Abstract

The enantioselective synthesis of a broad variety of novel differently functionalized α-halogenated α-aryl-β2,2-amino acid derivatives by means of an ammonium-salt-catalyzed asymmetric α-halogenation of isoxazolidin-5-ones was accomplished. Key to success to obtain high levels of enantioselectivities was the use of Maruoka’s spirocyclic binaphthyl-based ammonium salts, and detailed accompanying mechanistic studies using density functional theory methods revealed the key features for the catalyst–substrate interactions.
Keywords: Asymmetric Catalysis, Ammonium Salt Catalysis, Kinetic Resolution, DFT Calculations
Introduction
Investigations focusing on the asymmetric synthesis and further utilization of chiral non-natural amino acid derivatives have for decades been among the most prominent research topics in organic and bioorganic chemistry.1−5 A broad variety of conceptually different (catalytic) approaches to access non-natural amino acids (AA) with high levels of stereocontrol have been established, and the development of new synthesis strategies is still a highly contemporary field of research.1,6−12 In addition to the more classical focus on synthesis and applications of α-amino acids (α-AA) and α-AA-based peptides,1−7 non-natural β-AA have emerged as targets of significant interest over the past decades.8−17 The introduction of β-AA into the peptides, as well as the preparation of chiral β-AA-based heterocycles, can lead to peptidomimetics displaying unique (improved) biological properties,8−17 which makes the development of novel asymmetric approaches toward (masked) β-AA derivatives an important task.
Depending on their substitution pattern, different classes of β-AA can be defined (Scheme 1A). While several highly efficient strategies for the catalytic enantioselective synthesis of β3-, β2,3-, and β2-AA have been reported,8−12 the asymmetric construction of β2,2-AA remains challenging. In 2013, the Brière group reported the direct synthesis of isoxazolidin-5-ones 1 starting from Meldrum’s acid derivatives (Scheme 1B).18 Compounds 1 are versatile masked β2-AA derivatives which can be reacted in an asymmetric manner with different electrophiles to access the β2,2-AA derivatives 2 straightforwardly.19−29 These chiral heterocycles subsequently allow for the synthesis of free β2,2-AA and small peptides20−29 as well as for the synthesis of heterocyclic amino acids,30,31 to mention three potential applications only.
Scheme 1. β-AA, Recently Established Strategy for β2,2-AA, α-Halogenated β-AA, and the Herein Investigated α-Halogenation of β2,2-AA Derivatives.

Over the past few years, this powerful concept has successfully been used for a handful of asymmetric C–C bond forming reactions (conjugate additions to classical Michael acceptors, MBH-carbonates, and quinone methides;23−25 Mannich-type reactions;26,27 Pd-catalyzed allylations28,29) as well as asymmetric α-sulfanylations,20 α-trifluoromethylthiolations,21,22 and one α-amination example.23 Apart from these few recent reports, however, the suitability of compounds 1 to access a broader variety of α-(hetero)-functionalized β2,2-AA has so far not systematically been explored.
The asymmetric α-heterofunctionalization of amino acids32 has been a very versatile strategy to access novel AA derivatives with promising biological properties or may serve as useful building blocks for further manipulations. Interestingly, while asymmetric approaches toward α-halogenated β3-, β2,3-, and β2-AA have been well-described,32 stereoselective syntheses of α-halogenated β2,2-AA have so far very sparingly been reported33−37 (Scheme 1C). Considering the unique potential of compounds 1 to serve as precursors for novel masked β2,2-AA derivatives 2, we now became interested in exploring the suitability of compounds 1 for asymmetric α-halogenation reactions with different electrophilic halogen-transfer agents. This should give access to a new family of so far unprecedented α-halogenated α-arylated-β2,2-AA in a unique and direct manner by utilizing the easily available starting materials 1. Based on our own previous experience with compounds 1,22,24,25 as well as taking inspiration from Brière’s early reports,20,23 we focused on the use of chiral ammonium salt ion pairing catalysts38−42 to control compounds 1 in the herein targeted asymmetric α-halogenation approaches (chiral ammonium salt catalysts were also successfully used by Della Sala and Alemán for α-trifluoromethylthiolations of compounds 1(21)). In addition, we also thought about gathering a deeper understanding of these reactions by carrying out detailed density functional theory (DFT) studies within the context of this project.
Results and Discussion
Asymmetric α-Chlorination
The synthesis of chiral α-Cl-β2,2-AA has been very sparingly reported so far,37,43,44 and to the best of our knowledge, a reliable asymmetric catalysis approach to access (masked) α-Cl-β2,2-AA derivatives is yet missing. Considering the general value of enantioenriched α-Cl-carbonyl compounds to serve as building blocks for further manipulations (i.e., stereospecific SN2-type reactions),45−47 we now became interested in developing a protocol for the asymmetric electrophilic α-chlorination48,49 of isoxazolidin-5-ones 1 using the established chiral ammonium salt ion pairing catalysts A–C (Figure 1).
Figure 1.
Chiral ammonium salt ion pairing catalysts tested for the asymmetric α-halogenations of compounds 1.
As summarized in Table 1, a variety of different conditions and catalysts were tested for the α-chlorination of the α-phenyl-substituted parent substrate 1a using N-chlorosuccinimide (NCS, 3) as a readily available and established electrophilic Cl-transfer agent.48,49 Based on the recently observed privileged application potential of Maruoka’s spirocyclic ammonium salt catalysts A1 and A2(50) for asymmetric transformations of isoxazolidin-5-ones 1,20−25 we started our screening using 5 mol % of the ammonium salt A1 (R,R-configuration as depicted in Figure 1) in toluene in the presence of different mild bases (entries 1–3). Gratifyingly, in all cases, a complete conversion of 1a was observed, and the targeted product 2aCl could be obtained in reasonable isolated yields and with promising initial enantioselectivities up to 85:15 (favoring the (+)-isomer; please see the discussion below concerning the assignment of the depicted S-configuration). A further screening of different carbonate bases in different solvents did not allow for any improvement (results not given in the table), and in some cases, we also observed formation of the elimination product 4. Surprisingly, however (considering our previous observations with compounds 1 where weaker inorganic bases were beneficial20−25), it was possible to obtain 2aCl with a high er of 94:6 when using sodium phenoxide (PhONa) as a base instead (entry 4).51 Interestingly, despite the fact that we observed full conversion of 1a, product 2aCl could only be obtained in around 50% isolated yield, accompanied by formation of a, at this time not characterized, hardly soluble white precipitate. Initially, we suspected a problem with elimination and decomposition of product 2aCl in the presence of this base as well as homogenization difficulties of the base in toluene. We therefore tested the use of an ultrasonic bath, different temperatures and reaction times, and order of addition of reagents (conditions A vs conditions B) next (entries 4–8). The overall transformation turned out to be much faster when carried out in an ultrasonic bath with more or less identical yield and er (entry 5). To achieve a better mixing and homogenization without using an ultrasonic bath, we next tested the stepwise addition of reagents (conditions B, entry 6). In addition, we also reduced the amount of the valuable catalyst to 2 mol % for the further optimization. Interestingly, even with this lower amount of catalyst, a full conversion of starting material 1a was observed within 4 h under these conditions, and product 2aCl was again obtained in around 50% isolated yield with a reasonable er of 91:9. Surprisingly, the reaction as such was found to be relatively clean, with no formation of elimination product 4, and no other significant byproducts were observable in the crude product 1H NMR spectrum (recorded in CDCl3). Thus, we had a closer look on the hardly soluble precipitate that forms during this reaction and which was so far assumed to contain succinimide 3- or phenoxide-originating byproducts. This precipitate was separated during workup by means of a simple filtration and was found to be insoluble in CDCl3. In contrast, however, it was well-soluble in H2O and DMSO and, upon closer analysis, was identified as the (relatively instable) acid 5aCl. Mechanistically, this compound most likely gets formed by ring-opening addition of phenoxide to 2aCl followed by subsequent hydrolysis of the phenylester of 5aCl. It should be noted that we were not able to isolate this postulated phenylester, but when we tested the stability and reactivity of isolated 2aCl in the presence of catalyst A1 and different bases (including hydroxides), we realized that the nucleophilic PhONa was the only one which allowed for the formation of 5aCl, whereas other bases led to quantitative formation of the elimination product 4 only.
Table 1. Optimization of the Asymmetric α-Chlorination of Isoxazolidin-5-one 1aa.

| entry | cat. (mol %) | solvent | base (equiv) | conditionb | t (h) | conv. (%)c | yield (%)d | ere |
|---|---|---|---|---|---|---|---|---|
| 1 | A1 (5) | toluene | K2CO3 (1.2) | A | 18 | 100 | 64 | 85:15 |
| 2 | A1 (5) | toluene | K2HPO4 (1.2) | A | 18 | 100 | 82 | 80:20 |
| 3 | A1 (5) | toluene | NaOAc (1.2) | A | 18 | 100 | 90 | 81:19 |
| 4 | A1 (5) | toluene | PhONa (1.1) | A | 18 | 100 | 47 | 94:6 |
| 5f | A1 (5) | toluene | PhONa (1.1) | A | 1.5 | 100 | 53 | 93:7 |
| 6 | A1 (2) | toluene | PhONa (1.1) | B | 4 | 100 | 53 | 91:9 |
| 7 | A1 (2) | toluene | PhONa (1.1) | B | 6 | 100 | 32 | 97:3 |
| 8 | A1 (2) | toluene | PhONa (1.1) | B | 17 | 100 | 25 | 99.5:0.5 |
| 9 | A1 (2) | toluene | 4-NO2-C6H4ONa (1.1) | B | 5 | 100 | 71 | 84:16 |
| 10 | A1 (2) | THF | PhONa (1.1) | B | 4 | 100 | 34 | 81:19 |
| 11 | A1 (2) | Et2O | PhONa (1.1) | B | 4 | 100 | 36 | 85:15 |
| 12 | A1 (5) | toluene | PhONa (0.5) | B | 24 | 100 | 59 | 92:8 |
| 13 | A1 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 52 (33)h | 95:5 |
| 14 | A1 (2) | toluene | PhONa (0.5) | B | 72 | 100 | 46 | 91:9 |
| 15g | A1 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 52 | 95:5 |
| 16 | A2 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 52 | 85:15 |
| 17 | A3 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 54 | 62:38 |
| 18 | A4 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 46 | 73:27 |
| 19 | B1 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 54 | 55:45 |
| 20 | B2 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 54 | 55:45 |
| 21 | C1 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 51 | 53:47 |
| 22 | C2 (5) | toluene | PhONa (0.5) | B | 72 | 100 | 62 | 55:45 |
Unless otherwise stated, all reactions were carried out at 25 °C in the indicated solvent using 0.1 mmol 1a and 0.11 mmol 3 (0.05 M with respect to 1a).
Conditions A: 1a, 3, catalyst, and base were all placed in a flask followed by addition of the solvent. Conditions B: 1a, 3, and catalyst were dissolved in 50 vol % of the total solvent amount, followed by addition of a finely suspended mixture of the base (PhONa) in the remaining 50 vol % of the indicated solvent.
Conversion of 1a (determined by 1H NMR of the crude product mixture).
Isolated yield of 2aCl (after column chromatography).
Given as (+)/(−)-2aCl ratio (determined by HPLC using a chiral stationary phase).
Carried out in an ultrasonic bath.
Carried out at 0 °C.
Isolated yield of 5aCl.
Interestingly, Birman’s group recently reported a very appealing kinetic resolution (KR) of β-substituted isoxazolidin-5-ones with aliphatic alcohols in the presence of chiral squaramide catalysts, which gave access to various acyclic β3-AA esters in a mechanistically related manner.52 To probe if a conceptually similar KR may also account for our observations, we next treated racemic 2aCl with 0.5 equiv of PhONa in the presence of catalyst A1, which resulted in the formation of 5aCl and the recovery of enantioenriched (+)-2aCl (45% yield, er = 69:31, s = 3). In addition, when carrying out the overall α-chlorination protocol for prolonged reaction times with 1.1 equiv of PhONa (compare entries 6–8), the isolated yield of cyclic 2aCl constantly decreased, combined with a significantly increasing enantiopurity up to er = 99.5:0.5 (entry 8), thus substantiating the involvement of a chiral ammonium-salt-catalyzed resolution step. Additionally, the use of the less nucleophilic 4-NO2-C6H4-ONa resulted in a higher 2aCl yield but with lower selectivity (entry 9), which supports our proposal that the aryloxide serves as a nucleophile in the resolution step. Based on these results, it can therefore be postulated that the overall transformation most likely proceeds via two distinct steps, a relatively fast asymmetric ammonium-salt-catalyzed α-chlorination first, followed by a subsequent (slower) ammonium phenoxide-controlled kinetic resolution of the already enantioenriched 2a. These two asymmetric processes match each other, resulting in a reasonably selective two-step one-pot strategy to access enantioenriched 2aCl (overall, this reaction is best carried out in toluene while different ether solvents turned out to be not beneficial (entries 10 and 11)).
In our recent investigations concerning the asymmetric α-trifluoromethylthiolation of compounds 1 with succinimide or phthalimide-based SCF3-transfer reagents, we found that catalytic amounts of external bases may be sufficient, as the in situ formed succinimide or phthalimide can serve as a base, as well.22 Analogously, when we carried out the α-chlorination of 1a with 10–20 mol % of K2CO3 only, we obtained the same yield and enantioselectivity as observed for the use of 1.2 equiv of this base (compare with entry 1), demonstrating that the α-chlorination step itself is a fast autocatalytic process where the nature of the external base has a less pronounced effect only. We therefore speculated that it should be possible to use only 50 mol % of the phenoxide (in order to primarily control the KR step), which should allow for a synthetically useful compromise between isolated yield and enantiopurity. In addition, control of the reaction progress/reaction time should be less critical compared to the use of an excess of phenoxide. As shown in entries 12 and 13, the use of 50 mol % of PhONa in the presence of 5 mol % of A1 results in 59% isolated 2aCl yield after 24 h already (er = 92:8) and allows for a further enantioenrichment when stirring for a prolonged reaction time (52% isolated yield in combination with a satisfying er of 95:5 after 3 days, entry 13). As expected, lowering the catalyst amount to 2 mol % had a slightly detrimental effect on the overall selectivity (entry 14), whereas lower temperatures had no influence at all (entry 15), provided the reaction was run long enough to allow for a satisfying progress of the KR (substantiating that the resolution step is the slower process in this two-step protocol). Finally, other ammonium salt catalysts A–C were tested, but in close analogy to previous observations,20−25 only the spirocyclic Maruoka ammonium salts A1 and A2 allowed for reasonable selectivities (compare entries 16–22), while the other well-established systems failed to allow for any reasonable selectivities.
Having identified reliable conditions for this combined α-chlorination–kinetic resolution approach to access enantioenriched 2aCl, we next investigated a series of further manipulations of this masked α-chlorinated β-AA derivative (Scheme 2). First, it was possible to directly replace the chlorine with a NO2 (product 2aNO2), an EtO (product 2aOEt), and a N3 group (product 2aN3). Here, it should be noted that the α-NO2-containing 2aNO2 turned out to be a fairly unstable compound, which rapidly undergoes ring opening and decarboxylation to compound 8 upon exposure to silica gel. In contrast, 2aN3 and 2aOEt could be obtained with excellent levels of enantiospecificity when carrying out the nucleophilic SN2 displacement on enantioenriched 2aCl. In addition to these α-substitutions, it was also possible to carry out nucleophilic ring-opening reactions, as shown for the synthesis of the Me ester 6aCl or the amide 7aCl (analogous reactions could be carried out with 2aN3, as well53). Interestingly, however, while it was possible to hydrogenate the N–O bond of other α,α-disubstituted isoxazolidinones 2 with classical Pd-catalyzed approaches (either with H2 or HCOONH4) in the past,20−29 this was not possible for 2aCl, as illustrated for the formation of the dehalogenated ester 6aH under established heterogeneous Pd-catalyzed hydrogenation conditions (other methods were tested as well, but we were not able to reduce the N–O bond without cleaving the C–Cl bond).
Scheme 2. Further Manipulations of the Masked α-Chlorinated β2,2-AA Derivative 2aCl.

Conditions (a) NaN3 (1.1 equiv), acetone, 25 °C, 24 h [with (+)-2aCl (er = 99.0:0.5)]; (b) CHCl3 (0.6 vol % of EtOH), CsF (3 equiv), 18-crown-6, 25 °C, 1 h [with (+)-2aCl (er = 95:5)]; (c) NaNO2 (1.1 equiv), DMSO, 25 °C, 1 h [with rac-2aCl]; (d) Y(OTf)3 (10 mol %), MeOH, 25 °C, 72 h [with rac-2aCl]; (e) H2 (1 atm), Pd/C, MeOH, 25 °C, 24 h [with rac-2aCl]; (f) 4-ClC6H4CH2NH2 (5 equiv), MeOH, 25 °C, 24 h [with rac-2aCl].
When testing the α-chlorination of a variety of alternatively substituted α-arylated starting materials 1 under the optimized conditions next (Scheme 3), it turned out that this protocol, in general, tolerates different substitution patterns (like the halogenated derivatives 2eCl–2hCl), but some interesting limitations also became obvious. The thiophene-containing 2dCl was obtained with more than 50% yield but a lower enantioselectivity, indicating that the KR step is less efficient for this substrate as compared to others. In addition, the p-OMe- and p-OTBDMS-containing 2iCl and 2jCl could not be isolated as they decomposed very quickly, forming colored byproducts which most likely possess p-quinone methide-type structures. It was however possible to add NaN3 directly after completion of the α-chlorination, resulting in formation of the α-azidated products 2iN3 and 2jN3 instead. Unfortunately, enantioselectivities were not very high, which can be rationalized by a partial erosion of the enantiopurity of the primary reaction products 2iCl and 2jCl because of the aforementioned formation of quinone methide-type intermediates (to which NaN3 can add, as well54,55). It should be noted that we also tried to carry out this chlorination on α-alkyl-substituted derivatives 1 (e.g., Bn instead of Ar), but unfortunately, these turned out to be less reactive and gave trace amounts of the product only (the same outcome was obtained for the analogous fluorination reaction), which underscores the strong influence of the nature of the α-substituent on the reactivity of compounds 1.
Scheme 3. Asymmetric Application Scope for the Synthesis of Masked α-Chlorinated β2,2-AA Derivatives (+)-2Cl.

Asymmetric α-Fluorination
Based on the knowledge gathered for the asymmetric α-chlorination of isoxazolidin-5-ones 1, we next investigated the (analogous) α-fluorination of these compounds. Although a handful of previous reports described successful examples for the asymmetric synthesis of some α-F-β2,2-AA derivatives,33−36 the general enantioselective synthesis of these valuable targets is still far from being a solved challenge. Thus, we focused on the asymmetric ammonium-salt-catalyzed α-fluorination of the parent substrate 1a with N-fluorobenzenesulfonimide (NFSI, 9) as the electrophilic F-transfer agent next (Table 2 gives an overview about the most significant results obtained in a detailed screening of different catalysts and conditions). First attempts trying to apply our chlorination-inspired α-heterofunctionalization–kinetic resolution strategy with NaOPh failed, resulting in full decomposition of starting 1a, without any product 2aF formation (entry 1). We next changed for “more common” asymmetric ammonium salt conditions using Cs2CO3 as a solid inorganic base. This allowed for a promising first hit, giving (+)-2aF with reasonable conversion and a good er of 85:15 when using 5 mol % of the Maruoka catalyst A1 in toluene (entry 2). Noteworthy, at this point, we already observed a rather pronounced sensitivity of product 2aF to prolonged exposure to base or acid (including silica gel), leading to formation of the elimination product 4 as well as other unidentified decomposition products. This made purification of 2aF a bit tricky, requiring either a rather fast column chromatographic isolation or recrystallization from cyclohexane to obtain 2aF in reasonable purity and yield (although some loss of material was observed, as well, especially after silica gel column chromatography). For that reason, we calculated in situ yields using an internal NMR standard in all cases and carried out further isolation attempts only once suited overall conditions were identified.
Table 2. Optimization of the Asymmetric α-Fluorination of Isoxazolidin-5-one 1aa.

| entry | cat. (mol %) | solvent | base (equiv) | T (°C) | t (h) | yield (%)b | erc |
|---|---|---|---|---|---|---|---|
| 1 | A1 (5) | toluene | PhONa (1.1) | 25 | 24 | ||
| 2 | A1 (5) | toluene | Cs2CO3 (1.5) | 25 | 24 | 65 | 85:15 |
| 3 | A1 (5) | MTBE | Cs2CO3 (1.5) | 25 | 24 | 70 | 93:7 |
| 4 | A1 (5) | Et2O | Cs2CO3 (1.5) | 25 | 24 | 75 | 92:8 |
| 5 | A1 (5) | Et2O | K2CO3 (1.5) | 25 | 24 | <5 | |
| 6 | A1 (10) | Et2O | Cs2CO3 (1.5) | 25 | 65 | 85 (71d) | 92:8 |
| 7 | A2 (10) | Et2O | Cs2CO3 (1.5) | 25 | 65 | 60 | 91:9 |
| 8 | A1 (1) | Et2O | Cs2CO3 (1.5) | 25 | 65 | 45 | 85:15 |
| 9 | A1 (5) | Et2O | Cs2CO3 (1.5) | –20 | 65 | 75 | 94:6 |
| 10 | A1 (5) | Et2O | Cs2CO3 (1.5) | –40 | 65 | 60 | 95:5 |
| 11 | A1 (5) | Et2O | Cs2CO3 (1.5) | –60 | 144 | 15 | 96:4 |
| 12 | A1 (5) | MTBE | Cs2CO3 (1.5) | 25 | 40 | 85 (73d, 55e) | 93:7 |
Unless otherwise stated, all reactions were carried out using 0.1 mmol 1a and 0.25 mmol 9 in the indicated solvent (0.017 M with respect to 1a) under the given conditions.
In situ yields of 2aF determined using 4-fluoroanisol as an internal NMR standard.
Given as (+)/(−)-2aF ratio (determined by HPLC using a chiral stationary phase); please see the discussion below concerning the assignment of the S-configuration for the (+)-enantiomer.
Isolated yield after precipitation of excess of reagents, off-products, and catalyst with cyclohexane (containing less than 5 mol % of remaining diphenylsulfonimide, the given yield has been corrected for this “contamination”).
Isolated yield after column chromatography.
Conversion and er using Cs2CO3 could be improved by changing for ethereal solvents next (entries 3 and 4). While MTBE allowed for a marginally higher er, reactions in Et2O showed a slightly better conversion, and a further screening of conditions was carried out in Et2O then. Other bases were tested, as well, but, as exemplified for K2CO3 (entry 5), turned out to be not suitable, and we therefore relied on Cs2CO3 for the remaining optimization (variations of reagent and base ratios were also tested but without any improvement). To increase yield and er, we next used 10 mol % of A1 (65 h overall reaction time, entry 6). This allowed for a high er of 92:8 accompanied by a satisfying in situ yield of 85% and an isolated yield of 71% (after precipitation of reagents, off-products and catalyst with cyclohexane). Using other catalysts, the alternatively substituted Maruoka catalyst A2 gave almost the same selectivity (entry 7), while all the other scaffolds shown in Figure 1 again gave more or less racemic 2aF only (results not given in Table 2). Lowering the catalyst loading to 1 mol % (entry 8) led to a reduced yield and selectivity, and we thus again used 5 mol % of A1 for further attempts at lower temperatures (entries 9–11). Although it was possible to increase the er up to 96:4 at −60 °C, this increase in selectivity came with a significantly reduced conversion/yield. Therefore, to obtain a practical balance of yield and er, we finally opted for room temperature conditions and carried out the α-fluorination of 1a in MTBE for a slightly prolonged reaction time of 40 h (entry 12). This allowed for the synthesis of (+)-2aF in 85% in situ yield (isolated yields 73% after crystallization or 55% after column chromatography) and with an er of 93:7.
With these conditions in hand, we next investigated the application scope for the α-fluorination of starting materials 1 and the suitability of products 2F for further manipulations (Scheme 4). A variety of different aryl substituents were well-tolerated, resulting in reasonable enantioselectivities and in situ yields for products 2F. Unfortunately, the pronounced sensitivity of these compounds, however, made isolation by silica gel column chromatography difficult, especially for electron-rich aryl derivatives like 2iF (it should, however, be emphasized that we did not try to develop crystallization methods for each derivative as we did for the parent 2aF).
Scheme 4. Asymmetric Application Scope for the Synthesis of Masked α-Fluorinated β2,2-AA Derivatives (+)-2F and (Attempted) Further Transformations,

IST is the yield determined using 4-fluoroanisol as an internal NMR standard; SG is isolated yield after silica gel column chromatography.
Conditions: (a) Different hydrogenation conditions (with or without previous TFA-mediated Boc-deprotection of 2aF); (b) ArCH2NH2 in t-BuOH, 90 °C; (c) TFA in CH2Cl2 followed by addition of 12 in DMF.
Finally, we also investigated the use of the masked α-F-β-AA 2aF to carry out further transformations (Scheme 4, lower part). Hereby, we first investigated the reductive ring opening toward the free carboxylic acids 10 as well as the nucleophilic ring opening with benzylamine derivatives to access products 7. While the latter could be isolated in relatively low yields (accompanied by decomposition of 2aF under the basic reaction conditions), formation of the acids 10 could only be detected by direct LRMS analysis of the crude products (which contained significant amounts of unspecified side products already), but all attempts to isolate these products failed. Similar results were unfortunately obtained when testing the well-established KAHA ligation of 2aF with the ketoacid 12.56,57 Formation of the dipeptide 11aF could be confirmed by LRMS analysis, but again all attempts to isolate this interesting target failed because of its high sensitivity.
Asymmetric α-Bromination and Stereochemical Considerations
Having investigated the asymmetric synthesis of masked α-Cl and α-F-β2,2-AA derivatives 2Cl and 2F in much detail, we became interested in testing if analogous α-Br derivatives 2Br may be accessible, as well. Obviously, considering the observed sensitivity of compounds 2Cl and 2F under acidic and/or basic conditions, we expected an even more pronounced lability of the related Br target 2Br. Thus, we were also not too much surprised that we did not succeed in carrying out the direct electrophilic α-bromination of the parent substrate 1a with N-bromosuccinimide. Under several conditions that were tried, the starting material remained either unreacted or decomposed, and we therefore opted for an alternative approach to access 2aBr next. Recently the groups of Ibrahim and Adamo described the stereospecific SN2-type substitution of enantioenriched alkylphenylsulfides with Cl or Br,58−60 which provides an appealing entry to halogenated alkanes with good levels of stereocontrol (inversion of configuration). Inspired by these reports,58−60 and considering the fact that α-benzyl-substituted isoxazolidin-5-ones 1 were successfully α-sulfanylated under asymmetric ammonium salt catalysis by Brière before,20 we became interested if an asymmetric α-sulfanylation–desulfurylative bromination sequence may allow us to access the target α-Br derivative 2Br. In analogy to Brière’s pioneering report,20 the α-sulfanylation of the phenyl-substituted 1a could be carried out with good enantioselectivity with catalyst A2 (using succinimide 13 as the PhS-transfer agent; Scheme 5). Gratifyingly, utilizing the reported desulfurylation–bromination conditions,59 it was possible to access (−)-2aBr with good in situ yield and moderate levels of enantiospecificity (the loss in enantiopurity can be attributed to a rapid epimerization of product 2aBr under the reaction conditions). As expected, this compound turned out to be relatively unstable, resulting in the fast formation of unidentified decomposition products as well as in the elimination of HBr (giving alkene 4), which made further purifications (e.g., by column chromatography) not possible. Interestingly, however, crude (−)-2aBr can directly be reacted with NaN3 to access (−)-2aN3 under conditions similar to those established starting from (+)-2aCl already (which gave (+)-2aN3, as shown in Scheme 2).
Scheme 5. Asymmetric Sulfanylation and Stereospecific Desulfurylation–Halogenation Procedures.

The opposite sense of optical rotation (as well as HPLC retention orders) of products 2aN3 obtained via these two different approaches clearly confirm the opposite absolute configuration of (−)-2aBr relative to (+)-2aCl and (+)-2aF. In addition, it was also possible to convert (+)-2aSPh into (−)-2aCl upon treatment with SO2Cl2. This process again proceeds with good enantiospecificity, and the optical rotations (as well as HPLC retention orders) of all these products accessed by different paths now confirm that (+)-2aSPh as well as (+)-2aCl and (+)-2aF prepared by means of an α-heterofunctionalization of 1a in the presence of (R,R)-A1 or (R,R)-A2 have identical absolute configurations. This high level of catalyst-controlled face-selectivity, independent of the nature of the employed electrophile, is also in full accordance with previous observations.20−25 There, it was always found that the R,R-enantiomers of catalysts A1 and A2 efficiently block the Re-face of compounds 1 and thus favor Si-face approaches of the electrophiles (proven by single-crystal X-ray analysis for various enantiomerically enriched analogous products).20−25 Accordingly, when considering these earlier observations as well as the above-described chemical correlation, and based on our additional computational studies (vide infra), the absolute configuration of the major (+)-enantiomers of products 2Cl,F,SPh can be assigned to be S, despite of the fact that we were unfortunately not able to obtain crystals of enantioenriched products 2 suited for X-ray analysis.
Computational Studies
To better understand the origin of selectivity and to elaborate on the importance of the Maruoka catalysts A in catalyzing these reactions efficiently, we performed DFT studies on the chlorination and fluorination reactions catalyzed by A1 (as well as the slightly less selective derivative A2(53)). In addition, these calculations will help us in further supporting the proposed absolute configurations for the favored enantiomers of products 2. First, we modeled the competing major and minor enantiomeric transition state structures (TSS) for the A1-catalyzed chlorination reaction of 1c with reagent 3. In line with the outcome of our chemical correlation (vide supra) and previous observations,20−25 (R,R)-A1 efficiently favors the Si-face chlorination of starting material 1c (resulting in (S)-2cCl). The lowest-lying TSS for the major S-enantiomer was found to be favored by 2.3 kcal/mol at 298 K (Figure 2). Closer inspection of the TSS revealed that the transferring electrophilic chlorine was found to nearly be at the same distance in both TSS. However, the TS(S)major enjoys stronger hydrogen bonding interactions between the reactant fragments and the Maruoka ammonium catalyst compared to TS(R)minor.61 Distortion–interaction analysis decomposed the 2.6 kcal/mol electronic energy difference between the competing TSS into 1.1 kcal/mol of activation strain/distortion and 1.5 kcal/mol of interaction energy favoring the major enantiomer. Furthermore, decomposition of the interaction energy revealed strong electrostatic interactions (+1.8 kcal/mol) between the catalyst and the reactants favoring TS(S)major (note that this electrostatic interaction was lower for the A2-catalyzed chlorination, thus substantiating the importance of the CF3 groups53). Finally, TS(S)major was found to enjoy dispersion interactions (0.5 kcal/mol) more favorable than those of TS(R)minor.
Figure 2.

Competing enantiomeric TSS for the A1-catalyzed α-chlorination computed at PCM(toluene)-UFF:M062X/6-31+G(d,p)//PM7:B3LYP/6-31G*.
The steric cavity provided by the Maruoka catalyst was visualized by help from the SambVca algorithm (Figure 3). We observed that the TS leading to the major enantiomer suffered less steric interactions owing to a higher percentage of free volume (43.1% for TS(S)major vs 42.7% for TS(R)minor) in the cavity. Furthermore, the area affected by the steric interactions of the catalyst arms indicated by area in red (Figure 3) is smaller in TS(S)major.
Figure 3.

Buried volume plots for the major and minor enantiomeric TSS for the (R,R)-A1-catalyzed α-chlorination.
Additionally, we also investigated the α-fluorination of 1c catalyzed by (R,R)-A1. In line with the α-chlorination, computations clearly support the Si-face attack as well (favoring (S)-2cF), as TS-F-(S)major was found to be favored by 1.8 kcal/mol over the minor enantiomer (Figure 4; this energy difference corresponds to a theoretical er = 96:4, which is slightly higher than the experimental outcome (er = 93:7)). Overall, similar key interactions between the catalyst and the substrates as observed for the chlorination were identified, thus underscoring the rather general activation mode of ammonium salt A1 when used for asymmetric α-functionalizations of isoxazolidinones 1.
Figure 4.

Competing enantiomeric TSS for the A1-catalyzed α-fluorination computed at PCM(diethyl ether)-UFF:M062X/6-31+G(d,p)//PM7:B3LYP/6-31G*.
A slightly reduced cavity volume was observed in the case of the A1-catalyzed fluorination reaction (44% for TS(S)major vs 44.4% for TS(R)minor), presumably due to the slightly longer hydrogen bonding distances observed in the fluorination reaction. Based on these results, the Si-face preference by the Maruoka’s catalyst can be attributed to both the uniquely confined steric pocket generated by the arms of the biphenyl groups and the electrostatic environment generated by the electron-withdrawing substituents on the arms, as these groups were found to play a key role in modulating the hydrogen bond strength of the spirocyclic Maruoka-type catalysts.
Conclusion
A detailed experimental and computational study on the enantioselective synthesis of (masked) α-halogenated β2,2-amino acid derivatives by means of asymmetric α-halogenation strategies of α-arylisoxazolidin-5-ones 1 has been carried out. High levels of enantioselectivities were possible by carrying out the electrophilic α-chlorination and α-fluorination in the presence of Maruoka’s spirocyclic binaphthyl-based ammonium salts. Noteworthy, while the α-fluorination followed a classical α-functionalization pathway, the α-chlorination protocol was most selective when carried out as a tandem process consisting of the electrophilic α-chlorination first, followed by a direct kinetic resolution via a nucleophilic ring opening. In addition, the α-bromination was possible, as well, via an alternative strategy by carrying out an enantioselective α-sulfanylation first, followed by a stereospecific desulfurylative bromination. All of the accessed targets 2 were investigated for their potential to undergo further manipulations. Moreover, detailed accompanying mechanistic studies using DFT methods revealed the key features for the catalyst–substrate interactions and provided an explanation for the high potential of the used catalysts to facilitate reactions of substrates 1 with a broad variety of electrophiles.
Experimental Details
General Methods
1H, 13C, and 19F NMR spectra were recorded on a Bruker Avance III 300 MHz spectrometer with a broad band observation probe and a sample changer for 16 samples, a Bruker Avance DRX 500 MHz spectrometer, and on a Bruker Avance III 700 MHz spectrometer with an Ascend magnet and TCI cryoprobe, which are both property of the Austro Czech NMR Research Center “RERI uasb”. NMR spectra were referenced on the solvent peak and chemical shifts are given in parts per million.
High-resolution mass spectra (HRMS) were obtained using a Thermo Fisher Scientific LTQ Orbitrap XL with an Ion Max API source. Analyses were made in the positive ionization mode if not otherwise stated. HPLC was performed using a Thermo Scientific Dionex Ultimate 3000 or a Shimadzu Prominence system with diode array detector with a CHIRALPAK AD-H, OD-H, CHIRAL ART amylose-SA or cellulose-SB (250 × 4.6 mm, 5 μm) chiral stationary phase. Optical rotations were recorded on a Schmidt + Haensch polarimeter model UniPol L1000 at 589 nm.
All chemicals were purchased from commercial suppliers and used without further purification unless otherwise stated. Starting materials 1 were synthesized as described previously.20−22 Dry solvents were obtained from an MBraun-SPS-800 solvent purification system. All reactions were carried out under argon atmosphere unless stated otherwise.
General α-Chlorination Procedure
A flame-dried reaction vial was charged with catalyst A1 (5.6 mg, 5 mol %), 4-aryl isoxazolidin-5-one 1 (0.1 mmol, 1.0 equiv), and dry toluene (1 mL). After complete dissolution, N-chlorosuccinimide 3 (15.1 mg, 1.1 equiv) and PhONa (5.9 mg, 0.5 equiv; finely suspended in 1 mL of toluene) were added successively. The reaction mixture was layered with argon and stirred for 72 h at room temperature. Afterward, the reaction was quenched by addition of saturated NH4Cl solution and diluted with EtOAc and H2O. The aqueous phase was extracted with EtOAc (3×), and the combined organic phases were washed with brine, dried over anhydrous Na2SO4, filtered, and concentrated under reduced pressure. The crude product was subjected to flash column chromatography (silica gel, heptanes/EtOAc) to obtain products 2Cl in the given yields and enantiopurities.
(+)-2aCl:
Obtained by α-chlorination of 1a (26.2 mg, 0.100 mmol) in 52% isolated yield (15.4 mg, 0.052 mmol) with er = 95:5; Rf (heptanes/EtOAc = 5/1) = 0.37; [α]D23 = +64.2 (c = 1.0, CHCl3); 1H NMR (300 MHz, δ, CDCl3, 298 K) 7.64–7.60 (m, 2H), 7.49–7.40 (m, 3H), 4.86 (d, J = 13.2 Hz, 1H), 4.49 (d, J = 13.2 Hz, 1H), 1.51 (s, 9H); 13C NMR (75 MHz, δ, CDCl3, 298 K) 170.2, 156.1, 133.8, 130.3, 129.4 (2C), 127.2 (2C), 85.3, 64.5, 63.5, 28.2 (3C); HRMS (ESI) m/z [M + NH4]+ calcd for C14H20ClN2O4+ 315.1106; found 315.1113; HPLC (CHIRALCEL OD-H, eluent: hexanes/i-PrOH = 4/1, 0.5 mL/min, 10 °C) tr = 14.2 min (major), 16.5 min (minor).
General α-Fluorination Procedure
A flame-dried reaction vial was charged with N-fluorobenzenesulfonimide 9 (81.3 mg, 2.5 equiv), A1 (5.6 mg, 5 mol %), Cs2CO3 (48.9 mg, 1.5 equiv), and 4-arylisoxazolidin-5-one 1 (0.1 mmol, 1 equiv). Then the vial was flushed with argon, and anhydrous MTBE (6 mL) was added counter-currently to the gas flow. After the reaction mixture was stirred at room temperature for 40 h, the mixture was filtered through a bed of Na2SO4 and washed with DCM, and the solvent was removed under reduced pressure. The crude product was purified via column chromatography (silica gel, heptanes/Et2O = 2/1) to obtain products 2F in the given yields and enantiopurities.
(+)-2aF:
Obtained from 1a (26.3 mg, 0.100 mmol) in 85% NMR yield, 73% after precipitation of byproducts with cyclohexane, and 55% isolated yield after silica gel column chromatography (15.8 mg, 0.056 mmol) with er = 93:7; Rf (heptanes/Et2O = 2/1) = 0.44; [α]D22 = +12.9 (c = 0.44, CHCl3); 1H NMR (300 MHz, δ, CDCl3, 298 K) 7.48 (s, 5H), 4.62 (dd, JHF = 17.8 Hz, JHH = 13.1 Hz, 1H), 4.44 (dd, JHF = 21.9 Hz, JHH = 13.1 Hz, 1H), 1.50 (s, 9H); 13C NMR (75 MHz, δ, CDCl3, 298 K) 169.1 (d, JCF = 25.5 Hz), 155.9, 132.4 (d, JCF = 23.2 Hz), 130.5 (d, JCF = 6.6 Hz), 129.2, 125.6 (d, JCF = 6.6 Hz), 93.1 (d, JCF = 190.1 Hz), 59.9 (d, JCF = 26.7 Hz), 28.0; 19F NMR (282 MHz, δ, CDCl3, 298 K) −155.2 (dd, JFH = 21.9 Hz, JFH = 17.8 Hz); HRMS (ESI) m/z [M + NH4]+ calcd for C14H20FN2O4 299.1401; found 299.1413; HPLC (YMC CHIRAL ART Cellulose-SA, eluent: hexane/i-PrOH = 100:1, 0.5 mL/min, 10 °C) tr = 27.2 min (major), 35.2 min (minor).
Acknowledgments
We are grateful to Thomas Bögl, JKU Linz, for support with HRMS measurements.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsorginorgau.1c00025.
Author Contributions
The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
This work was supported by the Austrian Science Funds (FWF), Project No. P31784. S.C.M. acknowledges support from the XSEDE Science Gateways Program (allocation ID CHE210031), which is supported by the National Science Foundation Grant No. ACI-1548562.
The authors declare no competing financial interest.
Supplementary Material
References
- Hughes A. B., Ed. Amino Acids, Peptides and Proteins in Organic Chemistry; Wiley-VCH, 2009; Vols. 1–5. [Google Scholar]
- Lang K.; Chin J. W. Cellular Incorporation of Unnatural Amino Acids and Bioorthogonal Labeling of Proteins. Chem. Rev. 2014, 114, 4764–4806. 10.1021/cr400355w. [DOI] [PubMed] [Google Scholar]
- Maza J. C.; Jacobs T. H.; Uthappa D. M.; Young D. D. Employing Unnatural Amino Acids in the Preparation of Bioconjugates. Synlett 2016, 27, 805–813. 10.1055/s-0035-1560551. [DOI] [Google Scholar]
- Nödling A. R.; Spear L. A.; Williams T. L.; Luk L. Y. P.; Tsai Y.-H. Using genetically incorporated unnatural amino acids to control protein functions in mammalian cells. Essays Biochem. 2019, 63, 237–266. 10.1042/EBC20180042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Narancic T.; Almahboub S. A.; O’Connor K. E. Unnatural Amino Acids: Production and Biotechnological Potential. World J. Microbiol. Biotechnol. 2019, 35, 67. 10.1007/s11274-019-2642-9. [DOI] [PubMed] [Google Scholar]
- Najera C.; Sansano J. M. Catalytic Asymmetric Synthesis of α-Amino Acids. Chem. Rev. 2007, 107, 4584–4671. 10.1021/cr050580o. [DOI] [PubMed] [Google Scholar]
- Williams R. M.Synthesis of Optically Active α-Amino Acids; Perlmutter P., Ed.; Tetrahedron Organic Chemistry; Pergamon, 2016; Vol. 7. [Google Scholar]
- Abele S.; Seebach D. Preparation of Achiral and of Enantiopure Geminally Disubstituted, beta-Amino Acids for beta-Peptide Synthesis. Eur. J. Org. Chem. 2000, 2000, 1–15. . [DOI] [Google Scholar]
- Juaristi E., Soloshonok V. A., Eds. Enantioselective Synthesis of β-Amino Acids, 2nd ed.; John Wiley & Sons, 2005. [Google Scholar]
- Weiner B.; Szymanski W.; Janssen D. B.; Minnaard A. J.; Feringa B. L. Recent advances in the catalytic asymmetric synthesis of β-amino acids. Chem. Soc. Rev. 2010, 39, 1656–1691. 10.1039/b919599h. [DOI] [PubMed] [Google Scholar]
- Ashfaq M.; Tabassum R.; Ahmad M. M.; Hassan N. A.; Oku H.; Rivera G. Enantioselective Synthesis of β-amino acids: A Review. Med. Chem. 2015, 5, 295–309. 10.4172/2161-0444.1000278. [DOI] [Google Scholar]
- Noda H.; Shibasaki M. Recent Advances in the Catalytic Asymmetric Synthesis of β2- and β2,2-Amino Acids. Eur. J. Org. Chem. 2020, 2020, 2350–2361. 10.1002/ejoc.201901596. [DOI] [Google Scholar]
- Cheng R. P.; Gellman S. H.; DeGrado W. F. β-Peptides: From structure to function. Chem. Rev. 2001, 101, 3219–3232. 10.1021/cr000045i. [DOI] [PubMed] [Google Scholar]
- Lelais G.; Seebach D. β-Amino Acids-Syntheses, Occurrence in Natural Products, and Components of β-Peptides. Biopolymers 2004, 76, 206–243. 10.1002/bip.20088. [DOI] [PubMed] [Google Scholar]
- Seebach D.; Gardiner J. β-Peptidic Peptidomimetics. Acc. Chem. Res. 2008, 41, 1366–1375. 10.1021/ar700263g. [DOI] [PubMed] [Google Scholar]
- Seebach D.; Beck A. K.; Capone S.; Deniau G.; Grošelj U.; Zass E. Enantioselective Preparation of β2-Amino Acid Derivatives for β-Peptide Synthesis. Synthesis 2009, 2009, 1–32. 10.1055/s-0028-1087490. [DOI] [Google Scholar]
- Cabrele C.; Martinek T. A.; Reiser O.; Berlicki Ł. Peptides Containing β-Amino Acid Patterns: Challenges and Successes in Medicinal Chemistry. J. Med. Chem. 2014, 57, 9718–9739. 10.1021/jm5010896. [DOI] [PubMed] [Google Scholar]
- Tite T.; Sabbah M.; Levacher V.; Brière J.-F. Organocatalysed decarboxylative protonation process from Meldrum’s acid: enantioselective synthesis of isoxazolidinones. Chem. Commun. 2013, 49, 11569–11571. 10.1039/c3cc47695b. [DOI] [PubMed] [Google Scholar]
- Macchia A.; Eitzinger A.; Brière J.-F.; Waser M.; Massa A. Asymmetric Synthesis of Isoxazol-5-ones and Isoxazolidin-5-ones. Synthesis 2021, 53, 107–122. 10.1055/s-0040-1706483. [DOI] [Google Scholar]
- Cadart T.; Berthonneau C.; Levacher V.; Perrio S.; Brière J.-F. Enantioselective Phase-Transfer Catalyzed α-Sulfanylation of Isoxazolidin-5-ones: An Entry to β2,2-Amino Acid Derivatives. Chem. - Eur. J. 2016, 22, 15261–15264. 10.1002/chem.201603910. [DOI] [PubMed] [Google Scholar]
- Capaccio V.; Sicignano M.; Rodríguez R. I.; Della Sala G.; Alemán J. Asymmetric Synthesis of α-Trifluoromethylthio-β-Amino Acids under Phase Transfer Catalysis. Org. Lett. 2020, 22, 219–223. 10.1021/acs.orglett.9b04195. [DOI] [PubMed] [Google Scholar]
- Eitzinger A.; Brière J. F.; Cahard D.; Waser M. Enantioselective Catalytic Synthesis of α-Aryl-α-SCF3-β2,2-Amino Acids. Org. Biomol. Chem. 2020, 18, 405–408. 10.1039/C9OB02666E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cadart T.; Levacher V.; Perrio S.; Brière J.-F. Construction of Isoxazolidin-5-ones with a Tetrasubstituted Carbon Center: Enantioselective Conjugate Addition Mediated by Phase-Transfer Catalysis. Adv. Synth. Catal. 2018, 360, 1499–1509. 10.1002/adsc.201800009. [DOI] [Google Scholar]
- Capaccio V.; Zielke K.; Eitzinger A.; Massa A.; Palombi L.; Faust K.; Waser M. Asymmetric phase-transfer catalysed β-addition of isoxazolidin-5-ones to MBH carbonates. Org. Chem. Front. 2018, 5, 3336–3340. 10.1039/C8QO01057A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eitzinger A.; Winter M.; Schörgenhumer J.; Waser M. Quaternary β2,2-Amino Acid Derivatives by Asymmetric Addition of Isoxazolidin-5-ones to para-Quinone Methides. Chem. Commun. 2020, 56, 579–582. 10.1039/C9CC09239K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu J.-S.; Noda H.; Shibasaki M. Exploiting β-Amino Acid Enolates in Direct Catalytic Diastereo- and Enantioselective C–C Bond-Forming Reactions. Chem. - Eur. J. 2018, 24, 15796–15800. 10.1002/chem.201804346. [DOI] [PubMed] [Google Scholar]
- Amemiya F.; Noda H.; Shibasaki M. Lewis Base Assisted Lithium Brønsted Base Catalysis: A New Entry for Catalytic Asymmetric Synthesis of β2,2-Amino Acids. Chem. Pharm. Bull. 2019, 67, 1046–1049. 10.1248/cpb.c19-00569. [DOI] [PubMed] [Google Scholar]
- Yu J.-S.; Noda H.; Shibasaki M. Quaternary β2,2-Amino Acids: Catalytic Asymmetric Synthesis and Incorporation into Peptides by Fmoc-Based Solid-Phase Peptide Synthesis. Angew. Chem., Int. Ed. 2018, 57, 818–822. 10.1002/anie.201711143. [DOI] [PubMed] [Google Scholar]
- Nascimento de Oliveira M.; Arseniyadis S.; Cossy J. Palladium-Catalyzed Asymmetric Allylic Alkylation of 4-Substituted Isoxazolidin-5-ones: Straightforward Access to β2,2-Amino Acids. Chem. - Eur. J. 2018, 24, 4810–4814. 10.1002/chem.201800641. [DOI] [PubMed] [Google Scholar]
- Yu J.-S.; Espinosa M.; Noda H.; Shibasaki M. Traceless Electrophilic Amination for the Synthesis of Unprotected Cyclic β-Amino Acids. J. Am. Chem. Soc. 2019, 141, 10530–10537. 10.1021/jacs.9b05476. [DOI] [PubMed] [Google Scholar]
- Espinosa M.; Noda H.; Shibasaki M. Synthesis of Unprotected Spirocyclic β-Prolines and β-Homoprolines by Rh-Catalyzed C–H Insertion. Org. Lett. 2019, 21, 9296–9299. 10.1021/acs.orglett.9b03198. [DOI] [PubMed] [Google Scholar]
- 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. 10.1002/ejoc.202001077. [DOI] [Google Scholar]
- Avenoza A.; Busto J. H.; Corzana F.; Jimenez-Oses G.; Peregrina J. M. SN2 vs. E2 on quaternary centres: an application to the synthesis of enantiopure β2,2-amino acids. Chem. Commun. 2004, 980–981. 10.1039/B400282B. [DOI] [PubMed] [Google Scholar]
- Edmonds M. K.; Graichen F. H. M.; Gardiner J.; Abell A. D. Enantioselective Synthesis of α-Fluorinated β2-Amino Acids. Org. Lett. 2008, 10, 885–887. 10.1021/ol703045z. [DOI] [PubMed] [Google Scholar]
- Peddie V.; Pietsch M.; Bromfield K. M.; Pike R. N.; Duggan P. J.; Abell A. D. Fluorinated β2- and β3-Amino Acids: Synthesis and Inhibition of α-Chymotrypsin. Synthesis 2010, 2010, 1845–1859. 10.1055/s-0029-1218743. [DOI] [Google Scholar]
- Sun A. W.; Hess S. N.; Stoltz B. M. Enantioselective synthesis of gem-disubstituted N-Boc diazaheterocycles via decarboxylative asymmetric allylic alkylation. Chem. Sci. 2019, 10, 788–792. 10.1039/C8SC03967D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lakner F. J.; Hager L. P. Chloroperoxidase-mediated asymmetric epoxidation. Synthesis of (R)-dimethyl 2-methvlaziridine-1,2-dicarboxvlate—a potential α-methylamino acid synthon. Tetrahedron: Asymmetry 1997, 8, 3547–3550. 10.1016/S0957-4166(97)00467-9. [DOI] [Google Scholar]
- Shirakawa S.; Maruoka K. Recent Developments in Asymmetric Phase-Transfer Reactions. Angew. Chem., Int. Ed. 2013, 52, 4312–4348. 10.1002/anie.201206835. [DOI] [PubMed] [Google Scholar]
- Tan J.; Yasuda N. Contemporary Asymmetric Phase Transfer Catalysis: Large-Scale Industrial Applications. Org. Process Res. Dev. 2015, 19, 1731–1746. 10.1021/acs.oprd.5b00304. [DOI] [Google Scholar]
- Kaneko S.; Kumatabara Y.; Shirakawa S. A new generation of chiral phase-transfer catalysts. Org. Biomol. Chem. 2016, 14, 5367–5376. 10.1039/C5OB02446C. [DOI] [PubMed] [Google Scholar]
- Schörgenhumer J.; Tiffner M.; Waser M. Chiral Phase-Transfer Catalysis in the Asymmetric α-Heterofunctionalization of Prochiral Nucleophiles. Beilstein J. Org. Chem. 2017, 13, 1753–1769. 10.3762/bjoc.13.170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qian D.; Sun J. Recent Progress in Asymmetric Ion-Pairing Catalysis with Ammonium Salts. Chem. - Eur. J. 2019, 25, 3740–3751. 10.1002/chem.201803752. [DOI] [PubMed] [Google Scholar]
- Žukauskaitė A.; Mangelinckx S.; Šačkus A.; De Kimpe N. Synthesis of Alkyl 3-Chloroazetidine-3-carboxylates via Regioselective Ring Transformation of Alkyl 2-(Bromomethyl)aziridine-2-carboxylates. Heterocycles 2014, 88, 731–740. 10.3987/COM-13-S(S)35. [DOI] [Google Scholar]
- Pu X.-Q.; Zhao H.-Y.; Lu Z.-H.; He X.-P.; Yang X.-J. Aminochlorination of Alkenes with CFBSA. Eur. J. Org. Chem. 2016, 2016, 4526–4533. 10.1002/ejoc.201600709. [DOI] [Google Scholar]
- Shibatomi K.; Soga Y.; Narayama A.; Fujisawa I.; Iwasa S. Highly Enantioselective Chlorination of β-Keto Esters and Subsequent SN2 Displacement of Tertiary Chlorides: A Flexible Method for the Construction of Quaternary Stereogenic Centers. J. Am. Chem. Soc. 2012, 134, 9836–9839. 10.1021/ja304806j. [DOI] [PubMed] [Google Scholar]
- Shibatomi K.; Kotozaki M.; Sasaki N.; Fujisawa I.; Iwasa S. Williamson Ether Synthesis with Phenols at a Tertiary Stereogenic Carbon: Formal Enantioselective Phenoxylation of β-Keto Esters. Chem. - Eur. J. 2015, 21, 14095–14098. 10.1002/chem.201502042. [DOI] [PubMed] [Google Scholar]
- Liu R. Y.; Wasa M.; Jacobsen E. N. Enantioselective synthesis of tertiary α-chloro esters by non-covalent catalysis. Tetrahedron Lett. 2015, 56, 3428–3430. 10.1016/j.tetlet.2015.01.124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shibatomi K.; Narayama A. Catalytic Enantioselective α-Chlorination of Carbonyl Compounds. Asian J. Org. Chem. 2013, 2, 812–823. 10.1002/ajoc.201300058. [DOI] [Google Scholar]
- Gomez-Martinez M.; Alonso D. A.; Pastor I. M.; Guillena G.; Baeza A. Organocatalyzed Assembly of Chlorinated Quaternary Stereogenic Centers. Asian J. Org. Chem. 2016, 5, 1428–1437. 10.1002/ajoc.201600404. [DOI] [Google Scholar]
- Ooi T.; Kameda M.; Maruoka K. Molecular Design of a C2-Symmetric Chiral Phase-Transfer Catalyst for Practical Asymmetric Synthesis of α-Amino Acids. J. Am. Chem. Soc. 1999, 121, 6519–6520. 10.1021/ja991062w. [DOI] [PubMed] [Google Scholar]
- Godemert J.; Oudeyer S.; Levacher V. Chiral Ammonium Aryloxides: Efficient Multipurpose Basic Organocatalysts. ChemCatChem 2016, 8, 74–85. 10.1002/cctc.201500616. [DOI] [Google Scholar]
- Straub M. R.; Birman V. B. Organocatalytic Kinetic Resolution of N-Boc-Isoxazolidine-5-ones. Org. Lett. 2021, 23, 984–988. 10.1021/acs.orglett.0c04196. [DOI] [PubMed] [Google Scholar]
- Further details can be found in the online Supporting Information.
- Lima C. G. S.; Pauli F. P.; Costa D. C. S.; de Souza A. S.; Forezi L. S. M.; Ferreira V. F.; de Carvalho da Silva F. para-Quinone Methides as Acceptors in 1,6-NucleophilicConjugate Addition Reactions for the Synthesis of Structurally Diverse Molecules. Eur. J. Org. Chem. 2020, 2020, 2650–2692. 10.1002/ejoc.201901796. [DOI] [Google Scholar]
- Wang J.-Y.; Hao W.-J.; Tu S.-J.; Jiang B. Recent developments in 1,6-addition reactions of para-quinone methides (p-QMs). Org. Chem. Front. 2020, 7, 1743–1778. 10.1039/D0QO00387E. [DOI] [Google Scholar]
- Bode J. W.; Fox R. M.; Baucom K. D. Chemoselective Amide Ligations by Decarboxylative Condensations of N-Alkylhydroxylamines and α-Ketoacids. Angew. Chem., Int. Ed. 2006, 45, 1248–1252. 10.1002/anie.200503991. [DOI] [PubMed] [Google Scholar]
- Wucherpfennig T. G.; Pattabiraman V. R.; Limberg F. R. P.; Ruiz-Rodriguez J.; Bode J. W. Traceless Preparation of C-Terminal α-Ketoacids for Chemical Protein Synthesis by α-Ketoacid–Hydroxylamine Ligation: Synthesis of SUMO2/3. Angew. Chem., Int. Ed. 2014, 53, 12248–12252. 10.1002/anie.201407014. [DOI] [PubMed] [Google Scholar]
- Canestrari D.; Lancianesi S.; Badiola E.; Strinna C.; Ibrahim H.; Adamo M. F. A. Desulfurative Chlorination of Alkyl Phenyl Sulfides. Org. Lett. 2017, 19, 918–921. 10.1021/acs.orglett.7b00077. [DOI] [PubMed] [Google Scholar]
- Canestrari D.; Cioffi C.; Biancofiore I.; Lancianesi S.; Ghisu L.; Ruether M.; O’Brien J.; Adamo M. F. A.; Ibrahim H. Sulphide as a leaving group: highly stereoselective bromination of alkyl phenyl sulphides. Chem. Sci. 2019, 10, 9042–9050. 10.1039/C9SC03560E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alletto F.; Adamo M. F. A. Enantiospecific on-water bromination: a mild and efficient protocol for the preparation of alkyl bromides. Green Chem. 2020, 22, 8692–8698. 10.1039/D0GC02855J. [DOI] [Google Scholar]
- Bencivenni G.; Salazar Illera D.; Moccia M.; Houk K. N.; Izzo J. A.; Novacek J.; Grieco P.; Vetticatt M. J.; Waser M.; Adamo M. F. A. Study of ground state interactions of enantiopure chiral quaternary ammonium salts and amides, nitroalkanes, nitroalkenes, esters, heterocycles, ketones and fluoroamides. Chem. - Eur. J. 2021, 27, 11352–11366. 10.1002/chem.202100908. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.

