disubstituted malonates. The effects of different chiral units and various cavity sizes of crown ethers on catalytic activity and enantioselectivity were investigated. In these catalysts, the squaramide unit is directly linked to the nitrogen atom of an aza-crown ether; thus, one amino group of the squaramide unit is tertiary. This means that the squaramide unit serves as a linker and a single hydrogen bond donor unit [28,29]. By applying this catalyst design, we have anticipated that one NH group and the cation complexed by the crown ether can form secondary interactions of sufficient strength during the catalysis.

2. Results and Discussion

2.1. Synthesis of Crown Ether-Squaramide Phase-Transfer Catalysts

We have prepared a series of crown ethers with a squaramide hydrogen bond donor unit (Scheme 1) for catalyzing the α-alkylation reaction of malonates, which may be an important method for obtaining α,α-disubstituted α-amino acid intermediates. Different cinchona alkaloids [quinine (Q), hydroquinine (HQ) and cinchonine (C)] and a d-glucose derivative (G) were chosen as chiral starting materials for the syntheses of the catalysts. Previously, we have shown that the substituent on the nitrogen atom of glucose-based aza-crown ether G influences the catalytic effect [30].
The cinchona alkaloid-based catalysts were prepared in 3 steps. The C9 hydroxyl group of the commercially available starting materials (Q, HQ, C) was converted to an amino group as reported earlier [31]. In the next step, half squaramide derivatives quinine-half squaramide (Q-HSQ), hydroquinine-half squaramide (HQ-HSQ) and cinchonine-half squaramide (C-HSQ) were obtained by the addition of these amines to dimethyl squarate (DMSQ) [32,33]. Finally, the addition of 1-aza-15-crown-5 or 1-aza-18-crown-6 ethers to the previously mentioned half squaramides afforded the corresponding cinchona-crown ether-squaramide derivatives (Q5, Q6, C5, C6, HQ5).
The glucose-based catalyst was prepared in a two-step-synthesis from glucose-aza-crown ether derivative G, which can be prepared in 5 steps from d-glucose [34,35]. First, dimethyl squarate (DMSQ) was reacted with 3,5-bis(trifluoromethyl)aniline to obtain a half squaramide derivative HSQ [36]; then, the treatment of aza-crown ether G with this half squaramide (HSQ) led to the appropriate glucose-crown ether-squaramide derivative (G5).
During the synthetic procedures, all compounds were characterized by well-established methods, including HRMS, IR, 1H and 13C NMR spectroscopies.

2.2. Application of the Catalysts

The catalysts were applied in the asymmetric α-alkylation of tert-butyl methyl α-benzylmalonate [3] (1) under phase-transfer conditions (Scheme 2).
First, we compared the activities and enantioselectivities of the catalysts; these results are shown in Table 1. The catalysts provided good yields except catalysts bearing a 1-aza-18-crown-6 ether macroring (Q6 and C6, Table 1, Entries 5–6), and only poor or no enantioselectivity. The highest enantiomeric excess of (R)-3 was obtained with catalyst C5, which was still as low as 9% (Table 1, Entry 3). In the absence of a base while applying C5 as catalyst, no product was observed according to thin-layer chromatography (TLC) analysis. When 50% aq. NaOH was applied as a base in the absence of a phase-transfer catalyst, the yield was only 33%.
Next, we conducted a parameter study with catalyst C5 to maximize the enantioselectivity. First, we examined the effect of the solvent while applying 50% aq. NaOH or solid NaOH as base (Table 2). In the case of aqueous NaOH, six different solvents were tested (Table 2, Entries 1–6); however, only the most suitable solvent (DCM: dichloromethane; Table 2, Entry 7) and two other polar aprotic solvents (MeCN; THF: tetrahydrofuran; Table 2, Entries 8–9) were used with solid NaOH. Dichloromethane proved to be the best solvent in terms of yield and enantiomeric excess by applying 50% aq. NaOH as base (Table 2, Entry 1, 73% yield, 9% ee).
Then, the effect of the quantity of the applied base or its concentration was investigated. By decreasing the quantity of the 50% aq. NaOH from 50 to 25 eq, the yield slightly increased (from 73% to 85%), but the enantiomeric excess did not change. By using 1 eq of 50% aq. NaOH or 50 eq of 30% aq. NaOH, both yield and enantioselectivity declined (from 73% yield and 9% ee to 26–28% yield and <5% ee).
In the next step of the parameter study, we examined the effect of the catalyst quantity (10 or 20 mol%), the reagent quantity (1 or 1.2 eq) and the solvent volume (1.2 or 2.4 mL) (Table 3). This set of experiments indicated that the application of 10 mol% catalyst, 1.2 eq of allyl bromide and 2.4 mL DCM is the most beneficial as it gave (R)-3 in an increased yield (83%) with a slightly better enantiomeric excess (11%, Table 3, Entry 3).
The reaction was also carried out using allyl iodide instead of allyl bromide as a reagent under the same conditions as at the beginning of the study (1 eq reagent, 10 mol% C5 catalyst, 50 eq 50% aq. NaOH base, DCM solvent, 25 °C, 24 h). In this case, a better yield (98% instead of 73%) and a slightly higher, but still low, enantiomeric excess (12% instead of 9%) were obtained. Consequently, we continued the study by applying allyl iodide.
Next, the reaction was carried out at 0 °C as an endeavor to improve enantioselectivity. As increasing the quantity of the reagent to 1.2 eq and decreasing the quantity of the base (50% aq. NaOH) to 25 eq seemed to be able to ameliorate yield and enantiomeric excess, we also varied these conditions in the following experiments (Table 4). By setting a lower reaction temperature, the enantiomeric excess did not improve significantly in any case. Based on these experiments, the yield was found to deteriorate when only 25 eq base is applied, while it does not depend on the temperature and the reagent quantity in the applied ranges. We also conducted the reaction at −78 °C to examine whether the enantioselectivity could be improved by setting a drastically lower reaction temperature, but no product (3) was formed based on TLC analysis.
Furthermore, the use of other bases did not improve the outcome: with 2 eq Na2CO3 in THF or MeCN solvent, there was no product at 0 °C, and by applying 50% aq. NaOH base in the cases of Q6 and C6, and 50% aq. KOH in the cases of Q5, HQ5, C5 and G5 catalysts (to examine the effect of both bases with each catalyst), enantioselective alkylation occurred only with C5 catalyst at 0 °C (17% ee, but only 47% yield, with 50% aq. KOH).
As another attempt to increase enantioselectivity, we also investigated the activity and enantioselectivity of catalyst C5 in Michael addition reaction of tert-butyl methyl α-benzylmalonate (1) to benzyl acrylate (Scheme 3). We anticipated higher enantioselectivity than in the alkylation reaction as benzyl acrylate has a hydrogen bond acceptor carbonyl oxygen, which could form stronger interactions with the hydrogen bond donor groups of the catalyst as opposed to the allyl halide reagents in the alkylation reaction. However, there was practically no enantioselectivity (2% ee) in this reaction, although product 5 was obtained in a good yield (74%).

2.3. Proposed Mechanism

Based on our previous research, we propose a mechanism for the α-alkylation reaction of tert-butyl methyl α-benzylmalonate (1) in the presence of our crown ether-squaramide catalysts. According to our preceding study, the Na+ ion of multiple monosaccharide-based aza-crown ethers’ Na+ complexes can be coordinated by a substrate’s carbonyl group [37]. In our other work, we have found that it is possible that only one NH group of a cinchona-based organocatalyst’s squaramide unit forms a hydrogen bond with a substrate’s carbonyl group [29]. Extending these results to our catalysts, we suggest that after the deprotonation of malonate 1 by the hydroxide ion, the phase-transfer catalyst coordinates the malonate anion in a way that one carbonyl of the malonate binds to the Na+ ion and the other to the NH group of the squaramide (Scheme 4). Then, the nucleophilic α-carbon of the malonate attacks the allyl halide, either directly on the electrophilic saturated carbon or in a conjugate addition on the unsaturated carbon, resulting in allylic rearrangement. Naturally, in the case of the applied unsubstituted allyl halides, the product is the same in both cases. Thus, the halide anion leaves, and an allyl group is built into the substrate.

3. Materials and Methods

3.1. General

Starting materials were purchased from commercially available sources (Sigma-Aldrich, Merck, and Alfa Aesar). Infrared spectra were recorded on a Bruker Alpha-T FT-IR spectrometer (Bruker, Ettlingen, Germany). Optical rotations were measured on a Perkin-Elmer 241 polarimeter (Perkin-Elmer, Waltham, MA, USA) that was calibrated by measuring the optical rotations of both enantiomers of menthol. Silica gel 60 F254 (Merck) plates were used for TLC. Silica gel 60 (70–230 mesh, Merck) was used for column chromatography. Ratios of solvents for the eluents are given in volumes (mL mL−1). Melting points were taken on a Boetius micro-melting point apparatus (VEB Dresden Analytik, Dresden, Germany), and they were uncorrected.
NMR spectra were recorded on a Bruker Avance 500 MHz spectrometer( Bruker, Ettlingen, Germany), (500.13 MHz for 1H, 125.76 MHz for 13C and 50.68 MHz for 15N) in DMSO-d6 and were referenced to residual solvent proton signals(δH = 2.50) and solvent carbon signals (δC = 39.51). The approximate 15N chemical shifts were detected by 1H,15N-gs-HMBC method optimized for J(N,H) = 5 Hz long-range couplings, and 15N chemical shifts are given on NH3 scale. The pulse programs of one-dimensional (1H, 13C and DEPTQ) and two-dimensional (1H,1H-gs-COSY, 1H,13C-gs-HSQC, 1H,13C-gs-HMQC, 1H,13C-gs-HMBC, 1H,15N-gs-HMBC and 1H,1H-gs-ROESY) measurements were utilized. All chemical shifts are reported in parts per million (ppm). Abbreviations used in the description of resonances are: a (methylene hydrogen with higher 1H chemical shift), b (methylene hydrogen with lower 1H chemical shift), o (overlapping signal), s (singlet), d (doublet), t (triplet), q (quartet), br (broad), m (multiplet) and nd: not detectable. Coupling constants (J) are given in Hz.
The broad signals, the non-integer integral values in the 1H NMR spectra (seen in Supplementary Materials) of compounds and the duplication of some signals in the 1H and 13C NMR spectra (e.g., 1H and 13C NMR signals of N-methylene groups in aza-crown ether units) of compounds indicate inhibited motions in the molecule. Also, the conformational properties of the molecules resulted in many cases in which not all signals appeared in the NMR spectra due to signal broadening. In these cases, the approximate chemical shift of the above-mentioned signals could be determined based on their HSQC or HMBC correlations. Previous 1H NMR measurements of similar compounds at higher temperatures showed that not all conformational motion inhibitions were removed.
In the process of structure elucidation, we assigned the 1H and 13C NMR signals of the main conformer with the exception of compound G5, in which the chemical shifts of several minor signals were also determined.
The starting points of signal assignment were easily identifiable units of molecules: the methyl and methoxy groups, the terminal olefinic methylene unit, the position 2′ in the quinoline ring and the methylidine groups next to nitrogen. The remainder of the molecular structures was elucidated using the same NMR methods mentioned above.
HPLC-MS was performed using a Shimadzu LCMS-2020 (Shimadzu Corp., Kyoto, Japan) device, equipped with a Reprospher (Altmann Analytik Corp., München, Germany) 100 C18 (5 µm; 100 × 3 mm) column and a positive/negative double ion source (DUIS±) with a quadrupole MS analyzer in a range of 50–1000 m/z. The samples were eluted with gradient elution, using eluent A (0.1% HCOOH in H2O) and eluent B (0.1% HCOOH in MeCN). The flow rate was set to 1.5 mL min−1. The initial condition was 5% eluent B, followed by a linear gradient to 100% eluent B by 1.5 min; from 1.5 to 4.0 min, 100% eluent B was retained; and from 4 to 4.5 min, it went back by a linear gradient to 5% eluent B, which was retained from 4.5 to 5 min. The column temperature was kept at room temperature, and the injection volume was 1 µL. The purity of the compounds was assessed by HPLC with UV detection at 215 and 254 nm. High-resolution mass measurements were performed on a Q-TOF Premier mass spectrometer (Waters Corp., Milford, MA, USA). The ionization method was electrospray ionization (ESI) operated in positive ion mode. The enantiomeric excess (ee) values were determined by chiral HPLC with a PerkinElmer Series 200 instrument (PerkinElmer, Waltham, MA, USA). The applied column, eluent, flow rate, column temperature, and detector wavelength are indicated at the corresponding procedure.

3.2. Procedures

3.2.1. 3-(+)-Methoxy-4-(((R)-quinolin-4-yl((1S,2R,4S,5R)-5-vinylquinuclidin-2-yl)methyl)amino)cyclobut-3-ene-1,2-dione (C-HSQ)

A solution of C-N (535 mg, 1.55 mmol) in DCM (2 mL) was added to a solution of dimethyl squarate (DMSQ) (242 mg, 1.70 mmol) in DCM (2 mL) under argon atmosphere. This mixture was stirred for 3 h at room temperature. The solvent was removed under reduced pressure, then the crude product was purified by column chromatography (SiO2; DCM:methanol = 10:1) to obtain C-HSQ as pale-yellow crystals (494 mg, 79% yield). TLC (SiO2 TLC; DCM:methanol = 5:1, Rf = 0.62). M.p. 122 °C (decomposed). [ α ] D 25 : +220.0 (c = 1.0, chloroform).
IR (KBr): 3233, 3075, 2938, 2869, 1803, 1706, 1606, 1510, 1460, 1393, 1375, 1263, 1240, 1174, 1079, 1053, 927, 851, 826, 768 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 9.20 (br m, 1H, NH), 8.94 (d, J = 3.8 Hz, 1H, H-2′), 8.35 (d, J = 7.9 Hz, 1H, H-5′), 8.07 (d, J = 8.2 Hz, 1H, H-8′), 7.80 (br m, 1H, H-7′), 7.72 (o m, 1H, H-6′), 7.71 (o m, 1H, H-3′), 6.08 (br m, 1H, H-9), 5.80 (br m, 1H, H-10), 5.13 (d, J = 17.1 Hz, 1H, H-11a), 5.08 (d, J = 10.5 Hz, 1H, H-11b), 4.25 (s, 3H, H-5′′), 3.28 (br m, 1H, H-8), 2.99 (o m, 1H, H-2a), 2.90 (o m, 1H, H-6a), 2.84 (o m, 1H, H-6b), 2.80 (o m, 1H, H-2b), 2.21 (br m, 1H, H-3), 1.51 (o m, 1H, H-4), 1.49 (o m, 2H, H-5ab), 0.91 (o m, 1H, H-7a), 0.84 (o m, 1H, H-7b).
13C NMR (125.76 MHz, DMSO-d6): δ = 189.8 (C-1′′ or C-2′′), 181.6 (C-1′′ or C-2′′), 177.5 (C-3′′), 171.2 (C-4′′), 150.4 (C-2′), 148.0 (C-8a′), 144.7 (C-4′), 140.6 (C-10), 130.1 (C-8′), 129.6 (C-7′), 127.3 (C-6′), 126.3 (C-4a′), 122.9 (C-5′), 119.8 (C-3′), 114.6 (C-11), 60.1 (C-5′′), 59.2 (C-8), 52.2 (C-9), 48.8 (C-6), 45.9 (C-2), 38.6 (C-3), 27.3 (C-4), 26.0 (C-5), 24.7 (C-7).
HRMS (ESI+): m/z [M + H]+ calcd for C24H25N3O3: 404.1974; found: 404.1974.
To the best of our knowledge, the synthesis of C-HSQ has not been reported so far.

3.2.2. General Procedure for the Preparation of Cinchona Alkaloid-Based Catalysts

To a solution of cinchona half squaramide (Q-HSQ or HQ-HSQ or C-HSQ, 0.482 mmol) in methanol (2.1 mL), a solution of aza-crown ether (1-aza-15-crown-5 ether or 1-aza-18-crown-6 ether, 0.439 mmol) in methanol (2.1 mL) was added under argon atmosphere. The resulting mixture was warmed up to 60 °C, then it was stirred for 6 h at this temperature. The solvent was removed under reduced pressure, then the crude product was purified by column chromatography (SiO2; DCM:methanol = 10:1) to obtain Q5 or Q6 or HQ5 or C5 or C6.

3.2.3. 3-(+)-(1,4,7,10-Tetraoxa-13-azacyclopentadecan-13-yl)-4-(((S)-(6-methoxyquinolin-4-yl)((1S,2S,4S,5R)-5-vinylquinuclidin-2-yl)methyl)amino)cyclobut-3-ene-1,2-dione (Q5)

Yellowish white crystals (93 mg, 34% yield). TLC (SiO2 TLC; DCM:methanol = 10:1, Rf = 0.46). M.p. 97–100 °C. [ α ] D 25 : +53.1 (c = 1.0, chloroform).
IR (KBr): 3323, 2934, 2863, 1786, 1664, 1621, 1516, 1471, 1432, 1354, 1313, 1298, 1245, 1230, 1176, 1124, 1082, 1057, 1028, 851 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 8.77 (d, J = 4.3 Hz, 1H, H-2′), 7.95 (d, J = 9.2 Hz, 1H, H-8′), 7.89 (br m, 1H, H-5′), 7.69 (d, J = 4.3 Hz, 1H, H-3′), 7.64 (br m, 1H,NH), 6.26 (br m, 1H, H-9), 7.42 (dd, J = 9.2, and 2.3 Hz, 1H, H-7′), 5.93 (ddd, J = 17.2, 10.3, and 7.6 Hz, 1H, H-10), 5.04 (d, J = 17.2 Hz, 1H, H-11a), 4.99 (d, J = 10.3 Hz, 1H, H-11b), 3.98 (br m, 1H, H-2′′′a or H-11′′′a), 3.93 (s, 3H, H-9′), 3.65 (br m, 2H, H-2′′′ and/or H-11′′′), 3.00–3.60 (o m, 16H, H-3′′′, H-4′′′, H-5′′′, H-6′′′, H-7′′′, H-8′′′, H-9′′′, and H-10′′′), 3.47 (o m, 1H, H-2′′′b or H-11′′′b), 3.50 (o m, 1H, H-8), 3.33 (o m, 1H, H-6a), 3.17 (o m, 1H, H-2a), 2.72 (dm, J = 13.3 Hz, 1H, H-2b), 2.62 (br m, 1H, H-6b), 2.27 (br m, 1H, H-3), 1.57 (br m, 1H, H-4), 1.49 (o m, 2H, H-5ab), 1.42 (tm, J = 12.2 Hz, 1H, H-7a), 0.55 (dm, J = 13.2, and 7.1 Hz, 1H, H-7b).
13C NMR (125.76 MHz, DMSO-d6): δ = 182.2 (C-1′′ or C-2′′), 181.8 (C-1′′ or C-2′′), 168.3 (C-3′′ or C-4′′), 166.5 (C-3′′ or C-4′′), 157.8 (C-6′), 144.2 (C-4′ or C-8a′), 144.0 (C-4′ or C-8a′), 147.8 (C-2′), 142.2 (C-10), 131.4 (C-8′), 127.7 (C-4a′), 121.9 (C-7′), 119.9 (C-3′), 114.4 (C-11), 101.8 (C-5′), 68.5–70.5 (C-3′′′, C-4′′′, C-5′′′, C-6′′′, C-7′′′, C-8′′′, C-9′′′, and C-10′′′), 58.6 (C-8), 55.7 (C-9′), 55.5 (C-2), 52.8 (C-9), 51.8 (C-2′′′ or C-13′′′), 51.4 (C-2′′′ or C-13′′′), 40.1 (C-6), 39.2 (C-3), 27.4 (C-5), 27.3 (C-4), 26.1 (C-7).
HRMS (ESI+): m/z [M + H]+ calcd for C34H44N4O7: 621.3288; found: 621.3285.
To the best of our knowledge, the synthesis of Q5 has not been reported so far.

3.2.4. 3-(+)-(1,4,7,10,13-Pentaoxa-16-azacyclooctadecan-16-yl)-4-(((S)-(6-methoxyquinolin-4-yl)((1S,2S,4S,5R)-5-vinylquinuclidin-2-yl)methyl)amino)cyclobut-3-ene-1,2-dione (Q6)

Pale-brown oil (64 mg, 22% yield). TLC (SiO2 TLC; DCM:methanol = 10:1, Rf = 0.19). [ α ] D 25 : +64.5 (c = 1.0, chloroform).
IR (KBr): 3311, 2909, 2867, 1784, 1666, 1622, 1574, 1521, 1473, 1431, 1347, 1315, 1303, 1248, 1231, 1134, 1107, 1028, 980, 833 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 8.78 (d, J = 4.3 Hz, 1H, H-2′), 7.96 (d, J = 9.1 Hz, 1H, H-8′), 7.90 (br m, 1H, H-5′), 7.71 (br m, 1H, H-3′), 7.66 (br m, 1H, NH), 6.32 (br m, 1H, H-9), 7.42 (dd, J = 9.1, and 2.2 Hz, 1H, H-7′), 5.95 (ddd, J = 17.3, 10.5, and 6.7 Hz, 1H, H-10), 5.08 (d, J = 17.3 Hz, 1H, H-11a), 5.03 (d, J = 10.5 Hz, 1H, H-11b), 3.99 (br m, 1H, H-2′′′a or H-11′′′a), 3.98 (br m, 1H, H-8), 3.93 (s, 3H, H-9′), 3.71 (br m, 2H, H-2′′′ and/or H-11′′′), 3.00–3.64 (o m, 18H, H-4′′′, H-5′′′, H-6′′′, H-7′′′, H-8′′′, H-9′′′, H-10′′′, H-11′′′, and H-12′′′), 3.52 (o m, 1H, H-2′′′b or H-11′′′b), 3.38 (o m, 1H, H-6a), 3.27 (o m, 1H, H-2a), 2.78 (br m, 1H, H-2b), 2.69 (br m, 1H, H-6b), 2.34 (br m, 1H, H-3), 1.62 (o m, 1H, H-4), 1.50 (o m, 1H, H-7a), 0.60 (br m, 1H, H-7b), and (2H, H-5).
13C NMR (125.76 MHz, DMSO-d6): δ = 182.1 (C-1′′ or C-2′′), 181.9 (C-1′′ or C-2′′), 168.2 (C-3′′ or C-4′′), 166.2 (C-3′′ or C-4′′), 157.9 (C-6′), 144.2 (C-8a′), 147.7 (C-2′), 142.0 (C-10), 131.4 (C-8′), 127.7 (C-4a′), 121.9 (C-7′), 120.0 (C-3′), 114.7 (C-11), 101.7 (C-5′), 68.5–70.5 (C-3′′′, C-4′′′, C-5′′′, C-6′′′, C-7′′′, C-8′′′, C-9′′′, C-10′′′, C-11′′′, and C-12′′′), 55.7 (C-8, C-9′), 55.3 (C-2), 52.6 (C-9), 50.6 (C-2′′′ or C-13′′′), 49.7 (C-2′′′ or C-13′′′), 40.2 (C-6), 38.9 (C-3), 27.7 (C-4), 25.8 (C-7), and (C-5, C-4′).
15N NMR (50.68 MHz, DMSO-d6): 312.3 (N-1′′).
HRMS (ESI+): m/z [M + H]+ calcd for C36H48N4O8: 665.3550; found: 665.3550.
To the best of our knowledge, the synthesis of Q6 has not been reported so far.

3.2.5. 3-(+)-(1,4,7,10-Tetraoxa-13-azacyclopentadecan-13-yl)-4-(((S)-((1S,2S,4S,5R)-5-ethylquinuclidin-2-yl)(6-methoxyquinolin-4-yl)methyl)amino)cyclobut-3-ene-1,2-dione (HQ5)

White powder (208 mg, 76% yield). TLC (SiO2 TLC; DCM:methanol = 10:1, Rf = 0.19). M.p. 107 °C. [ α ] D 25 : +50.8 (c = 1.0, chloroform).
IR (KBr): 3271, 2930, 2862, 1789, 1672, 1622, 1579, 1519, 1475, 1432, 1353, 1297, 1257, 1241, 1230, 1122, 1029, 939, 854, 831 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 8.78 (d, J = 4.3 Hz, 1H, H-2′), 7.95 (d, J = 9.2 Hz, 1H, H-8′), 7.89 (br, 1H, H-5′), 7.73 (br m, 1H, H-3′), 7.67 (br m, 1H,NH), 6.28 (br m, 1H, H-9), 7.42 (dd, J = 9.2, and 2.3 Hz, 1H, H-7′), 4.02 (br m, 1H, H-2′′′a or H-11′′′a), 3.93 (s, 3H, H-9′), 3.65 (br m, 2H, H-2′′′ and/or H-11′′′), 3.10–3.62 (o m, 14H, H-4′′′, H-5′′′, H-6′′′, H-7′′′, H-8′′′, H-9′′′, and H-10′′′), 3.57 (o m, 1H, H-8), 3.45 (o m, 1H, H-2′′′b or H-11′′′b), 3.36 (o m, 1H, H-6a), 3.20 (o m, 1H, H-2a), 2.67 (br m, 1H, H-6b), 2.51 (o m, 1H, H-2b), 1.58 (o m, 1H, H-4), 1.56 (o m, 1H, H-5a), 1.45 (o m, 1H, H-5b), 1.44 (o m, 1H, H-3), 1.41 (o m, 1H, H-7a), 1.34 (o m, 2H, H-10ab), 0.81 (t, J = 7.0 Hz, 3H, H-11), 0.58 (br m, 1H, H-7b).
13C NMR (125.76 MHz, DMSO-d6): δ = 182.2 (C-1′′ or C-2′′), 181.8 (C-1′′ or C-2′′), 168.4 (C-3′′ or C-4′′), 166.4 (C-3′′ or C-4′′), 157.8 (C-6′), 144.2 (C-8a′), 147.7 (C-2′), 131.4 (C-8′), 127.6 (C-4a′), 121.9 (C-7′), 119.9 (C-3′), 101.7 (C-5′), 69.0–70.0 (C-3′′′, C-4′′′, C-5′′′, C-6′′′, C-7′′′, C-8′′′, C-9′′′, and C-10′′′), 58.5 (C-8), 56.7 (C-2), 55.7 (C-9′), 52.4 (C-9), 51.8 (C-2′′′ or C-11′′′), 51.4 (C-2′′′ or C-11′′′), 39.8 (C-6), 36.5 (C-3), 27.6 (C-5), 27.0 (C-10), 25.5 (C-7), 24.8 (C-4), 11.9 (C-11), and (C-4′).
15N NMR (50.68 MHz, DMSO-d6): 312.1 (N-1′′).
HRMS (ESI+): m/z [M + H]+ calcd for C34H46N4O7: 623.3445; found: 623.3450.
To the best of our knowledge, the synthesis of HQ5 has not been reported so far.

3.2.6. 3-(+)-(1,4,7,10-Tetraoxa-13-azacyclopentadecan-13-yl)-4-(((R)-quinolin-4-yl((1S,2R,4S,5R)-5-vinylquinuclidin-2-yl)methyl)amino)cyclobut-3-ene-1,2-dione (C5)

White powder (122 mg, 47% yield). TLC (SiO2 TLC; DCM:methanol = 10:1, Rf = 0.46). M.p. 204 °C. [ α ] D 25 : +45.2 (c = 1.0, chloroform).
IR (KBr): 3308, 2934, 2895, 2856, 1790, 1664, 1575, 1526, 1475, 1433, 1384, 1352, 1317, 1265, 1249, 1124, 1080, 1058, 770 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 8.94 (d, J = 4.2 Hz, 1H, H-2′), 8.51 (d, J = 8.5 Hz, 1H, H-5′), 8.05 (d, J = 8.3 Hz, 1H, H-8′), 7.79 (o m, 1H, H-3′), 7.78 (o m, 1H, H-7′), 7.67 (o m, 1H, NH), 7.70 (o m, 1H, H-6′), 6.36 (br m, 1H, H-9), 5.81 (ddd, J = 17.3, 10.5, and 6.7 Hz, 1H, H-10), 5.17 (d, J = 17.3 Hz, 1H, H-11a), 5.08 (d, J = 10.5 Hz, 1H, H-11b), 3.93 (br m, 1H, H-2′′′a or H-11′′′a), 3.67 (br m, 2H, H-2′′′ and/or H-11′′′), 3.26–3.62 (o m, 14H, H-4′′′ H-5′′′ H-6′′′ H-7′′′ H-8′′′ H-9′′′ and H-10′′′), 3.58 (o m, 4H, H-3′′′ and H-10′′′), 3.52 (br m, 1H, H-2′′′b or H-11′′′b), 3.40 (o m, 1H, H-8), 3.14 (br m, 1H, H-2a), 2.96 (br m, 1H, H-6a), 2.88 (o m, 1H, H-6b), 2.82 (br m, 1H, H-2b), 2.23 (br m, 1H, H-3), 1.52 (o m, 1H, H-4), 1.51 (o m, 2H, H-5ab), 0.95 (br m, 1H, H-7a), 0.88 (br m, 1H, H-7b).
13C NMR (125.76 MHz, DMSO-d6): δ = 182.2 (C-1′′ or C-2′′), 182.1 (C-1′′ or C-2′′), 168.2 (C-3′′ or C-4′′), 166.8 (C-3′′ or C-4′′), 150.3 (C-2′), 148.0 (C-8a′), 145.7 (C-4′), 140.5 (C-10), 129.9 (C-8′), 129.4 (C-7′), 126.9 (C-6′), 126.6 (C-4a′), 123.5 (C-5′), 119.7 (C-3′), 114.6 (C-11), 69.0–70.0 (C-3′′′, C-4′′′, C-5′′′, C-6′′′, C-7′′′, C-8′′′, C-9′′′, and C-10′′′), 59.0 (C-8), 51.8 (C-9 and C-2′′′ or C-11′′′), 51.4 (C-2′′′ or C-11′′′), 48.8 (C-6), 45.8 (C-2), 38.6 (C-3), 27.3 (C-4), 25.9 (C-5), 24.9 (C-7).
15N NMR (50.68 MHz, DMSO-d6): 312.4 (N-1′′), 105.8 (9-NH, 1J = 95 Hz).
HRMS (ESI+): m/z [M + H]+ calcd for C33H42N4O6: 591.3183; found: 591.3179.
To the best of our knowledge, the synthesis of C5 has not been reported so far.

3.2.7. 3-(+)-(1,4,7,10,13-Pentaoxa-16-azacyclooctadecan-16-yl)-4-(((R)-quinolin-4-yl((1S,2R,4S,5R)-5-vinylquinuclidin-2-yl)methyl)amino)cyclobut-3-ene-1,2-dione (C6)

Pale-brown oil (137 mg, 49% yield). TLC (SiO2 TLC; DCM:methanol = 5:1, Rf = 0.62). [ α ] D 25 : +25.2 (c = 1.0, chloroform).
IR (KBr): 3275, 2935, 2867, 1789, 1671, 1637, 1577, 1520, 1477, 1431, 1350, 1297, 1259, 1204, 1116, 987, 916, 852, 828, 771 cm–1.
1H NMR (500.13 MHz, DMSO-d6): δ = 8.94 (d, J = 4.0 Hz, 1H, H-2′), 8.52 (d, J = 8.3 Hz, 1H, H-5′), 8.06 (d, J = 8.3 Hz, 1H, H-8′), 7.80 (br m, 1H, H-3′), 7.78 (t, J = 7.7 Hz, 1H, H-7′), 7.68 (t, J = 7.4 Hz, 1H, H-6′), 7.65 (o m, 1H,NH), 6.41 (br m, 1H, H-9), 5.80 (ddd, J = 17.3, 10.4, and 6.3 Hz, 1H, H-10), 5.18 (d, J = 17.3 Hz, 1H, H-11a), 5.08 (d, J = 10.4 Hz, 1H, H-11b), 3.96 (br m, 1H, H-2′′′a or H-11′′′a), 3.71 (br m, 2H, H-2′′′ and/or H-11′′′), 3.25–3.64 (o m, 18H, H-4′′′, H-5′′′, H-6′′′, H-7′′′, H-8′′′, H-9′′′, H-10′′′, H-11′′′, and H-12′′′), 3.57 (o m, 1H, H-2′′′b or H-11′′′b), 3.47 (br m, 1H, H-8), 3.17 (o m, 1H, H-2a), 2.98 (o m, 1H, H-6a), 2.91 (o m, 1H, H-6b), 2.83 (o m, 1H, H-2b), 2.26 (br m, 1H, H-3), 1.56 (o m, 3H, H-4, H-5ab), 1.01 (br m, 1H, H-7a), 0.89 (br m, 1H, H-7b).
13C NMR (125.76 MHz, DMSO-d6): δ = 182.13 (C-1′′ or C-2′′), 182.11 (C-1′′ or C-2′′), 168.0 (C-3′′ or C-4′′), 166.6 (C-3′′ or C-4′′), 150.4 (C-2′), 145.5 (C-4′), 148.0 (C-8a′), 140.4 (C-10), 129.9 (C-8′), 129.5 (C-7′), 127.1 (C-6′), 126.6 (C-4a′), 123.6 (C-5′),119.9 (C-3′), 114.7 (C-11), 68.5–71.0 (C-3′′′, C-4′′′, C-5′′′, C-6′′′, C-7′′′, C-8′′′, C-9′′′, C-10′′′, C-11′′′, and C-12′′′), 58.9 (C-8), 52.0 (C-9), 51.6 (C-2′′′ or C-13′′′), 49.7 (C-2′′′ or C-13′′′), 48.8 (C-6), 45.8 (C-2), 38.6 (C-3), 27.2 (C-4), 25.8 (C-5), 25.0 (C-7).
HRMS (ESI+): m/z [M + H]+ calcd for C35H46N4O7: 635.3445; found: 635.3459.
To the best of our knowledge, the synthesis of C6 has not been reported so far.

3.2.8. Procedure for the Preparation of the Glucose-Based Catalyst: 3-(+)-((3,5-bis(trifluoromethyl)phenyl)amino)-4-((2R,4aR,6S,6aR,19aS,19bR)-6-methoxy-2-phenyltetradecahydro-13H-[1,3]dioxino[4′,5′:5,6]pyrano[3,4-e][1,4,7,10]tetraoxa[13]azacyclopentadecin-13-yl)cyclobut-3-ene-1,2-dione (G5)

To a solution of HSQ (119 mg, 0.350 mmol) in methanol (2.5 mL), a solution of crown ether derivative G (140 mg, 0.319 mmol) in methanol (2.5 mL) was added under argon atmosphere. The resulting mixture was warmed up to 60 °C, then stirred for 6 h at this temperature. The solvent was removed under reduced pressure, then the crude product was purified by column chromatography (SiO2; DCM:methanol = 40:1) to obtain G5 as a yellowish-white powder (131 mg, 55% yield). TLC (SiO2 TLC; DCM:methanol = 40:1, Rf = 0.19). M.p. 209–211 °C. [ α ] D 25 : +32.5 (c = 1.0, chloroform).
IR (KBr): 3262, 3080, 2931, 2868, 1785, 1687, 1593, 1562, 1475, 1458, 1427, 1383, 1278, 1174, 1093, 1059, 1018, 992, 885, 766 cm–1.
1H NMR (500.13 MHz, DMSO-d6):
Major signals: δ = 9.55 (br s, 0.6H, NH), 7.92 (s, 2H, H-2′′′ and H-6′′′), 7.56 (br m, 0.55H, H-4′′′), 7.41 (o m, 2H, H-2′′ and H-6′′), 7.37 (o m, 2H, H-3′′ and H-5′′), 7.36 (o m, 1H, H-4′′), 5.61 (s, 0.65H, H-2), 4.92 (br m, 0.55H, H-6), 4.16 (o m, 2H, H-4ab), 3.30–3.90 (o m, 12H, H-8, H-9, H-10, H-15, H-17 and H-18), 3.55 (o m, 2H, H-4′′′ and H-5′′′), 3.52 (o m, 1H, H-19a), 3.35 (o m, 1H, H-6a), 3.31 (o s, 3H, CH3).
Minor signals: δ = 7.65 (br m, 0.45H, H-4′′′), 7.38 (o m, H-3′′ and H-5′′), 7.35 (o m, 1H, H-4′′), 7.32 (o m, H-3′′ and H-5′′), 7.27 (o m, H-2′′ and H-6′′), 5.45 (s, 0.5H, H-2), 3.24 (s, 1.7H, CH3), 3.19 (s, 0.5H, CH3)
13C NMR (125.76 MHz, DMSO-d6):
Major signals: δ = 185.7 (C-1′ or C-2′), 180.9 (C-1′ or C-2′), 171.1 (C-3′ or C-4′), 162.4 (C-3′ or C-4′), 141.0 (C-1′′′), 137.6 (C-1′′), 130.8 (q, J = 32.8 Hz, C-3′′′ and C-5′′′), 128.8 (C-4′′), 128.0 (C-3′′ and C-5′′), 126.1 (C-2′′ and C-6′′), 123.3 (q, J = 272.5 Hz, CF3), 119.1 (C-2′′′ and C-6′′′), 114.6 (C-4′′′), 100.5 (C-2), 97.3 (C-6), 81.1 (C-19b), 78.7 (C-6a), 77.8 (C-19a), 68.3–72.0 (C-8, C-9, C-11, C-15, C-17 and C-18), 68.1 (C-4), 62.0 (C-4a), 54.6 (CH3), 52.0 (C-2′′ or C-11′′), 51.7 (C-12 or C-14).
Minor signals: δ = 137.8 (C-1′′), 137.7 (C-1′′), 128.7 (C-4′′), 128.15 (C-3′′ and C-5′′), 128.12 (C-3′′ and C-5′′), 125.9 (C-2′′ and C-6′′), 119.1 (C-2′′′ and C-6′′′), 114.6 (C-4′′′), 100.5 (C-2), 98.4 (C-6), 97.8 (C-6), 80.8 (C-19b), 78.8 (C-6a), 62.2 (C-4a), 58.1 (CH3), 58.0 (CH3), 54.7 (CH3).
HRMS (ESI+): m/z [M + H]+ calcd for C34H36F6N2O10: 747.2352; found: 747.2355.
To the best of our knowledge, the synthesis of G5 has not been reported so far.

3.2.9. General Procedure for the α-alkylation of tert-butyl methyl α-benzylmalonate with allyl halides (3)

To a solution of tert-butyl methyl α-benzylmalonate (1, 26 mg, 0.1 mmol) in the indicated solvent, catalyst Q5 or Q6 or HQ5 or C5 or C6 or G5, and a base were added (exact reaction conditions are shown in Table 1, Table 2, Table 3 and Table 4). Then, the allyl halide (2, allyl bromide or allyl iodide) was added, and the resulting mixture was stirred vigorously at the temperature shown in Table 1, Table 2, Table 3 and Table 4. After 24 h, water (1 mL) was added to the reaction mixture, and it was extracted with DCM (2 × 1 mL). The combined organic phase was dried over MgSO4 and filtered (except in the cases of MeCN and THF solvents, when the mixture was only dried over MgSO4, then filtered). The volatile components were removed under reduced pressure. The residue was purified by preparative TLC (SiO2; hexan:EtOAc = 4:1) to obtain product 3 as a colorless oil. Yields and enantiomeric excess (ee) values can be seen in Table 1, Table 2, Table 3 and Table 4. Spectroscopic data are fully consistent with those reported in the literature [3]. TLC (SiO2 TLC; hexane:EtOAc = 10:1, Rf = 0.47). [ α ] D 25 : –1.7 (c = 1.0, chloroform, 15% ee, (R) abs. config.) (lit. α D 25 : –7.6 (c = 1.0, chloroform, 94% ee, (R) abs. config.). The enantiomeric excess values were determined by chiral HPLC using a Phenomenex Lux® 5-µm Cellulose-2 column (250 × 4.6 mm ID), and a 95:5 mixture of hexane/ethanol was used as the eluent with a flow rate of 0.8 mL min−1. The column temperature was 20 °C. UV detector λ = 254 nm. Retention time for (S)-3: 5.0 min, for (R)-3: 5.4 min.
MS (ESI+): m/z (%) = 249 (100) [M-t-Bu + 2H]+.

3.2.10. General Procedure for the Michael Addition of tert-butyl methyl α-benzylmalonate to benzyl acrylate (5)

To a solution of tert-butyl methyl α-benzylmalonate (1, 26 mg, 0.1 mmol) in DCM (4 mL), catalyst C5 (6 mg, 0.01 mmol) and aq. 50% NaOH (35 µL, 0.65 mmol) were added. Then, benzyl acrylate (4, 23 µL 24 mg, 0.15 mmol) was added, and the resulting mixture was stirred vigorously at 0 °C. After 24 h, water (1 mL) was added to the reaction mixture, then it was extracted with DCM (2 × 1 mL), dried over MgSO4, and filtered. The volatile components were removed under reduced pressure. The residue was purified by preparative TLC (SiO2; hexane:EtOAc = 10:1) to obtain product 5 (32 mg, 74% yield, 2% ee) as a colorless oil. Spectroscopic data are fully consistent with those reported in the literature [2]. TLC (SiO2 TLC; hexane:EtOAc = 10:1, Rf = 0.23). In this case, the enantiomeric excess values were determined by chiral HPLC using a Kromasil® 5-µm AmyCoat® column (250 × 4.6 mm ID), an 85:15 mixture of hexane/ethanol was used as the eluent with a flow rate of 0.8 mL min−1. The column temperature was 20 °C. UV detector λ = 254 nm. Retention times: 6.8 min and 8.7 min.
MS (ESI+): m/z (%) = 371 (100) [M-t-Bu + 2H]+.

4. Conclusions

In conclusion, we have described the synthesis of six new crown ether-squaramide phase-transfer organocatalysts bearing four different chiral units. To the best of our knowledge, this is the first successful direct coupling of squaramide and aza-crown-ether units and the first application of crown ether-squaramides as phase-transfer catalysts. We have tested their performance in the asymmetric α-alkylation of tert-butyl methyl α-benzylmalonate with extensive parameter investigation, after which reaction of the products could be converted to α,α-disubstituted amino acids through Curtius rearrangement. The alkylation reactions afforded the products in excellent yields, but with only low enantiomeric excess values. Thus, despite the low enantioselectivity, the new crown ether-based catalysts can catalyze the often-difficult construction of quaternary carbon centers. With the catalyst having the best enantioselectivity, a Michael addition reaction of tert-butyl methyl α-benzylmalonate was also conducted; however, no enantioselectivity was observed. Based on our results, we anticipate that a linker between the crown ether and the squaramide unit may potentially increase the enantioselectivity of this new type of phase-transfer catalyst, as two NH groups may form stronger hydrogen bonds during catalysis.

Supplementary Materials

The following are available online: NMR spectra of the new compounds and HPLC chromatograms of products 3 and 5.

Author Contributions

Conceptualization, Z.F., S.N. and J.K.; methodology, Z.F., S.N. and J.K.; synthesis of compounds, Z.F., D.R. and Z.R.; performing NMR experiments, data analysis, A.S.; determination of ee values by chiral HPLC measurements, P.B. (Péter Bagi); performing HRMS experiments, data analysis, L.D.; writing—original draft preparation, Z.F.; writing—review and editing, J.K., P.H., P.B. (Péter Bakó), P.B. (Péter Bagi), Z.F. and Z.R.; project administration, J.K.; funding acquisition, P.H. and J.K.; resources, P.H. and J.K.; supervision, J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the New National Excellence Program of the Ministry of Human Capacities, grant numbers ÚNKP-20-3-II-BME-325 (S.N.), ÚNKP-21-2-I-BME-209 (D.R.) and ÚNKP-20-5-BME-322 (J.K.), and the János Bolyai Research Scholarship of the Hungarian Academy of Science (J.K.). It was also supported by the National Research, Development, and Innovation Office (grant numbers FK138037 and K128473), the Servier–Beregi PhD Research Fellowship (S.N.), and the Gedeon Richter′s Talentum Foundation (Z.F.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to Hajnalka Szabó-Szentjóbi from Budapest University of Technology and Economics for her assistance with the synthesis of crown ether derivatives, and Róbert Ritter from Budapest University of Technology and Economics for his assistance with the NMR measurements.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Hong, S.; Lee, J.; Kim, M.; Park, Y.; Park, C.; Kim, M.; Jew, S.; Park, H. Highly Enantioselective Synthesis of α,α-Dialkylmalonates by Phase-Transfer Catalytic Desymmetrization. J. Am. Chem. Soc. 2011, 133, 4924–4929. [Google Scholar] [CrossRef]
  2. Odanaka, Y.; Kanemitsu, T.; Iwasaki, K.; Mochizuki, Y.; Miyazaki, M.; Nagata, K.; Kato, M.; Itoh, T. Asymmetric Michael addition of malonic diesters to acrylates by phase-transfer catalysis toward the construction of quaternary stereogenic α-carbons. Tetrahedron 2019, 75, 209–219. [Google Scholar] [CrossRef]
  3. Kanemitsu, T.; Koga, S.; Nagano, D.; Miyazaki, M.; Nagata, K.; Itoh, T. Asymmetric Alkylation of Malonic Diester Under Phase-Transfer Conditions. ACS Catal. 2011, 1, 1331–1335. [Google Scholar] [CrossRef]
  4. Ha, M.W.; Lee, H.; Yi, H.Y.; Park, Y.; Kim, S.; Hong, S.; Lee, M.; Kim, M.; Kim, T.; Park, H. Enantioselective Phase-Transfer Catalytic α-Benzylation and α-Allylation of α-tert-Butoxycarbonyllactones. Adv. Synth. Catal. 2013, 355, 637–642. [Google Scholar] [CrossRef]
  5. Ha, M.W.; Hong, S.; Park, C.; Park, Y.; Lee, J.; Kim, M.; Lee, J.; Park, H. Enantioselective phase-transfer catalytic α-alkylation of 2-methylbenzyl tert-butyl malonates. Org. Biomol. Chem. 2013, 11, 4030–4039. [Google Scholar] [CrossRef] [PubMed]
  6. Park, C.; Ha, M.W.; Kim, B.; Hong, S.; Kim, D.; Park, Y.; Kim, M.; Lee, J.K.; Lee, J.; Park, H. Enantioselective α-Alkylation of Benzylideneamino tert-Butyl Malonates by Phase-Transfer Catalysis. Adv. Synth. Catal. 2015, 357, 2841–2848. [Google Scholar] [CrossRef]
  7. Kanemitsu, T.; Sato, M.; Yoshida, M.; Ozasa, E.; Miyazaki, M.; Odanaka, Y.; Nagata, K.; Itoh, T. Enantioselective α-Benzoyloxylation of Malonic Diesters by Phase-Transfer Catalysis. Org. Lett. 2016, 18, 5484–5487. [Google Scholar] [CrossRef] [PubMed]
  8. Tan, J.; Yasuda, N. Contemporary Asymmetric Phase Transfer Catalysis: Large-Scale Industrial Applications. Org. Process Res. Dev. 2015, 19, 1731–1746. [Google Scholar] [CrossRef]
  9. Hashimoto, T.; Maruoka, K. Recent Development and Application of Chiral Phase-Transfer Catalysts. Chem. Rev. 2007, 107, 5656–5682. [Google Scholar] [CrossRef] [PubMed]
  10. Shirakawa, S.; Maruoka, K. Recent Developments in Asymmetric Phase-Transfer Reactions. Angew. Chem. Int. Ed. 2013, 52, 4312–4348. [Google Scholar] [CrossRef]
  11. Majdecki, M.; Niedbała, P.; Tyszka-Gumkowska, A.; Jurczak, J. Assisted by Hydrogen-Bond Donors: Cinchona Quaternary Salts as Privileged Chiral Catalysts for Phase-Transfer Reactions. Synthesis 2021, 53, 2777–2786. [Google Scholar] [CrossRef]
  12. Brak, K.; Jacobsen, E.N. Asymmetric Ion-Pairing Catalysis. Angew. Chem. Int. Ed. 2013, 52, 534–561. [Google Scholar] [CrossRef] [Green Version]
  13. Schettini, R.;  Sicignano, M.; De Riccardis, F.;  Izzo, I.; Della Sala, G. Macrocyclic Hosts in Asymmetric Phase-Transfer Catalyzed Reactions. Synthesis 2018, 50, 4777–4795. [Google Scholar] [CrossRef]
  14. Wang, H. Chiral Phase-Transfer Catalysts with Hydrogen Bond: A Powerful Tool in the Asymmetric Synthesis. Catalysts 2019, 9, 244. [Google Scholar] [CrossRef] [Green Version]
  15. Majdecki, M.; Niedbala, P.; Jurczak, J. Amide-Based Cinchona Alkaloids as Phase-Transfer Catalysts: Synthesis and Potential Application. Org. Lett. 2019, 21, 8085–8090. [Google Scholar] [CrossRef] [PubMed]
  16. Majdecki, M.; Grodek, P.; Jurczak, J. Stereoselective α-Chlorination of β-Keto Esters in the Presence of Hybrid Amide-Based Cinchona Alkaloids as Catalysts. J. Org. Chem. 2021, 86, 995–1001. [Google Scholar] [CrossRef]
  17. Wang, B.; Liu, Y.; Sun, C.; Wei, Z.; Cao, J.; Liang, D.; Lin, Y.; Duan, H. Asymmetric Phase-Transfer Catalysts Bearing Multiple Hydrogen-Bonding Donors: Highly Efficient Catalysts for Enantio-and Diastereoselective Nitro-Mannich Reaction of Amidosulfones. Org. Lett. 2014, 16, 6432–6435. [Google Scholar] [CrossRef] [PubMed]
  18. Craig, R.; Litvajova, M.; Cronin, S.A.; Connon, S.J. Enantioselective acyl-transfer catalysis by fluoride ions. Chem. Commun. 2018, 54, 10108–10111. [Google Scholar] [CrossRef]
  19. Lu, N.; Fang, Y.; Gao, Y.; Wei, Z.; Cao, J.; Liang, D.; Lin, Y.; Duan, H. Bifunctional Thiourea–Ammonium Salt Catalysts Derived from Cinchona Alkaloids: Cooperative Phase-Transfer Catalysts in the Enantioselective Aza-Henry Reaction of Ketimines. J. Org. Chem. 2018, 83, 1486–1492. [Google Scholar] [CrossRef]
  20. Cui, D.; Li, Y.; Huang, P.; Tian, Z.; Jia, Y.; Wang, P. Bifunctional phase-transfer catalysts for synthesis of 2-oxazolidinones from isocyanates and epoxides. RSC Adv. 2020, 10, 12360–12364. [Google Scholar] [CrossRef]
  21. Alemán, J.; Parra, A.; Jiang, H.; Jørgensen, K.A. Squaramides: Bridging from Molecular Recognition to Bifunctional Organocatalysis. Chem. Eur. J. 2011, 17, 6890–6899. [Google Scholar] [CrossRef]
  22. Frontera, A.; Orell, M.; Garau, C.; Quiñonero, D.; Molins, E.; Mata, I.; Morey, J. Preparation, Solid-State Characterization, and Computational Study of a Crown Ether Attached to a Squaramide. Org. Lett. 2005, 7, 1437–1440. [Google Scholar] [CrossRef]
  23. Zdanowski, S.; Piątek, P.; Romański, J. An ion pair receptor facilitating the extraction of chloride salt from the aqueous to the organic phase. New J. Chem. 2016, 40, 7190–7196. [Google Scholar] [CrossRef]
  24. Jagleniec, D.; Siennicka, S.; Dobrzycki, Ł.; Karbarz, M.; Romański, J. Recognition and Extraction of Sodium Chloride by a Squaramide-Based Ion Pair Receptor. Inorg. Chem. 2018, 57, 12941–12952. [Google Scholar] [CrossRef] [PubMed]
  25. Jagleniec, D.; Dobrzycki, Ł.; Karbarz, M.; Romański, J. Ion-pair induced supramolecular assembly formation for selective extraction and sensing of potassium sulfate. Chem. Sci. 2019, 10, 9542–9547. [Google Scholar] [CrossRef]
  26. Yu, X.-H.; Cai, X.-J.; Hong, X.-Q.; Tam, K.Y.; Zhang, K.; Chen, W.-H. Synthesis and biological evaluation of aza-crown ether-squaramide conjugates as anion/cation symporters. Future Med. Chem. 2019, 11, 1091–1106. [Google Scholar] [CrossRef]
  27. Zaleskaya, M.; Jagleniec, D.; Karbarz, M.; Dobrzycki, Ł.; Romański, J. Squaramide based ion pair receptors possessing ferrocene as a signaling unit. Inorg. Chem. Front. 2020, 7, 972–983. [Google Scholar] [CrossRef]
  28. Yan, L.; Huang, G.; Wang, H.; Xiong, F.; Peng, H.; Chen, F. Squaramide-Linked Chloramphenicol Base Hybrid Catalysts for the Asymmetric Michael Addition of 2,3-Dihydrobenzofuran-2-carboxylates to Nitroolefins. Eur. J. Org. Chem. 2018, 1, 99–103. [Google Scholar] [CrossRef]
  29. Kisszekelyi, P.; Alammar, A.; Kupai, J.; Huszthy, P.; Barabas, J.; Holtzl, T.; Szente, L.; Bawn, C.; Adams, R.; Szekely, G. Asymmetric synthesis with cinchona-decorated cyclodextrin in a continuous-flow membrane reactor. J. Catal. 2019, 371, 255–261. [Google Scholar] [CrossRef]
  30. Rapi, Z.; Bakó, P.; Drahos, L.; Keglevich, G. Side-Arm Effect of a Methyl α-D-Glucopyranoside Based Lariat Ether Catalysts in Asymmetric Syntheses. Heteroatom Chem. 2015, 26, 63–71. [Google Scholar] [CrossRef]
  31. McCooey, S.H.; Connon, S.J. Readily Accessible 9-epi-amino Cinchona Alkaloid Derivatives Promote Efficient, Highly Enantioselective Additions of Aldehydes and Ketones to Nitroolefins. Org. Lett. 2007, 9, 599–602. [Google Scholar] [CrossRef]
  32. Nagy, S.; Dargó, G.; Kisszékelyi, P.; Fehér, Z.; Simon, A.; Barabás, J.; Höltzl, T.; Mátravölgyi, B.; Kárpáti, L.; Drahos, L.; et al. New enantiopure binaphthyl-cinchona thiosquaramides: Synthesis and application for enantioselective organocatalysis. New J. Chem. 2019, 43, 5948–5959. [Google Scholar] [CrossRef] [Green Version]
  33. Nagy, S.; Fehér, Z.; Dargó, G.; Barabás, J.; Garádi, Z.; Mátravölgyi, B.; Kisszékelyi, P.; Dargó, G.; Huszthy, P.; Höltzl, T.; et al. Comparison of Cinchona Catalysts Containing Ethyl or Vinyl or Ethynyl Group at Their Quinuclidine Ring. Materials 2019, 12, 3034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bakó, P.; Tőke, L.J. Novel monoaza-and diazacrown ethers incorporating sugar units and their extraction ability towards picrate salts. Inclusion Phenom. Mol. Recognit. Chem. 1995, 23, 195–201. [Google Scholar] [CrossRef]
  35. Bakó, P.; Kiss, T.; Tőke, L. Chiral azacrown ethers derived from D-glucose as catalysts for enantioselective Michael addition. Tetrahedron Lett. 1997, 38, 7259–7262. [Google Scholar] [CrossRef]
  36. Yang, W.; Du, D.-M. Highly Enantioselective Michael Addition of Nitroalkanes to Chalcones Using Chiral Squaramides as Hydrogen Bonding Organocatalysts. Org. Lett. 2010, 12, 5450–5453. [Google Scholar] [CrossRef] [PubMed]
  37. Makó, A.; Menyhárd, D.K.; Bakó, P.; Keglevich, G.; Tőke, L. Theoretical study of the asymmetric phase-transfer mediated epoxidation of chalcone catalyzed by chiral crown ethers derived from monosaccharides. J. Mol. Struct. 2008, 892, 336–342. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of cinchona alkaloid-based (Q5, Q6, HQ5, C5, C6) and d-glucose-based (G5) crown ether-squaramide phase-transfer catalysts. DIAD: diisopropyl azodicarboxylate, DPPA: diphenylphosphoryl azide, DCM: dichloromethane.
Scheme 1. Synthesis of cinchona alkaloid-based (Q5, Q6, HQ5, C5, C6) and d-glucose-based (G5) crown ether-squaramide phase-transfer catalysts. DIAD: diisopropyl azodicarboxylate, DPPA: diphenylphosphoryl azide, DCM: dichloromethane.
Molecules 26 06542 sch001
Scheme 2. Synthesis of α,α-disubstituted malonate (R)-3 by asymmetric α-alkylation of tert-butyl methyl α-benzylmalonate (1) with allyl halides 2 under phase-transfer conditions.
Scheme 2. Synthesis of α,α-disubstituted malonate (R)-3 by asymmetric α-alkylation of tert-butyl methyl α-benzylmalonate (1) with allyl halides 2 under phase-transfer conditions.
Molecules 26 06542 sch002
Scheme 3. Synthesis of α,α-disubstituted malonate 5 by Michael addition of tert-butyl methyl α-benzylmalonate (1) to benzyl acrylate (4) (Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), benzyl acrylate reagent (1.5 eq), C5 catalyst (10 mol%), 50% aq. NaOH base (6.5 eq), DCM solvent (4 mL), 0 °C, 24 h reaction time).
Scheme 3. Synthesis of α,α-disubstituted malonate 5 by Michael addition of tert-butyl methyl α-benzylmalonate (1) to benzyl acrylate (4) (Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), benzyl acrylate reagent (1.5 eq), C5 catalyst (10 mol%), 50% aq. NaOH base (6.5 eq), DCM solvent (4 mL), 0 °C, 24 h reaction time).
Molecules 26 06542 sch003
Scheme 4. Proposed transition state for the α-alkylation reaction of tert-butyl methyl α-benzylmalonate (1) with catalyst C5. R1, R2 = tert-butyl or methyl.
Scheme 4. Proposed transition state for the α-alkylation reaction of tert-butyl methyl α-benzylmalonate (1) with catalyst C5. R1, R2 = tert-butyl or methyl.
Molecules 26 06542 sch004
Table 1. Comparison of the catalysts in the α-alkylation of malonate 1 a.
Table 1. Comparison of the catalysts in the α-alkylation of malonate 1 a.
EntryCatalystBase bYield c (%)ee d (%)
1Q550% aq. NaOH985
2HQ550% aq. NaOH93<5
3C550% aq. NaOH739
4G550% aq. NaOH865
5Q650% aq. KOH59<5
6C650% aq. KOH37<5
a Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), allyl bromide reagent (1 eq), catalyst (10 mol%), base (50 eq), 2.4 mL dichloromethane (DCM) solvent, 25 °C, 24 h reaction time. b Applied base chosen considering the phase-transfer catalyst (PTC) crown ether cavity size. c Yield of isolated product purified by preparative thin-layer chromatography (TLC). d Determined by chiral high-performance liquid chromatography (HPLC).
Table 2. Examination of the effect of the solvent in the presence of catalyst C5 a.
Table 2. Examination of the effect of the solvent in the presence of catalyst C5 a.
EntrySolventBaseBase Quantity (eq)Yield b (%)ee c (%)
1DCM50% aq. NaOH50739
2toluene50% aq. NaOH50607
3EtOAc50% aq. NaOH50655
4MTBE50% aq. NaOH5051<5
5CPME50% aq. NaOH5049<5
62-Me-THF50% aq. NaOH5051<5
7DCMsolid NaOH13810
8MeCNsolid NaOH149<5
9THFsolid NaOH138<5
a Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), allyl bromide reagent (1 eq), C5 catalyst (10 mol%), base, 2.4 mL solvent, 25 °C, 24 h reaction time. b Yield of isolated product purified by preparative TLC. c Determined by chiral HPLC. ee: enantiomeric excess, DCM: dichloromethane, MTBE: tert-butyl methyl ether, CPME: cyclopentyl methyl ether, 2-Me-THF: 2-methyltetrahydrofuran, THF: tetrahydrofuran.
Table 3. Examination of the effect of catalyst quantity, reagent quantity and solvent volume in the presence of catalyst C5 a.
Table 3. Examination of the effect of catalyst quantity, reagent quantity and solvent volume in the presence of catalyst C5 a.
EntryCatalyst Quantity (mol%)Allyl Bromide Quantity (eq)Solvent Volume (mL)Yield b (%)ee c (%)
11012.4739
22012.4798
3101.22.48311
41011.2926
a Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), allyl bromide reagent, C5 catalyst, 50% aq. NaOH base (50 eq), DCM solvent, 25 °C, 24 h reaction time. b Yield of isolated product purified by preparative TLC. c Determined by chiral HPLC. ee: enantiomeric excess.
Table 4. Examination of the effect of temperature and quantities of reagent and base in the presence of catalyst C5 a.
Table 4. Examination of the effect of temperature and quantities of reagent and base in the presence of catalyst C5 a.
EntryTemperature (°C)Reagent Quantity (eq)Base Quantity (eq)Yield b (%)ee c (%)
1251509812
201509415
301257915
401.2509715
501.2257415
a Reaction conditions: tert-butyl methyl α-benzylmalonate (1 eq), allyl iodide reagent, C5 catalyst (10 mol%), 50% aq. NaOH base, 2.4 mL DCM solvent, 24 h reaction time. b Yield of isolated product purified by preparative TLC. c Determined by chiral HPLC. ee: enantiomeric excess.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fehér, Z.; Richter, D.; Nagy, S.; Bagi, P.; Rapi, Z.; Simon, A.; Drahos, L.; Huszthy, P.; Bakó, P.; Kupai, J. Synthesis of Novel Crown Ether-Squaramides and Their Application as Phase-Transfer Catalysts. Molecules 2021, 26, 6542. https://doi.org/10.3390/molecules26216542

AMA Style

Fehér Z, Richter D, Nagy S, Bagi P, Rapi Z, Simon A, Drahos L, Huszthy P, Bakó P, Kupai J. Synthesis of Novel Crown Ether-Squaramides and Their Application as Phase-Transfer Catalysts. Molecules. 2021; 26(21):6542. https://doi.org/10.3390/molecules26216542

Chicago/Turabian Style

Fehér, Zsuzsanna, Dóra Richter, Sándor Nagy, Péter Bagi, Zsolt Rapi, András Simon, László Drahos, Péter Huszthy, Péter Bakó, and József Kupai. 2021. "Synthesis of Novel Crown Ether-Squaramides and Their Application as Phase-Transfer Catalysts" Molecules 26, no. 21: 6542. https://doi.org/10.3390/molecules26216542

Article Metrics

Back to TopTop