Next Article in Journal
Nutritional Composition, Efficacy, and Processing of Vigna angularis (Adzuki Bean) for the Human Diet: An Overview
Next Article in Special Issue
Coinage Metal-Catalyzed Asymmetric Reactions of ortho-Alkynylaryl and Heteroaryl Aldehydes and Ketones
Previous Article in Journal
Antioxidant and Hypoglycemic Activity of Sequentially Extracted Fractions from Pingguoli Pear Fermentation Broth and Identification of Bioactive Compounds
Previous Article in Special Issue
A Plausible Mechanism for the Iridium-Catalyzed Hydrogenation of a Bulky N-Aryl Imine in the (S)-Metolachlor Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of Pyrrolidinyl Ferrocene-Containing Ligands and Their Application in Highly Enantioselective Rhodium-Catalyzed Olefin Hydrogenation

1
Centre for Synthesis and Chemical Biology, School of Chemistry, University College Dublin, Belfield, D04 N2E5 Dublin, Ireland
2
MSD Ballydine, Kilsheelan, Co. Tipperary, E91 V091 Clonmel, Ireland
3
Synthesis and Solid State Pharmaceutical Centre, School of Chemistry, University College Dublin, Belfield, D04 N2E5 Dublin, Ireland
*
Author to whom correspondence should be addressed.
Dedicated with respect and admiration to Professor Henri Kagan, a pioneer in asymmetric catalysis.
Molecules 2022, 27(18), 6078; https://doi.org/10.3390/molecules27186078
Submission received: 13 August 2022 / Revised: 12 September 2022 / Accepted: 13 September 2022 / Published: 17 September 2022

Abstract

:
Herein, we report the design and synthesis of a series of chiral pyrrolidine-substituted ferrocene-derived ligands. The proficiency of this novel structural motif was demonstrated in the Rh-catalyzed asymmetric hydrogenation of dehydroamino acid esters and α-aryl enamides. The products were obtained with full conversions and excellent levels of enantioselectivities of up to >99.9% ee and 97.7% ee, respectively, using a BINOL-substituted phosphine-phosphoaramidite ligand which possesses planar, central, and axial chirality elements.

Graphical Abstract

1. Introduction

Phosphorous-based ligands have found extensive use in homogeneous transition metal catalysis [1,2,3]. Bidentate P,P based structures, exemplified by BINAP, DuPhos and Josiphos, represent a privileged ligand scaffold in transition-metal catalysis (Figure 1) [4]. Ferrocenyl-based Josiphos derivatives have proven to be efficient ligands across a range of enantioselective processes, particularly in the field of olefin hydrogenations [5,6]. Their value has been demonstrated in the large-scale Ir-catalyzed hydrogenation of an N-aryl imine in the synthesis of the herbicide (S)-metolachlor [7]. Thus, the increasing demand for efficient enantioselective technologies has led to significant investigation of ligand derivatives, as evidenced by the range of commercially available variants and sub-families such as Knochel’s Ferriphos ligands [8].
Focusing on planar chiral ferrocene compounds and their ease of preparation through the use of diastereoselective ortho-metalating groups, including amines, sulfoxides, acetals, oxazolines, azepines, sulfoximines, and hydrazones, has contributed to the wide range of planar chiral ferrocene ligands reported to date [9]. We reported the preparation of ferrocenylphosphinamine ligands of type 1, Figure 2, possessing both planar and central chirality obtained through diastereoselective metalation of trans-(2R,5R)-2,5-dialkyl-1-(ferrocenylmethyl)pyrrolidines and their application in Pd-catalysed allylic alkylation [10].
Thus, interested by the successful application of ferrocene-containing ligands, bearing an α-chiral center, and extending our work on pyrrolidine-containing P,N ligands [11,12,13], we developed novel ferrocene ligands in which the α-chiral center is incorporated into a pyrrolidine unit. We had previously reported the enantioselective preparation of ferrocenepyrrolidine (R)-2 and applied it in the diastereoselective formation of a series of N,O ligands of type 3 for the diethylzinc-mediated addition to aldehydes, affording enantioselectivities of up to 95% ee [14]. In addition, we reported the synthesis of novel ferrocene-phosphinamine ligands of type 4, again obtained through diastereoselective ortho-metalation of ferrocenepyrrolidine (R)-2, Figure 2, and their application in the Pd-catalyzed allylic alkylation of 1,3-diphenylprop-2-enyl acetate with dimethyl malonate gave enantioselectivities of up to 77% ee [15].
In 2002, Boaz reported the synthesis of air-stable ‘BoPhoz’ ferrocenylphosphine-aminophosphine ligands (5, Figure 3), which were highly effective in the asymmetric hydrogenation of dehydro—α-amino acids, itaconic acids, and α-ketoesters [16,17]. Two different phosphorus donor atoms generate electronic asymmetry at the metal centre which then provides unique modulation of the catalyst activity. Derivatization of the ligands led to an expansion of the range of suitable substrates for hydrogenation [18,19,20] and identified the proficiency of the ligands in the catalytic asymmetric synthesis of cyclohexenone-based atropisomers [21]. Chang and Zheng introduced a phosphine-phosphoramidite scaffold (6, Figure 3) which proved effective in the asymmetric hydrogenation of a broad range of substrates including both (Z)- and (E)-aryl and β-alkyl-β-(acylamino)acrylates [22,23,24]. With the success of the reported ligand derivatizations in mind, it occurred to us that the flexible amino sidechain in 5 and 6 could be modified to introduce the more rigid pyrrolidinyl motif found in 3 and 4. Herein, we present the preparation of a variety of ferrocenylphosphine-aminophosphine and ferrocenylphosphine-phosphoramidite ligands of type 7 and 8. The relatively facile introduction of the amino-phosphine moiety introduces a modular element, which enabled the preparation of a diverse range of novel ligands for investigation.

2. Results and Discussion

2.1. Synthesis and Characterization of N-Phosphorus-Substituted Pyrrolidine Based Ligands

The strategy for the preparation of ferrocenyl N-phosphinepyrrolidinyl ligands was adapted from our previous work on the ferrocene-phosphinamine ligands of type 4 [15]. The synthesis of pyrrolidine (R)-13 from ferrocene (9), consisted of a Friedel-Crafts acylation to afford ketone (10), a Corey-Bakshi-Shibata oxazoborolidine-mediated enantioselective reduction to afford alcohol (R)-11, its acetylation to afford acetate (R)-12 and finally treatment with allylamine to give (R)-13 in a 92% yield over four steps (Scheme 1).
The introduction of the ferrocenyl phosphine moiety was accomplished by non-selective ortho-lithiation of (R)-13, which was then quenched with chlorodiphenylphosphine (Scheme 2). Diastereomers (R,Sp)-14 and (R,Rp)-14 were separated by silica gel column chromatography and isolated in yields of 24% and 33%, respectively. To facilitate the introduction of the N-pyrrolidinyl phosphine substituent, deallylation of the amine was accomplished using palladium(tetrakistriphenylphosphine) and N,N-dimethyl barbituric acid (NDMBA) [25]. Deprotected pyrrolidines (R,Sp)-15 and (R,Rp)-15 were isolated in yields of 98% and 86%, respectively. With the synthetic precursors to the desired ligands in hand, the final coupling could now be performed using triethylamine and the appropriate chlorodiarylphosphine. The initial series of ligands L16 were synthesized using (R,Sp)-15. Ligands bearing neutral (L1), electron rich (L2), and electron poor (L3L4) aryl groups were synthesized in moderate to good yields (40–78%). The proficiency of a BINOL unit in ferrocenyl bisphosphonate [26], phosphoramidite [27,28], and phosphine-phosphoramidite ligands [24] for enantioselective rhodium-catalyzed hydrogenation has been well-documented. Therefore, phosphoramidites L5 and L6 were synthesized using (R)- or (S)-1,1′-binaphthyl-2,2′-diyl phosphorochloridate in good yields of 83 and 70%, respectively. In order to test the effect of a switch in planar chirality, isomeric ligands L7 and L8 were synthesized from (R,Rp)-15 in excellent yields of 90% and 98%, respectively (Figure 4).
Interestingly, 5J coupling between the phosphorus atoms was observed for ligands L1L6 with coupling constants ranging from 16.4 to 56.2 Hz. Cross peaks in the two-dimensional 31P{1H}–31P{1H} spectra (31P COSY) provided further evidence for this unexpected interaction (Figure 5). Although ‘long range couplings’ (across more than four bonds) between two phosphorus atoms are quite rare, the phenomenon has been observed with several ligands, such as Xantphos (6J coupling) and tetraphosphine ferrocenyl derivatives [29]. An excellent in-depth study of the ferrocenyl compounds attributed the 31P–31P nuclear spin-spin coupling to a through-space non-bonded interaction of the phosphorus lone pairs, as was previously observed in 19F–19F and 15N–19F couplings [30]. Due to the magnitude of the coupling constants observed for L1L6, it is unlikely a through-bond interaction (σ- and π-transmitted components) is taking place. Through-space coupling results from overlap of the phosphorus lone-pair orbitals. Although the interaction provides an adequate pathway to transmit spin information between the nuclei, it does not lead to chemical bonding because both orbitals are occupied.
A comparison of the JPP coupling constants in L1L8 revealed a considerable effect resulting from the choice of aryl phosphine-substitution (Table 1). The magnitude of the coupling constant depends on the extent of the lone-pair overlap, providing some information on the P-P orientation and proximity in solution. No coupling was observed for L7L8 indicating the planar and central chirality of the ligand prohibit favorable alignment of the lone pairs. These factors could provide useful information on the bite angle and play an important role in the future development and application of this class of ligand.
X-ray crystallographic analysis of L5 provided confirmation of the assignment of the planar, central, and axial chirality in the molecule; 15% oxidation of the ferrocenyl phosphorous atom (P1) was observed (Figure 6). The distance between the phosphorus atoms was measured at 3.9 Å, which favorably compared to Meunier’s observation of coupling constants over 20 Hz with P–P distances of 4.0 Å or below [30]. However, the solid-state conformation of L5 obtained from X-ray structural data should not reflect the conformation in solution due to effects of crystal-packing, meaning the true inter-phosphorus distance remains unknown.

2.1.1. Rhodium-Catalyzed Asymmetric Hydrogenation of Dehydroamino Acid Esters

Transition metal-catalyzed asymmetric hydrogenation of dehydroamino acid derivatives is a valuable method for the preparation of amino acid precursors [31]. As such, the field has undergone extensive investigation since the 1970s, from the seminal work of Knowles and Kagan [32,33,34], thereby providing a convenient test reaction to test the efficiency of the rhodium complexes of our novel ligands L1L8.

Reaction Condition Optimization

(Z)-Methyl-2-acetamido-3-phenylacrylate 16a was chosen as a suitable substrate to test the application of ligands L1L8 (Table 2). Initial conditions consisted of Rh(COD)2OTf as the rhodium source, with ligand L1 in THF at room temperature under an atmosphere of hydrogen for 12 h. Full conversion of the starting material was observed with an ee of 89.0 % ee for the product (Table 2, entry 2). When Rh(COD)2BF4 was employed a drop in the conversion of starting material was observed, although the product was formed in a similar ee (Table 2, entry 3). This effect was subsequently observed throughout the optimization process (Table 2, entries 11, 18). Increasing the hydrogen pressure was found to have a detrimental effect upon the ee of the product (Table 2, entries 4-5). Similarly, variation of the solvent had a deleterious effect upon the ee and in some cases, the conversion (Table 2, entries 6–10). Poor asymmetric induction and conversion were observed upon switching to ligand L2, which bears an o-tolyl substituted phosphine (Table 2, entries 12–13). The comparatively electron-poor p-fluoro substituted L3 gave a similar result to that obtained previously with ligand L1 (Table 2, entry 14). Switching to 3,5-di(trifluoromethyl)-substituted ligand L4, the product was formed in 92.2% ee with full conversion after 2 h, and the ee was further increased to 95.5% upon cooling the reaction mixture to 0 °C for 3 h (Table 2, entries 15–16). An (R)-BINOL-based system was next evaluated in the application of ligand L5. Gratifyingly, full conversion of 16a was found in 12 h, with the product formed in an excellent enantiomeric excess of 97.9% (Table 2, entry 18). In this case, upon repetition with Rh(COD)2BF4, no decrease in the conversion of the starting material was observed (Table 2, entry 19). By increasing the hydrogen pressure to 10 bar, full conversion of the starting material was observed within 2 h, but the ee of the product dropped slightly (Table 2, entry 20). Variation of the solvent did not have a beneficial effect upon the enantioselectivity of the reaction (Table 2, entries 21–22). Interestingly, changing the axial chirality using (S)-BINOL in ligand L6 resulted in a significantly lower ee for the product (Table 2, entry 23). The best result was obtained by switching the planar chirality, as illustrated in ligand L7, where full conversion of the starting material was observed in one hour and an ee of 99.5% was obtained (Table 2, entry 24). By increasing the hydrogen pressure to 20 bar and the reaction time to 4 h, the catalyst loading could be significantly lowered to only 0.02 mol %, with minimal effect on the enantioselectivity of product formation (Table 2, entry 25). Interestingly, in contrast to the planar isomers (ligand L5 and L6), when the opposite hand of BINOL was used with ligand L8, the opposite enantiomer of the product was formed, albeit with a much lower ee of 46.7% (Table 2, entry 26). For comparison purposes, the results obtained in the literature using rhodium complexes of Josiphos, DuPhos, and BoPhoz ligands are included (Table 1, entries 27–29). Our optimal ligand (99.5% ee) compares favorably with DuPhos (85% ee) and Josiphos (96% ee) and is identical to the level of enantioselectivity exhibited by the BoPhoz ligand (99.5% ee).

Substrate Scope

With optimum conditions in hand, the activity of a selection of ligands was investigated with a variety of amino acid precursors (Scheme 3). As the best ligand from the optimization process, L7 consistently gave excellent results of full conversion and over 99% ee regardless of any electron-rich (16b, 16c, 16j) or electron-poor substituents (16d16i) on the β-aryl moiety or the particular substitution pattern of the substrate. The excellent performance also extended to a dehydroamino acid ester without aryl substitution (16k). Phenyl-substituted ligand L1 was also tested across a selection of the substrates and gave consistent levels of asymmetric induction (88.5–92.0% ee) but variable conversions of the starting material from (60–>99%). The rhodium complex of ligand L4 bearing 3,5-di(trifluoromethyl)phenyl substituents also led to product formation, with full conversions and remarkably consistent, although lower, enantioselectivities regardless of the starting material (92.5–96.3% ee).

2.1.2. Rhodium-Catalyzed Asymmetric Hydrogenation of α-Aryl Enamides

With the success of the rhodium complexes in the highly enantioselective reduction of dehydroamino acids, the efficiency of the ligands was next tested in the rhodium-catalyzed hydrogenation of a selection of α-aryl enamides. This is a valuable process for the construction of a variety of amines.

Reaction Condition Optimization

N-(1-phenylvinyl)acetamide 18a was chosen as the model substrate for optimization studies. Ligand L1 was tested using Rh(COD)2OTf in THF at room temperature with 40 bar hydrogen pressure for 2 h. While full conversion of the starting material was observed, a disappointing ee for the product of 33.6% (S) was observed (Table 3, entry 2). Decreasing the hydrogen pressure did not have a significant effect on the level of asymmetric induction (Table 3, entry 3). Shortening the reaction time to 1 h and switching to MeOH as solvent provided a slight increase to 46.0% ee (Table 3, entry 5). While ligands L2 and L3 were similarly ineffective, a dramatic improvement was observed upon using the 3,5-di(trifluoromethyl)-substituted ligand L4 with an ee of 91.3% (Table 3, entries 6–8).
Switching to ligand L5 in THF at 10 bar hydrogen provided the product with similar levels of asymmetric induction (Table 3, entry 9). Decreasing the pressure further to 1 bar resulted in decreased conversion of the starting material (Table 3, entry 10). As in our previous optimization, the best result was observed with ligand L7, with an ee of 96.4% observed and full conversion after 2.5 h (Table 3, entry 11). A similar ee was observed with 10 bar hydrogen pressure after 1 h while decreasing the catalyst loading to 0.2 mol % provided the product in 91.0 % ee (Table 3, entries 12-13). Once again, switching the axial chirality of the BINOL moiety provided the opposite enantiomeric product in a much-lowered ee of 31.7 % (Table 3, entry 14). For comparison purposes, the results obtained in the literature using rhodium complexes of Me-BPE (related to DuPhos) and BoPhoz ligands are included (Table 2, entries 15–16). Our optimal ligand (96.4% ee) compares closely with Me-BPE (95.2% ee) but outperforms the level of enantioselectivity exhibited by the BoPhoz ligand (61.8% ee).

Substrate Scope

The optimized conditions with ligand L7 were tested across a range of α-aryl enamides (18af) and, in contrast to the investigation of dehydroamino acids, the results were substrate-dependent (Scheme 4). Naphthyl (18b) and 4-chloro (18c) substitution of the aryl ring resulted in full conversions and excellent enantioselectivities (97.4–97.7% ee), but a lower ee of 93.7% was observed with a 4-methoxy substituent (18d). Ligand L4 was also tested in the synthesis of 19b and 19c and the products were formed in slightly lower enantioselectivities (92.0–92.4% ee), while a significant drop in the level of asymmetric induction was seen for the formation of 19d (77.6% ee). Poor conversion of the starting material and low level of asymmetric induction were observed in the hydrogenation of bicyclic N-(3,4-dihydro-1-naphthyl)acetamide 18e with ligand L7 (23% conversion, 48.0% ee) and ligand L4 (12% conversion, 27.0% ee). A similarly poor result was observed in the hydrogenation of β-phenyl- β- (acylamino)acrylate 18f with ligand L7 (18% conversion, 15.3% ee). However, a drastic improvement to full conversion to the product with an ee of 61.3% was observed with ligand L1. The result demonstrates the need for substrate dependent optimization with a new class of substrate to maximize the potential of the ligand series developed.

3. Materials and Methods—Chemistry

Unless otherwise noted, reactions were performed with rigorous exclusion of air and moisture under an inert atmosphere of nitrogen in flame-dried glassware with magnetic stirring using anhydrous solvents. N2-flushed stainless steel cannulas or plastic syringes were used to transfer air and moisture-sensitive reagents. All reagents were obtained from commercial sources and used without further purification unless otherwise stated. All anhydrous solvents were obtained from commercial sources (Sigma Aldrich, Glasgow, United Kingdom) and used as received with the following exceptions: diethyl ether (Et2O), dichloromethane (CH2Cl2), and toluene (PhCH3) were dried by passing through activated alumina columns. Powdered activated 4 Å molecular sieves were purchased from Sigma Aldrich (Glasgow, United Kingdom)and were stored in an oven at 120 °C. In vacuo refers to the evaporation of solvent under reduced pressure on a rotary evaporator. Thin-layer chromatography (TLC) was performed on aluminium plates pre-coated with silica gel F254 (Merck, Darmstadt, Germany). They were visualised with UV-light (254 nm) fluorescence quenching, or by charring with Hanessian’s staining solution (cerium molybdate, H2SO4 in water), basic potassium permanganate staining solution (potassium permanganate, K2CO3 and NaOH in water), or an acidic vanillin staining solution (vanillin, H2SO4 in ethanol). Flash column chromatography was carried out using 40–63 μm, 230–400 mesh silica gel.
1H NMR spectra were recorded on a 300-, 400-, or 500-MHz spectrometer. 13C NMR spectra were recorded on a 400- or 500-MHz spectrometer (Agilent, Birmingham, United Kingdom) at 101 or 126 MHz. 19F NMR spectra were recorded on a 400-MHz spectrometer at 470 MHz. Chemical shifts (δ) are reported in parts per million (ppm) downfield from tetramethylsilane and for 1H NMR are referenced to residual proton in the NMR solvent (CDCl3 = δ 7.26 ppm). 13C NMR are referenced to the residual solvent peak (CDCl3 = δ 77.16 ppm). All 13C spectra are 1H decoupled. NMR data are represented as follows: chemical shift (δ ppm), integration, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, dd = double doublet, m = multiplet, app. d = apparent doublet, app. t. = apparent triplet), coupling constant (J) in Hertz (Hz). High resolution mass spectra [electrospray ionisation (ESI-TOF)] (HRMS) were measured on a micromass LCT orthogonal time-of-flight mass spectrometer with leucine enkephalin (Tyr-Gly-Phe-Leu) as an internal lock mass. Infrared spectra were recorded on a FT-IR spectrometer and are reported in terms of wavenumbers (νmax) with units of reciprocal centimetres (cm−1). Microwave experiments were conducted in a CEM Discover S-class microwave reactor with controlled irradiation at 2.45 GHz using standard microwave process Pyrex vials. Reaction time reflects time at the set reaction temperature maintained by cycling of irradiation (fixed hold times). Optical rotation (α) values were measured at room temperature and specific rotation ([α]D20) values are given in deg.dm−1.cm3.g−1. Melting points were determined in open capillary tubes. HPLC analysis was carried out on a Shimadzu LC-10AT vp machine and Schimadzu LC-2010A machine equipped with a UV-Vis detector employing Chiracel® OD (Sigma Aldrich) and AD columns from Diacel Chemical Industries (Illkirch, France).

3.1. 4-Chloro-Ferrocenylbutanone (10)

Ferrocene 9 (12.7 g, 68 mmol) was added to a dry 500-mL two-necked room-bottom flask (RBF) containing a magnetic stir bar under an inert atmosphere. Dry CH2Cl2 (120 mL) was added to the reaction flask which was cooled to 0 °C. 4-Chlorobutyryl chloride (95%, 7.3 mL, 62 mmol) was added to the reaction mixture followed by the slow addition of aluminium chloride (9.9 g, 74 mmol). The reaction mixture was warmed to room temperature and stirred for 18 h. Ice-cold H2O (100 mL) was added to reaction mixture followed by 10% Na2S2O4 solution (100 mL). The mixture was stirred for 30 min and the aqueous layer was extracted with CH2Cl2 (4 × 50 mL). The combined organic layers were washed with NaOH (2 M, 100 mL) and brine (100 mL), and dried with anhydrous Na2SO4. The solvent was removed in vacuo, and the crude product was purified by silica gel column chromatography (pentane/EtOAc) to yield 10 as an orange solid (17.3 g, 96%). Spectroscopic data are in good accordance to literature [14].

3.2. (R)-4-Chloro-1-Ferrocenylbutanol ((R)-11)

BH3.THF (3 mL, 1.0 M, 3 mmol) was added to (S)-(−)-2-methyl-CBS-oxazaborolidine (crude residue) in a 250 mL Schlenk flask containing a magnetic stir bar under an inert atmosphere, and the reaction flask was cooled to −55 °C. A solution of 4-Chloro-ferrocenylbutanone (11) (3.67 g, 12.6 mmol) in dry THF (90 mL) was added followed by another portion of BH3.THF (1.0 M, 6 mL, 6 mmol). The reaction mixture was stirred for 18 h. The reaction mixture was warmed to 0 °C and then quenched by slow dropwise addition of MeOH (20 mL). The solvent was removed in vacuo, and the crude product was purified by silica gel column chromatography (pentane/EtOAc) to yield (R)-11 as an orange oil (3.12 g, >99%, 94.5% ee). Spectroscopic data are in good accordance with the literature [14].

3.3. (R)-4-Chloro-2-Acetoxy-1-Ferrocenylbutane ((R)-12)

(R)-4-Chloro-1-ferrocenylbutanol ((R)-11) (3.60 g, 12.3 mmol), 4-dimethylaminopyridine (0.075 mg, 0.62 mmol) and triethylamine (75 mL) were added to a dry 300-mL RBF containing a magnetic stir bar under an inert atmosphere. Acetic anhydride (1.76 mL, 18.6 mmol) was added to the reaction flask which was stirred at room temperature for 18 h. Et2O (100 mL) was added to reaction mixture which was subsequently washed with H2O (100 mL), 10% aqueous NH4Cl (2 × 40 mL), H2O (50 mL) and dried with anhydrous Na2SO4. The reaction mixture was filtered, and the solvent was removed in vacuo. The resultant crude orange oil was used directly in the next step without further purification.

3.4. (R)-N-Allyl-Pyrrolidin-2′-ylferrocene ((R)-13)

(R)-4-Chloro-2-acetoxy-1-ferrocenylbutane (4.20 g, 12.5 mmol), allylamine (6.0 mL, 80.0 mmol) and dry MeOH (6.0 mL) were added to a dry sealed microwave vial containing a magnetic stir bar under an inert atmosphere. The reaction mixture was heated to reflux, stirred for 4 h then diluted with Et2O (10 mL), washed with sat. aqueous NaHCO3 (2 × 10 mL), brine (10 mL) and dried with anhydrous Na2SO4. The solvent was removed in vacuo and the crude product was purified by alumina column chromatography (pentane/EtOAc, 30:1 with 1% triethylamine) to yield (R)-13 as an orange oil (3.61 g, 97%, 92.2% ee). Spectroscopic data are in good accordance with the literature [14].

3.5. 2-[(2R)-N-Allyl-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (14) and 2-[(2R)-N-Allyl-Pyrrolidin-2′-yl]-(1R)-Diphenylphosphineferrocene (14)

(R)-N-allyl-pyrrolidin-2′-ylferrocene ((R)-13) (4.20 g, 12.5 mmol) and dry Et2O (6.0 mL) were added to a dry 250-mL RBF containing a magnetic stir bar under an inert atmosphere. The reaction mixture was cooled to −78 °C and s-BuLi (6.0 mL, 80.0 mmol) was added dropwise. After stirring for 3 h, the reaction mixture was warmed to 0 °C and stirred for an additional 1 h. Ph2PCl (6.0 mL, 80.0 mmol) was added, and the reaction mixture was stirred for 1.5 h and then quenched with aqueous NH4Cl (10%, 15 mL). The aqueous layer was separated and washed with CH2Cl2 (2 × 50 mL) brine (10 mL) and dried with anhydrous Na2SO4. The solvent was removed in vacuo and the crude product was purified by purified by silica gel column chromatography (pentane/EtOAc, 10:1 to 2:1) then alumina column chromatography (pentane/EtOAc with 0.1% triethylamine, 40:1 to 15:1) to yield (R,Sp)-14 as an orange solid (0.79 g, 24%, 93% d.e.) and (R,Rp)-14 as an orange solid (1.08 g, 33%, >99% d.e.). (R,Sp)-14 (0.56 g) was dissolved in 4.5 mL pentane then cooled to −20 °C for 30 min. The precipitate was filtered, washed with pentane, dried and collected to yield (R,Sp)-14 as an orange solid (0.43 g, >99% d.e.).

3.5.1. Spectroscopic Analysis of (R,Sp)-14

Rf = 0.22 (pentane/EtOAc 10:1); m.p. = 136–138 °C; [α]D20 = −232.8 (c 0.63, CH2Cl2); IR (neat): νmax = 3054, 994 (C=C-H), 2939, 2922, 1443 (sp3C-H), 1628 (Alkene: C=C), 1609, 1587, 1565 (Aromatic: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.66–7.57 (m, 2H), 7.41–7.35 (m, 3H), 7.35–7.18 (m, 5H), 5.67–5.48 (m, 1H), 4.82 (d, J = 10.1 Hz, 1H), 4.74 (d, J = 17.1 Hz, 1H), 4.57 (s, 1H), 4.37 (t, J = 2.3 Hz, 1H), 3.96 (s, 5H), 3.46 (td, J = 8.0, 3.2 Hz, 1H), 3.12–2.99 (m, 1H), 2.91 (dd, J = 13.3, 5.2 Hz, 1H), 2.50–2.34 (m, 1H), 2.27–2.14 (m, 1H), 2.13–1.96 (m, 2H), 1.93–1.68 (m, 2H) ppm; 13C NMR (126 MHz, CDCl3): δ 139.7 (d, J = 8.6 Hz), 137.9 (d, J = 8.5 Hz), 135.4 (d, J = 21.9 Hz), 132.9 (d, J = 18.9 Hz), 129.2, 128.2, 128.1 (d, J = 1.6 Hz), 128.1 (d, J = 1.6 Hz), 116.2, 75.4 (d, J = 8.7 Hz), 70.8 (d, J = 4.5 Hz), 70.8, 69.9, 69.7 (d, J = 4.3 Hz), 69.6, 62.1 (d, J = 9.5 Hz), 57.2, 54.0, 35.2, 22.6 ppm; 31P NMR (202 MHz, CDCl3) δ 25.7 ppm; HRMS (ESI-TOF): calcd. for C29H31NPFe [M + H]+ 480.1544; found 480.1536. See Supplementary Materials, pages 21–22 for 1H, 13C and 31P NMR spectra.

3.5.2. Spectroscopic Analysis of (R,Rp)-14

Rf = 0.34 (pentane/EtOAc 9:1); m.p. = 110–112 °C; [α]D20 = 92.7 (c 0.4, CH2Cl2); IR (neat): νmax = 3048, 979 (C=C-H), 1611 (Alkene: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.66–7.48 (m, 2H), 7.45–7.31 (m, 3H), 7.29–7.11 (m, 5H), 5.69–5.49 (m, 1H), 5.12 (d, J = 16.8 Hz, 1H), 4.99 (d, J = 9.9 Hz, 1H), 4.44 (s, 1H), 4.28 (s, 1H), 3.97 (s, 4H), 3.87 (s, 2H), 3.40 (t, J = 7.8 Hz, 1H), 3.04 (t, J = 7.1 Hz, 1H), 2.66 (dd, J = 12.7, 8.2 Hz, 1H), 2.17–1.47 (m, 6H) ppm; 13C NMR (126 MHz, CDCl3): δ 140.6 (d, J = 9.1 Hz), 138.5 (d, J = 9.2 Hz), 137.1, 135.4 (d, J = 21.7 Hz), 132.6 (d, J = 18.3 Hz), 129.1, 128.07 (d, J = 7.8 Hz), 127.9 (d, J = 6.2 Hz), 127.7, 115.9, 97.6 (d, J = 22.5 Hz), 73.3 (d, J = 11.8 Hz), 71.7 (d, J = 4.7 Hz), 71.0 (d, J = 5.1 Hz), 69.7, 68.6, 63.5 (d, J = 3.9 Hz), 58.2, 54.6, 35.3 (d, J = 7.6 Hz), 22.8 (d, J = 1.2 Hz) ppm; 31P NMR (202 MHz, CDCl3) δ 22.98 ppm; HRMS (ESI-TOF): calcd. for C29H31NPFe [M + H]+ 480.1544; found 480.1563. See Supplementary Materials, pages 23–24 for 1H, 13C and 31P NMR spectra.

3.6. 2-[(2R)-N-H-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene ((R,Sp)-15)

2-[(2R)-N-allyl-pyrrolidin-2′-yl]-(1S)-diphenylphosphineferrocene (R,Sp)-14) (0.41 g, 0.85 mmol), Pd(PPh3)4 (18.8 mg, 80.0 mmol), 1,3-dimethylbarbituric acid (NDMBA) (0.42 g, 2.69 mmol) and dry CH2Cl2 (9.0 mL) were added to a dry 50-mL Schlenk flask containing a magnetic stir bar under an inert atmosphere. The reaction mixture was heated to 35 °C, stirred for 5 h, and then quenched with sat. aqueous NaHCO3 (10 mL). The organic layer was separated and washed with sat. aqueous NaHCO3 (10 mL) and dried with anhydrous Na2SO4. The solvent was removed in vacuo and the crude product was purified by alumina column chromatography (pentane/EtOAc/MeOH/ triethylamine, 3:1:0.1:0.01) to yield (R,Sp)-15) as a yellow solid (0.37 g, 98%).

Spectroscopic Analysis of (R,Sp)-15)

Rf = 0.37 (pentane/EtOAc/MeOH, 3:1:0.1); m.p. = 131–132 °C; [α]D20 = −184.4 (c 0.8, CH2Cl2); IR (neat): νmax = 3049, 990 (C=C-H), 2989, 1432 (sp3C-H), 1662 (Alkene: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.59–7.46 (m, 2H), 7.42–7.33 (m, 3H), 7.31–7.15 (m, 5H), 4.46 (s, 1H), 4.34–4.20 (m, 2H), 4.04 (s, 5H), 3.74–3.66 (m, 1H), 2.93 (dd, J = 14.1, 7.7 Hz, 1H), 2.76 (dd, J = 15.3, 8.2 Hz, 1H), 2.27–2.09 (m, 1H), 2.03–1.70 (m, 3H) ppm; 13C NMR (126 MHz, CDCl3): δ 140.2 (d, J = 10.0 Hz), 137.4 (d, J = 8.9 Hz), 135.4 (d, J = 20.9 Hz), 132.7 (d, J = 18.4 Hz), 129.2, 128.4 (d, J = 6.1 Hz), 128.3, 128.2, 96.04 (d, J = 22.8 Hz), 75.8 (d, J = 6.5 Hz), 71.4 (d, J = 4.0 Hz), 69.6, 69.5, 69.1 (d, J = 3.9 Hz), 56.4, 56.3, 47.0, 31.6, 25.3 ppm; 31P NMR (202 MHz, CDCl3) δ 23.6 ppm; HRMS (ESI-TOF): calcd. for C26H27NPFe [M + H]+ 440.1231; found 440.1243. See Supplementary Materials, pages 25−26 for 1H, 13C and 31P NMR spectra.

3.7. 2-[(2R)-N-H-Pyrrolidin-2′-yl]-(1R)-Diphenylphosphineferrocene)-(R,Rp)-15)

Prepared according to the same procedure as for (R,Sp)-15 using (R,Rp)-14 (1.80 g, 3.75 mmol), to afford the product as a yellow solid (1.42 g, 86%).

3.7.1. Spectroscopic Analysis of (R,Rp)-15

Rf = 0.37 (pentane/EtOAc/MeOH, 3:1:0.1); m.p. = 139–140 °C; [α]D20 = 239.0 (c 0.75, CH2Cl2); IR (neat): νmax = 3054, 997 (C=C-H), 2987, 1444 (sp3C-H), 1636, 1590 (Aromatic: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.61–7.44 (m, 2H), 7.41–7.33 (m, 3H), 7.30–7.13 (m, 5H), 4.50 (s, 1H), 4.30–4.16 (m, 2H), 4.07 (s, 5H), 3.77–3.66 (m, 1H), 3.13 (dd, J = 13.3, 8.4 Hz, 1H), 2.93 (dd, J = 16.4, 7.7 Hz, 1H), 2.18 (s, 1H), 1.84–1.45 (m, 3H), 1.21–1.02 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 140.3 (d, J = 9.9 Hz), 137.4 (d, J = 9.0 Hz), 135.1 (d, J = 20.9 Hz), 132.7 (d, J = 18.5 Hz), 129.1, 128.2, 128.2, 128.1 (d, J = 2.9 Hz), 98.4 (d, J = 22.5 Hz), 74.8 (d, J = 7.6 Hz), 71.1 (d, J = 4.4 Hz), 69.5, 69.0, 67.8 (d, J = 4.1 Hz), 56.7 (d, J = 11.3 Hz), 46.8, 34.9, 25.9 ppm; 31P NMR (202 MHz, CDCl3) δ 23.9 ppm; HRMS (ESI-TOF): calcd. for C26H27NPFe [M + H]+ 440.1231; found 440.1209. See Supplementary Materials, pages 27–28 for 1H, 13C and 31P NMR spectra.

3.7.2. Typical Procedure A: Phosphine-Coupling

(R,Sp)-15 or (R,Rp)-15 (1.0 equiv.), Et3N (3.0 equiv.) and dry toluene (0.23 M) were added to a dry 25-mL Schlenk flask containing a magnetic stir bar under an inert atmosphere. The di-substituted chlorophosphine (1.0 equiv.) in dry toluene (0.23 M) was added, and the reaction mixture was stirred at room temperature for 18 h. Heptane (5 mL) was added, and the reaction mixture was filtered. The solvent was removed in vacuo and the crude product was purified by alumina column chromatography (pentane/EtOAc, 99:1 to pentane/EtOAc/MeOH, 3:1:0.1) to yield the product.

3.8. 2-[(2R)-N-Diphenylphosphine-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L1)

Prepared according to typical procedure A using chlorodiphenylphosphine 0.075 g, 0.340 mmol) to afford the product as an orange solid (0.165 g, 78%).

Spectroscopic Analysis of L1

Rf = 0.80 (pentane/EtOAc, 20:1); m.p. = 58–60 °C; [α]D20 = −146.0 (c 0.7, CH2Cl2); IR (neat): νmax = 3032 (C=C-H), 2969, 1422 (sp3C-H) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.67–7.58 (m, 2H), 7.40–7.34 (m, 3H), 7.29–7.13 (m, 10H), 7.11–6.99 (m, 5H), 4.78–4.66 (m, J = 7.1, 3.4 Hz, 1H), 4.49 (d, J = 1.1 Hz, 1H), 4.36 (t, J = 2.4 Hz, 1H), 3.92 (s, 5H), 3.09 (dtd, J = 10.0, 7.3, 2.6 Hz, 1H), 2.78–2.63 (m, 1H), 2.45–2.16 (m, 2H), 1.90–1.70 (m, 1H), 1.55–1.34 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 140.0 (d, J = 7.8 Hz, 2C), 139.7 (d, J = 22.8 Hz), 139.1 (d, J = 9.1 Hz), 135.6 (d, J = 22.2 Hz, 2C), 133.0 (d, J = 21.0 Hz, 2C), 132.8 (d, J = 17.7 Hz), 132.8 (d, J = 17.7 Hz), 131.4 (d, J = 18.5 Hz, 2C), 129.1, 128.4, 128.1 (d, J = 8.1 Hz, 2C), 127.9 (d, J = 6.1 Hz, 2C), 127.8 (d, J = 5.3 Hz, 2C), 127.7 (d, J = 6.0 Hz, 2C), 127.4, 127.4, 99.3 (dd, J = 24.7, 5.4 Hz), 75.8 (d, J = 10.0 Hz), 71.4 (d, J = 4.4 Hz), 69.8 (dd, J = 4.4, 2.5 Hz), 69.7, 69.6 (5C), 60.9 (dd, J = 31.1, 10.0 Hz), 47.7 (d, J = 10.0 Hz), 36.6 (d, J = 6.8 Hz), 26.0 ppm; 31P NMR (202 MHz, CDCl3) δ 44.8 (d, J = 16.4 Hz), -24.3 (d, J = 16.4 Hz) ppm; HRMS (ESI-TOF): calcd. for C38H36NP2Fe [M + H]+ 624.1672; found 624.1650. See Supplementary Materials, pages 29–31 for 1H, 13C and 31P and 31P-31P COSY NMR spectra.

3.9. 2-[(2R)-N-Bis(2-Methylphenyl)phosphine-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L2)

Prepared according to typical procedure A using bis(2-methylphenyl)chlorophosphine (0.057 g, 0.230 mmol) to afford the product as an orange solid (0.063 g, 42%).

Spectroscopic Analysis of L2

Rf = 0.79 (pentane/EtOAc 10:1); m.p. = 84–87 °C; [α]D20 = −220.7 (c 0.18, CH2Cl2); IR (neat): νmax = 3047, 923 (C=C-H), 2921, 2858, 1444 (sp3C-H), 1586, 1563, 1502 (Aromatic: C=C) cm−1; 1H NMR (400 MHz, CDCl3): δ 7.65–7.54 (m, 2H), 7.40–7.28 (m, 4H), 7.23–6.89 (m, 11H), 6.74 (dd, J = 7.3, 2.6 Hz, 1H), 4.95–4.84 (m, 1H), 4.47 (d, J = 1.1 Hz, 1H), 4.31 (t, J = 2.4 Hz, 1H), 4.03–3.95 (m, 1H), 3.86 (s, 5H), 3.12 (dd, J = 16.7, 6.8 Hz, 1H), 2.68–2.31 (m, 3H), 2.25 (s, 3H), 1.99–1.83 (m, 1H), 1.69 (s, 3H), 1.65–1.45 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 142.0 (d, J = 27.9 Hz), 140.5 (dd, J = 7.3, 1.3 Hz), 140.2 (d, J = 25.5 Hz), 139.8 (dd, J = 9.7, 2.1 Hz), 139.0 (d, J = 11.5 Hz), 136.9 (d, J = 18.2 Hz), 135.6 (d, J = 22.6 Hz, 2C), 132.4 (d, J = 17.2 Hz), 132.4 (d, J = 17.2 Hz), 131.4 (d, J = 1.1 Hz), 131.0 (d, J = 3.2 Hz), 129.8 (d, J = 24.6 Hz), 129.8 (d, J = 24.2 Hz), 128.9, 128.0 (d, J = 27.5 Hz, 2C), 127.8, 127.3 (d, J = 12.5 Hz, 2C), 127.3, 126.8, 125.1 (d, J = 15.2 Hz, 2C), 100.1 (dd, J = 25.5, 2.7 Hz), 75.8 (d, J = 10.7 Hz), 71.2 (d, J = 4.6 Hz), 69.8, 69.4 (5C), 69.0 (d, J = 4.3 Hz), 60.3 (dd, J = 32.9, 10.4 Hz), 48.7 (d, J = 9.9 Hz), 36.5 (dd, J = 6.5, 1.9 Hz), 26.9, 21.3 (dd, J = 20.8, 1.4 Hz), 20.7 (d, J = 18.8 Hz) ppm; 31P NMR (202 MHz, CDCl3) δ 31.6 (d, J = 26.0 Hz), −24.8 (d, J = 26.0 Hz) ppm; HRMS (ESI-TOF): calcd. for C40H40NP2Fe [M + H]+ 652.1985; found 652.1993. See Supplementary Materials, pages 32–33 for 1H, 13C and 31P NMR spectra.

3.10. 2-[(2R)-N-Bis(4-fluorophenyl)phosphine-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L3)

Prepared according to typical procedure A using bis(4-fluorophenyl)chlorophosphine (0.054 g, 0.210 mmol) to afford the product as a yellow solid (0.072 g, 50%).

Spectroscopic Analysis of L3

Rf = 0.655 (pentane/EtOAc, 20:1); m.p. = 72–74 °C; [α]D20 = –166.0 (c 0.2, CHCl3); IR (neat): νmax = 3022, 982 (C=C-H), 2936, 2911, 1486 (sp3C-H), 1647 (Alkene: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.67–7.57 (m, 2H), 7.41–7.34 (m, 3H), 7.26–7.18 (m, 2H), 7.10–6.82 (m, 11H), 4.83–4.68 (m, 1H), 4.43 (s, 1H), 4.38 (t, J = 2.4 Hz, 1H), 4.03 (d, J = 1.0 Hz, 1H), 3.91 (s, 5H), 3.10–2.96 (m, 1H), 2.76–2.64 (m, 1H), 2.41–2.15 (m, 2H), 1.88–1.72 (m, 1H), 1.55–1.37 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 164.1, 163.5, 162.1, 161.5, 139.9–139.7 (m), 138.7 (dd, J = 8.8, 1.3 Hz), 135.4 (d, J = 22.2 Hz, 2C), 134.6 (d, J = 22.7 Hz), 134.6 (d, J = 22.5 Hz), 133.0 (d, J = 19.8 Hz), 132.9 (d, J = 19.8 Hz), 132.7 (d, J = 17.8 Hz), 132.7 (d, J = 17.7 Hz), 129.0, 128.0 (d, J = 8.2 Hz, 2C), 127.5 (d, J = 6.1 Hz, 2C), 127.3, 115.0 (d, J = 6.7 Hz), 114.9 (d, J = 5.9 Hz), 114.8 (d, J = 6.7 Hz), 114.7 (d, J = 5.8 Hz), 98.5 (dd, J = 25.2, 6.3 Hz), 75.8 (d, J = 10.8 Hz), 71.38 (d, J = 4.6 Hz), 69.67, 69.47 (5C), 69.34 (dd, J = 4.4, 2.2 Hz), 60.85 (dd, J = 31.6, 10.6 Hz), 47.2 (d, J = 10.0 Hz), 35.9 (d, J = 3.9 Hz), 25.7 ppm; 19F NMR (470 MHz, CDCl3) δ −113.1–−113.2 (m), −114.7–−114.8 (m) ppm; 31P NMR (202 MHz, CDCl3) δ 43.1 (dt, J = 19.5, 5.4 Hz), −25.1 (d, J = 19.5 Hz) ppm; HRMS (ESI-TOF): calcd. for C38H33NP2 F2Fe [M + H]+ 660.1484; found 660.1481. See Supplementary Materials, pages 34–36 for 1H, 13C, 31P and 19F NMR spectra.

3.11. 2-[(2R)-N-Bis(3,5-Di-Trifluoromethylphenyl)phosphine-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L4)

Prepared according to typical procedure A using bis(3,5-di-trifluoromethylphenyl)chlorophosphine (0.103 g, 0.210 mmol) to afford the product as an orange solid (0.075 g, 40%).

Spectroscopic Analysis of L4

Rf = 0.37 (pentane/EtOAc/MeOH, 3:1:0.1); m.p. = 56–58 °C; [α]D20 = −184.3 (c 0.08, CH2Cl2); IR (neat): νmax = 3053, 987 (C=C-H), 2970, 1434 (sp3C-H), 1587, 1576 (Aromatic: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.76 (s, 2H), 7.68–7.57 (m, 2H), 7.44–7.33 (m, 7H), 7.27–7.17 (m, 2H), 6.93–6.83 (m, 3H), 5.07–4.94 (m, 1H), 4.51 (t, J = 2.4 Hz, 1H), 4.43 (s, 1H), 4.21 (t, J = 2.9 Hz, 1H), 3.91 (s, 5H), 2.89 (dd, J = 16.4, 8.3 Hz, 1H), 2.74–2.60 (m, 1H), 2.57–2.41 (m, 1H), 2.34–2.15 (m, 1H), 2.02–1.85 (m, 1H), 1.82–1.65 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 142.1 (d, J = 1.0 Hz), 141.9 (d, J = 1.3 Hz), 141.6, 141.4, 138.9 (dd, J = 6.1, 1.8 Hz), 138.3 (dd, J = 7.8, 1.7 Hz), 135.3 (d, J = 22.4Hz, 2C), 132.9 (d, J = 18.6 Hz), 132.8 (d, J = 18.6 Hz), 131.8–131.7 (m), 131.7–131.5 (m), 131.4 (d, J = 4.7 Hz), 131.3–131.1 (m, 2C), 131.0–130.9 (m, J = 23.5 Hz), 129.1, 128.1 (d, J = 8.4 Hz, 2C), 127.6, 127.3 (d, J = 6.7 Hz, 2C), 122.6–122.4 (m), 122.2–121.9 (m), 96.6 (dd, J = 26.5, 5.9 Hz), 76.5 (dd, J = 10.0, 1.4 Hz), 72.0 (d, J = 4.7 Hz), 70.6, 69.7 (5C), 68.4 (d, J = 3.5 Hz), 60.9 (dd, J = 32.6, 12.7 Hz), 47.3 (d, J = 10.8 Hz), 35.3 (d, J = 5.5 Hz), 26.0 ppm; 19F NMR (470 MHz, CDCl3) δ -62.3, −61.0 ppm; 31P NMR (202 MHz, CDCl3) δ 40.1 (d, J = 26.2 Hz), −27.1 (d, J = 26.2 Hz) ppm; HRMS (ESI-TOF): calcd. for C42H32NP2 F12Fe [M + H]+ 896.1168; found 896.1155. See Supplementary Materials, pages 37–39 for 1H, 13C, 31P and 19F NMR spectra.

3.12. 2-[(2R)-N-(R)-1,1′-Binaphthyl-2,2′-Diylphosphoro-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L5)

Prepared according to typical procedure A using (R)-1,1′-binaphthyl-2,2′-diyl phosphorochloridate (0.21 g, 0.60 mmol, 1.3 equiv.) to afford the product as a yellow solid (0.29 g, 83%).

Spectroscopic Analysis of L5

Rf = 0.47 (pentane/EtOAc, 9:1); m.p. = 219–220 °C; [α]D20 = −241.2 (c 0.29, CH2Cl2); IR (neat): νmax = 3054, 896 (C=C-H), 2987, 1421 (sp3C-H) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.93–7.78 (m, 4H), 7.67–7.57 (m, 2H), 7.47–7.27 (m, 10H), 7.25–7.13 (m, 5H), 6.92 (d, J = 8.8 Hz, 1H), 5.24–5.11 (m, 1H), 4.45 (s, 1H), 4.35 (d, J = 2.3 Hz, 1H), 3.96 (s, 5H), 3.90 (s, 1H), 2.86–2.70 (m, 1H), 2.66–2.54 (m, 1H), 2.51–2.27 (m, 2H), 1.87–1.60 (m, 2H) ppm; 13C NMR (126 MHz, CDCl3): δ 150.4, 149.9, 140.4 (d, J = 8.9 Hz), 138.5 (d, J = 6.4 Hz), 135.6 (d, J = 22.1 Hz, 2C), 132.7, 132.6 (d, J = 1.6 Hz), 132.5 (d, J = 1.9 Hz), 132.4 (d, J = 2.0 Hz), 131.20, 130.4, 129.8, 129.4, 129.1, 128.4, 128.2, 128.1, 128.1, 128.0, 128.0, 127.9, 127.4, 127.3, 127.0 (2C), 125.8 (d, J = 10.9 Hz, 2C), 124.5, 124.3, 124.3, 123.91–123.81 (m), 122.3 (d, J = 1.8 Hz), 122.2 (d, J = 1.6 Hz), 121.9, 97.1 (dd, J = 24.4, 7.4 Hz), 75.5 (d, J = 11.2 Hz), 71.5 (d, J = 4.4 Hz), 69.7–69.6 (m, 6C), 69.5, 56.1 (dd, J = 33.6, 8.8 Hz), 44.0 (d, J = 6.8 Hz), 34.0, 25.34 ppm; 31P NMR (202 MHz, CDCl3) δ 146.9 (d, J = 56.2 Hz), −23.8 (d, J = 56.2 Hz) ppm; HRMS (ESI-TOF): calcd. for C46H38NO2P2Fe [M + H]+ 754.1727; found 754.1719. See Supplementary Materials, pages 40–41 for 1H, 13C, and 31P NMR spectra and page 40 for X-ray crystallographic data.

3.13. 2-[(2R)-N-(S)-1,1′-Binaphthyl-2,2′-Diylphosphoro-Pyrrolidin-2′-yl]-(1S)-Diphenylphosphineferrocene (L6)

Prepared according to typical procedure A using (S)-1,1′-binaphthyl-2,2′-diyl phosphorochloridate (0.10 g, 0.29 mmol, 1.2 equiv.) to afford the product as a yellow solid (0.12 g, 70%).

Spectroscopic Analysis of L6

Rf = 0.83 (pentane/EtOAc, 4:1); m.p. = 208–209 °C; [α]D20 = −32.5 (c 0.17, CH2Cl2); IR (neat): νmax = 2939, 2926, 1443 (sp3C-H), 1613, 1588, 1503 (Aromatic: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 7.92–7.81 (m, 4H), 7.65–7.55 (m, 2H), 7.43–7.33 (m, 6H), 7.30–7.19 (m, 9H), 7.15 (d, J = 8.7 Hz, 1H), 5.07–4.93 (m, 1H), 4.60 (s, 1H), 4.42 (t, J = 2.4 Hz, 1H), 4.01–3.97 (m, 1H), 3.95 (s, 5H), 3.03–2.91 (m, 1H), 2.56–2.42 (m, 1H), 2.39–2.23 (m, 2H), 1.82–1.67 (m, 1H), 1.64–1.49 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 150.1 (d, J = 6.2 Hz), 149.8 (d, J = 1.8 Hz), 139.6 (d, J = 8.2 Hz), 138.5 (d, J = 9.1 Hz), 135.4 (d, J = 22.1 Hz, 2C), 132.8 (d, J = 4.4 Hz), 132.8, 132.6 (d, J = 2.8 Hz), 132.6 (d, J = 1.0 Hz), 131.1, 130.9, 130.6, 129.8, 129.4, 129.1, 128.8, 128.2, 128.1, 128.1, 128.0, 127.9, 127.9, 127.6, 126.9 (d, J = 9.1 Hz, 2C), 125.8 (d, J = 8.2 Hz, 2C), 124.4 (d, J = 7.6 Hz, 2C), 123.8, 122.9, 122.2, 122.2 (d, J = 1.5 Hz), 98.0 (dd, J = 24.2, 7.7 Hz), 75.6 (d, J = 11.2 Hz), 71.6 (d, J = 4.4 Hz), 69.7 (dd, J = 4.0, 2.2 Hz), 69.6 (5C), 68.2, 57.3 (dd, J = 35.3, 9.7 Hz), 44.1 (d, J = 4.8 Hz), 34.7, 26.0 ppm; 31P NMR (202 MHz, CDCl3) δ 150.8 (d, J = 29.3 Hz), -24.7 (d, J = 29.3 Hz) ppm; HRMS (ESI-TOF): calcd. for C46H38NO2P2Fe [M + H]+ 754.1727; found 754.1757. See Supplementary Materials, pages 42–43 for 1H, 13C, and 31P NMR spectra.

3.14. 2-[(2R)-N-(R)-1,1′-Binaphthyl-2,2′-Diylphosphoro-Pyrrolidin-2′-yl]-(1R)-Diphenylphosphineferrocene (L7)

Prepared according to typical procedure A using (R)-1,1′-binaphthyl-2,2′-diyl phosphorochloridate (0.15 g, 0.44 mmol, 1.3 equiv.) to afford the product as an orange solid (0.23 g, 90%).

Spectroscopic Analysis of L7

Rf = 0.86 (pentane/EtOAc 9:1); m.p. = 177–179 °C; [a]D20 = −83.5 (c 0.18, CH2Cl2); IR (neat): νmax = 3052, 947 (C=C-H), 2967, 1434 (sp3C-H), 1588, 1567 (Aromatic: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 8.04–7.83 (m, 5H), 7.68–7.55 (m, 3H), 7.47–7.35 (m, 8H), 7.33 (s, 1H), 7.31–7.26 (m, 3H), 7.25–7.19 (m, 2H), 5.32 (dd, J = 11.7, 7.4 Hz, 1H), 4.51 (d, J = 1.4 Hz, 1H), 4.31 (dd, J = 4.1, 1.7 Hz, 1H), 4.15 (s, 5H), 3.98–3.92 (m, 1H), 3.07–2.90 (m, 1H), 2.79–2.60 (m, 1H), 2.00–1.81 (m, 1H), 1.56–1.32 (m, 3H) ppm; 13C NMR (126 MHz, CDCl3): δ 151.0 (d, J = 4.7 Hz), 150.0, 140.3 (d, J = 9.1 Hz), 137.6 (d, J = 8.0 Hz), 135.3 (d, J = 21.6 Hz, 2C), 132.8 (d, J = 15.6 Hz), 132.5 (d, J = 18.1 Hz, 2C), 131.4, 130.5, 130.3, 129.7, 129.2, 128.3, 128.2, 128.1, 128.1, 128.0, 128.0, 127.9, 127.0 (d, J = 10.7 Hz, 2C), 126.1 (d, J = 4.1 Hz, 2C), 124.8, 124.4, 124.2 (d, J = 5.1 Hz), 122.1 (d, J = 2.0 Hz), 122.0 (2C), 99.9 (d, J = 28.1 Hz), 71.9 (d, J = 10.4 Hz), 71.5 (d, J = 4.1 Hz), 70.2 (d, J = 4.7 Hz), 70.0 (d, J = 5.7 Hz, 5C), 68.5, 57.4 (dd, J = 34.8, 8.6 Hz), 45.3 (d, J = 6.8 Hz), 36.1 (t, J = 3.7 Hz), 23.91 ppm; 31P NMR (202 MHz, CDCl3) δ 148.1, −24.4 ppm; HRMS (ESI-TOF): calcd. for C46H38NO2P2Fe [M + H]+ 754.1727; found 754.1719. See Supplementary Materials, pages 44–45 for 1H, 13C, and 31P NMR spectra.

3.15. 2-[(2R)-N-(S)-1,1′-Binaphthyl-2,2′-Diylphosphoro-Pyrrolidin-2′-yl]-(1R)-Diphenylphosphineferrocene (L8)

Prepared according to typical procedure A using (S)-1,1′-binaphthyl-2,2′-diyl phosphorochloridate (0.15 g, 0.44 mmol, 1.3 equiv.) to afford the product as an orange solid (0.25 g, 98%).

3.15.1. Spectroscopic Analysis of L8

Rf = 0.91 (pentane/EtOAc 9:1); m.p. = 185–187 °C; [a]D20 = 380.2 (c 1.3, CHCl3); IR (neat): νmax = 3067, 977 (C=C-H), 1619 (Alkene: C=C) cm−1; 1H NMR (300 MHz, CDCl3): δ 8.04–7.90 (m, 5H), 7.70–7.52 (m, 4H), 7.50–7.34 (m, 8H), 7.33–7.26 (m, 3H), 7.25–7.22 (m, 2H), 5.33 (dd, J = 9.6, 7.9 Hz, 1H), 4.68 (d, J = 1.3 Hz, 1H), 4.41 (t, J = 2.2 Hz, 1H), 4.18 (s, 5H), 4.01–3.95 (m, 1H), 3.27–3.11 (m, 1H), 2.53–2.40 (m, 1H), 1.79–1.61 (m, 1H), 1.55–1.34 (m, 2H), 1.23–1.12 (m, 1H) ppm; 13C NMR (126 MHz, CDCl3): δ 150.1 (d, J = 6.5 Hz), 149.9 (d, J = 2.1 Hz), 140.3 (d, J = 9.2 Hz), 137.61 (d, J = 8.1 Hz), 135.2 (d, J = 21.5 Hz, 2C), 132.9 (d, J = 1.3 Hz), 132.7 (d, J = 1.1 Hz), 132.5 (d, J = 18.3 Hz, 2C), 131.3, 130.7, 130.3, 129.9, 129.2, 128.4, 128.2, 128.0, 128.1, 128.0, 128.0, 127.9, 127.0 (d, J = 21.6 Hz, 2C), 126.1 (d, J = 4.9 Hz, 2C), 124.6 (d, J = 5.2 Hz, 2C), 123.8 (d, J = 4.9 Hz), 123.3 (d, J = 2.3 Hz), 122.0, 100.9 (dd, J = 24.9, 1.9 Hz), 71.8 (d, J = 10.4 Hz), 71.4 (d, J = 4.2 Hz), 69.9 (d, J = 4.7 Hz, 5C), 69.7 (d, J = 4.9 Hz), 68.4, 59.5 (dd, J = 41.3, 9.3 Hz), 44.0 (d, J = 5.9 Hz), 36.0–35.3 (m), 24.6 ppm; 31P NMR (202 MHz, CDCl3) δ 150.8, −24.7 ppm; HRMS (ESI-TOF): calcd. for C46H38NO2P2Fe [M + H]+ 754.1727; found 754.1765. See Supplementary Materials, pages 46–48 for 1H, 13C and 31P and 31P-31P COSY NMR spectra.

3.15.2. Rhodium-Catalyzed Asymmetric Hydrogenation

Preparation of Substrates/Characterization Data for Substrates and Products

The substrates for catalysis were prepared according to the literature procedures and all characterization data for the substrates and products were in accordance with those reported. 16a/17a16c/7c and 16e/17e and 16g/17g, [27] 16d/16d and 16i/17i and 17k and 17a/19a-18e/19e, [37] 16k is commercially available, 16f/17f, [38] 16h/17h and 16j/17j, [39] and 16f/17f [24].

3.15.3. Rhodium-Catalyzed Asymmetric Hydrogenation of Dehydroamino Acid Esters

Typical Procedure B: Optimization and Substrate Scope

Optimization
The Rh source (0.005 mmol), ligand (0.006 mmol), substrate 16ak (0.5 mmol), and the solvent (2 mL) were added to a dry 10-mL Schlenk flask containing a magnetic stir bar under an inert atmosphere. The reaction mixture was cooled under liquid nitrogen, the atmosphere was evacuated (high vacuum), and then the reaction chamber was refilled with hydrogen using a balloon (reactions requiring higher pressures of hydrogen were quickly transferred to an autoclave). The reaction was stirred for the designated time, filtered through a plug of Celite®, and washed with the solvent of choice. The solvent was removed in vacuo to yield the crude product. The ee was determined by chiral HPLC and conversion of starting material to product by 1H NMR spectroscopy.
Substrate Scope
The reactions were performed with 0.5 mmol of the substrate using the procedure outlined above for the optimization process, with the following exceptions. Racemic reactions were performed with (±)-BINAP (1.1 mol %) using Rh(COD)2OTf (1.0 mol %), 2.3 bar H2 for substrate 16a and 40 bar H2 for 16bk, in THF for 1-18h h at room temperature. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 12 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 4 h at 0 °C. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 12 h at room temperature. For the methods and chiral columns used to determine the enantiomeric excess, and chromatograms for racemic and enantioenriched products, see Supporting Materials (Table S1) and pages 4–14, respectively.

3.16. (S)-Methyl 2-Acetamido-3-Phenylpropanoate (17a)

Prepared according to typical procedure B to afford the product (> 99 % conversion, 99.5% ee with L7) with all characterization analysis in good accordance with the literature.

3.17. (S)-Methyl 2-Acetamido-3-(4-Methoxyphenyl)propanoate (17b)

Prepared according to typical procedure B to afford the product (>99% conversion, >99.9% ee with L7) with all characterization analysis in good accordance with the literature.

3.18. (S)-Methyl 2-Acetamido-3-(p-Tolyl)propanoate (17c)

Prepared according to typical procedure B to afford the product (>99 % conversion, 99.4 % ee with L7) with all characterization analysis in good accordance with the literature.

3.19. (S)-Methyl 2-Acetamido-3-(4-Chlorophenyl)propanoate (17d)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.4% ee with L7) with all characterization analysis in good accordance with the literature.

3.20. (S)-Methyl 2-Acetamido-3-(4-Fluorophenyl)propanoate (17e)

Prepared according to typical procedure B to afford the product (>99% conversion, 98.5% ee with L7) with all characterization analysis in good accordance with the literature.

3.21. (S)-Methyl 2-Acetamido-3-(4-Nitrophenyl)propanoate (17f)

Prepared according to typical procedure B to afford the product (>99% conversion, 98.6% ee with L7) with all characterization analysis in good accordance with the literature.

3.22. (S)-Methyl 2-Acetamido-3-(3-Chlorophenyl)propanoate (17g)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.4% ee with L7) with all characterization analysis in good accordance with the literature.

3.23. (S)-Methyl 2-Acetamido-3-(3-Bromophenyl)propanoate (17h)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.0% ee with L7) with all characterization analysis in good accordance with the literature.

3.24. (S)-Methyl 2-Acetamido-3-(2-Chlorophenyl)propanoate (17i)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.7% ee with L7) with all characterization analysis in good accordance with the literature.

3.25. (S)-Methyl 2-Acetamido-3-(Naphthalen-1-yl)propanoate (17j)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.2% ee with L7) with all characterization analysis in good accordance with the literature.

3.26. (S)-Methyl 2-Acetamidopropanoate (17k)

Prepared according to typical procedure B to afford the product (>99% conversion, 99.1% ee with L7) with all characterization analysis in good accordance with the literature.

3.26.1. Rhodium-Catalyzed Asymmetric Hydrogenation of α-Aryl Enamides

Typical Procedure C: Optimization and Substrate Scope

Optimization
Reactions were set up using a glovebox. Rh(COD)2OTf (0.005 mmol), ligand (0.006 mmol), N-(1-phenylvinyl)acetamide 18a (0.108 g, 0.5 mmol), and the solvent (2 mL) were added to a dry 10-mL Schlenk flask containing a magnetic stir bar under an inert atmosphere. The reaction mixture was cooled under liquid nitrogen, the atmosphere was evacuated (high vacuum), and the reaction chamber was refilled with hydrogen (balloon, reactions requiring higher pressures of hydrogen were quickly transferred to an autoclave). The reaction was stirred for the designated time, filtered through a plug of Celite®, and washed with the solvent of choice. The solvent was removed in vacuo to yield the crude product. The ee was determined by HPLC and conversion of starting material to product by 1H NMR spectroscopy.
Substrate Scope
The reactions were performed with 0.5 mmol of the substrate using the procedure outlined above for the optimization process, with the following exceptions. Racemic reactions were performed with (±)-BINAP (1.2 mol %) using Rh(COD)2OTf (1.0 mol %), 40 bar H2 in THF for 18 h for substrate 18a and Pd/C (1.0 mol %, 10 wt. % loading), 20 bar H2 in methanol for 0.5–2 h for 18bf. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 40 bar H2, in CH2Cl2 for 24 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 20 bar H2, in MeOH for 1 h at room temperature. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 1 h at room temperature except for substrate 18f, which was subjected to Rh(COD)2OTf (1.0 mol %), L7 (1.1 mol %), 60 bar H2, in THF for 2 h at room temperature. For the methods and chiral columns used to determine the enantiomeric excess, and chromatograms for racemic and enantioenriched products, see Supplementary Materials (Table S1) and pages 15–20, respectively.

3.27. (S)-N-(1-Phenylethyl)acetamide (19a)

Prepared according to typical procedure C to afford the product (>99% conversion, 96.4% ee with L7) with all characterization analysis in good accordance with the literature. The absolute configuration of the product was determined by comparison of the [α]D20 value to the literature [40]. Reference value; [α]D20 = 129.5 (c 1.00, CHCl3) for the (R)-enantiomer (99% ee). Value obtained; [α]D20 = −52.7 (c 0.33, CHCl3).

3.28. (S)-N-(1-(Naphthalen-2-yl)ethyl)acetamide (19b)

Prepared according to typical procedure B to afford the product (>99% conversion, 97.4% ee with L7) with all characterization analysis in good accordance with the literature.

3.29. (S)-N-(1-(4-Chlorophenyl)ethyl)acetamide (19c)

Prepared according to typical procedure B to afford the product (>99% conversion, 97.7% ee with L7) with all characterization analysis in good accordance with the literature.

3.30. (S)-N-(1-(4-Methoxyphenyl)ethyl)acetamide (19d)

Prepared according to typical procedure B to afford the product (>99% conversion, 93.7% ee with L7) with all characterization analysis in good accordance with the literature.

3.31. (S)-N-(1,2,3,4-Tetrahydronaphthalen-1-yl)acetamide (19e)

Prepared according to typical procedure B to afford the product (23% conversion, 48.0% ee with L7) with all characterization analysis in good accordance with the literature.

3.32. (S)-Ethyl 3-Acetamido-3-Phenylpropanoate (19f)

Prepared according to typical procedure B to afford the product (>99% conversion, 61.3% ee with L1) with all characterization analysis in good accordance with the literature.

4. Conclusions

In summary, we have reported the design and convenient modular synthesis of a series of novel P,P-ferrocenyl pyrrolidine-containing ligands. Through-space interphosphorus coupling was observed in the 31P-NMR spectra for ligands L1–6, which bear (S)-planar chirality, indicative of a close P–P proximity in the solution phase. The potential application of the ligands was displayed in the rhodium-catalyzed asymmetric hydrogenation of dehydroamino acids and α-aryl enamides with full conversion of the starting materials and excellent ee’s observed in almost all cases using the BINOL-substituted phosphine-phosphoramidite L7. Further investigations of other catalytic asymmetric transformations are currently underway using these ligands, and progress will be reported in due course.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27186078/s1, including 1H, 13C, 19F, and 31P NMR spectra for novel compounds (R,Sp)-14, (R,Rp)-14, (R,Sp)-15, and (R,Rp)-15 and ligands L1L8; Supercritical Fluid Chromatography chromatograms of racemic and enantioenriched compounds 17ak and 19af; and X-ray crystallographic details of L5.

Author Contributions

Conceptualization, P.J.G.; synthetic methodology, X.L.; Investigation, X.L., T.B.B., and Y.O.; writing—original draft preparation, C.K.; writing—review and editing, P.J.G.; supervision, P.J.G.; project administration, P.J.G.; funding acquisition, P.J.G. All authors have read and agreed to the published version of the manuscript.

Funding

X.L. is grateful for the award of an IRCSET Postgraduate Research Scholarship. This publication has emanated from research conducted with the financial support of the Synthesis and Solid State Pharmaceutical Centre (SSPC), funded by Science Foundation Ireland (SFI) under grant numbers 12\RC\2275. C.K. is grateful for the award of an SSPC Ph.D. Scholarship. Acquisition of mass spectra was supported by a Science Foundation Ireland Infrastructure Award (18/RI/5702).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Facilities were provided by the Centre for Synthesis and Chemical Biology (CSCB), funded by the Higher Education Authority’s PRTLI. The authors wish to thank Yannick Ortin of the UCD NMR Centre in the School of Chemistry/CSCB for help with NMR spectroscopic studies and Jimmy Muldoon for the acquisition of mass spectra.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Jacobsen, E.N.; Pfaltz, A.; Yamamoto, H. (Eds.) Comprehensive Asymmetric Catalyses; Springer: Berlin, Heilderberg, Germany, 1999. [Google Scholar]
  2. Ojima, I. (Ed.) Catalytic Asymmetric Synthesis, 3rd ed.; Wiley & Sons: New York, NY, USA, 2010. [Google Scholar]
  3. Lin, G.-Q.; Li, Y.-M.; Chan, A.S.C. Principles and Applications of Asymmetric Synthesis; Wiley & Sons: New York, NY, USA, 2001. [Google Scholar]
  4. Yoon, T.P.; Jacobsen, E.N. Privileged Chiral Catalysts. Science 2003, 299, 1691–1693. [Google Scholar] [CrossRef] [PubMed]
  5. Togni, A.; Breutel, C.; Schnyder, A.; Spindler, F.; Landert, H.; Tijani, A. A Novel Easily Accessible Chiral Ferrocenyldiphosphine for Highly Enantioselective Hydrogenation, Allylic Alkylation, and Hydroboration Reactions. J. Am. Chem. Soc. 1994, 116, 4062–4066. [Google Scholar] [CrossRef]
  6. Tang, W.; Zhang, X. New Chiral Phosphorus Ligands for Enantioselective Hydrogenation. Chem. Rev. 2003, 103, 3029–3070. [Google Scholar] [CrossRef] [PubMed]
  7. Blaser, H. The Chiral Switch of (S)-Metolachlor: A Personal Account of an Industrial Odyssey in Asymmetric Catalysis. Adv. Synth. Catal. 2002, 344, 17–31. [Google Scholar] [CrossRef]
  8. Knochel, P.; Polborn, K.; Lotz, M. New Ferrocenyl Ligands with Broad Applications in Asymmetric Catalysis. Angew. Chem. Int. Ed. 2002, 41, 4708–4711. [Google Scholar] [CrossRef]
  9. Cunningham, L.; Benson, A.; Guiry, P.J. Recent developments in the synthesis and applications of chiral ferrocene ligands and organocatalysts in asymmetric catalysis. Org. Biomol. Chem. 2020, 18, 9329–9370. [Google Scholar] [CrossRef]
  10. Farrell, A.; Goddard, R.; Guiry, P.J. The preparation of ferrocene-containing phosphinamine ligands possessing central and planar chirality and their application in palladium-catalysed asymmetric allylic alkylation. J. Org. Chem. 2002, 67, 4209–4217. [Google Scholar] [CrossRef]
  11. Cahill, J.P.; Guiry, P.J. The Application of Pd-Complexes of trans-2,5-Dialkylpyrrolidinylbenzyldiphenyl phosphines to Enantioselective Allylic Alkylation. Tetrahedron Asymmetry 1998, 9, 4301–4306. [Google Scholar] [CrossRef]
  12. Cahill, J.P.; Bohnen, F.; Goddard, R.; Krüger, C.; Guiry, P.J. The Preparation of trans-2,5-Dialkylpyrrolidinylbenzyldiphenylphosphines: New Phosphinamine Ligands for Asymmetric Catalysis. Tetrahedron Asymmetry 1998, 9, 3831–3839. [Google Scholar] [CrossRef]
  13. Cahill, J.P.; Cunneen, D.; Guiry, P.J. trans-2,5-Dialkylpyrrolidinyl-containing Phosphinamines; Synthetic and Mechanistic Studies in Pd-Catalysed Asymmetric Allylic Alkylation. Tetrahedron Asymmetry 1999, 10, 4157–4173. [Google Scholar] [CrossRef]
  14. Ahern, T.; Müller-Bunz, H.; Guiry, P.J. The Synthesis of N,O-Ferrocenyl Pyrrolidine-Containing Ligands And Their Application in Diethyl- and Diphenylzinc Addition to Aromatic Aldehydes. J. Org. Chem. 2006, 71, 7596–7602. [Google Scholar] [CrossRef] [PubMed]
  15. Meaney, K.; Goddard, R.; Bronger, R.P.J.; Guiry, P.J. The preparation of ferrocene-containing phosphinamine ligands possessing central and planar chirality and their application in palladium-catalysed allylic substitution. Tetrahedron 2021, 90, 132088. [Google Scholar] [CrossRef]
  16. Boaz, N.W.; Debenham, S.D.; MacKenzie, E.B.; Large, S.E. Phosphinoferrocenylaminophosphines as Novel and Practical Ligands for Asymmetric Catalysis. Org. Lett. 2002, 4, 2421–2424. [Google Scholar] [CrossRef] [PubMed]
  17. Boaz, N.W.; MacKenzie, E.B.; Debenham, S.D.; Large, S.E.; Ponasik, J.A. Synthesis and Application of Phosphinoferrocenylaminophosphine Ligands for Asymmetric Catalysis. J. Org. Chem. 2005, 70, 1872–1880. [Google Scholar] [CrossRef]
  18. Krabbe, S.W.; Hatcher, M.A.; Bowman, R.K.; Mitchell, M.B.; McClure, M.S.; Johnson, J.S. Copper-Catalyzed Asymmetric Hydrogenation of Aryl and Heteroaryl Ketones. Org. Lett. 2013, 15, 4560–4563. [Google Scholar] [CrossRef]
  19. Deng, J.; Hu, X.-P.; Huang, J.-D.; Yu, S.-B.; Wang, D.-Y.; Duan, Z.-C.; Zheng, Z. Enantioselective Synthesis of β2-Amino Acids via Rh-Catalyzed Asymmetric Hydrogenation with BoPhoz-Type Ligands:  Important Influence of an N−H Proton in the Ligand on the Enantioselectivity. J. Org. Chem. 2008, 73, 2015–2017. [Google Scholar] [CrossRef]
  20. Duan, Z.-C.; Hu, X.-P.; Zhang, C.; Zheng, Z. Enantioselective Rh-Catalyzed Hydrogenation of 3-Aryl-4-phosphonobutenoates with a P-Stereogenic BoPhoz-Type Ligand. J. Org. Chem. 2010, 75, 8319–8321. [Google Scholar] [CrossRef]
  21. Pan, C.; Gu, Z.; Zhang, M.; Zhu, Z. Palladium-Catalyzed Enantioselective Synthesis of 2-Aryl Cyclohex-2-enone Atropisomers: Platform Molecules for the Divergent Synthesis of Axially Chiral Biaryl Compounds. Angew. Chem. Int. Ed. 2017, 56, 4777–4781. [Google Scholar] [CrossRef]
  22. Hu, X.-P.; Zheng, Z. Unsymmetrical Hybrid Ferrocene-Based Phosphine-Phosphoramidites:  A New Class of Practical Ligands for Rh-Catalyzed Asymmetric Hydrogenation. Org. Lett. 2004, 6, 3585–3588. [Google Scholar] [CrossRef]
  23. Jia, X.; Li, X.; Lam, W.S.; Kok, S.H.L.; Xu, L.; Lu, G.; Yeung, C.-H.; Chan, A.S.C. The synthesis of new chiral phosphine–phosphinites, phosphine–phosphoramidite, and phosphine–phosphite ligands and their applications in asymmetric hydrogenation. Tetrahedron Asymmetry 2004, 15, 2273–2278. [Google Scholar] [CrossRef]
  24. Hu, X.-P.; Zheng, Z. Practical Rh(I)-Catalyzed Asymmetric Hydrogenation of β-(Acylamino)acrylates Using a New Unsymmetrical Hybrid Ferrocenylphosphine−Phosphoramidite Ligand:  Crucial Influence of an N−H Proton in the Ligand. Org. Lett. 2005, 7, 419–422. [Google Scholar] [CrossRef] [PubMed]
  25. Garro-Helion, F.; Merzouk, A.; Guibe, F. Mild and Selective Palladium(0)-Catalyzed Deallylation of Allylic Amines. Allylamine and Diallylamine as Very Convenient Ammonia Equivalents for the Synthesis of Primary Amines. J. Org. Chem. 1993, 58, 6109–6113. [Google Scholar] [CrossRef]
  26. Reetz, M.T.; Gosberg, A.; Goddard, R.; Kyung, S.-H. Diphosphonites as highly efficient ligands for enantioselective rhodium-catalyzed hydrogenation. Chem. Commun. 1998, 2077–2078. [Google Scholar] [CrossRef]
  27. Ouyang, G.; Xiang, J.; Li, Y.; He, Y.; Fan, Q. Cation-Triggered Switchable Asymmetric Catalysis with Chiral Aza-CrownPho. Angew. Chem. Int. Ed. 2015, 54, 4334–4337. [Google Scholar] [CrossRef] [PubMed]
  28. Meetsma, A.; de Vries, J.G.; Giacomina, F.; de Vries, A.H.M.; Lefort, L.; Panella, L. High Enantioselectivity Is Induced by a Single Monodentate Phosphoramidite Ligand in Iridium-Catalyzed Asymmetric Hydrogenation. Angew. Chem. Int. Ed. 2007, 46, 1497–1500. [Google Scholar] [CrossRef]
  29. Doskocz, M.; Malinowska, B.; Młynarz, P.; Lejczak, B.; Kafarski, P. Long range phosphorus–phosphorus coupling constants in bis(phosphorylhydroxymethyl)benzene derivatives. Tetrahedron Lett. 2010, 51, 3406–3411. [Google Scholar] [CrossRef]
  30. Hierso, J.-C.; Fihri, A.; Ivanov, V.V.; Hanquet, B.; Pirio, N.; Donnadieu, B.; Rebière, B.; Amardeil, R.; Meunier, P. “Through-Space” Nuclear Spin−Spin JPP Coupling in Tetraphosphine Ferrocenyl Derivatives:  A 31P NMR and X-ray Structure Correlation Study for Coordination Complexes. J. Am. Chem. Soc. 2004, 126, 11077–11087. [Google Scholar] [CrossRef]
  31. Dai, L.-X.; Tu, T.; You, S.-L.; Deng, W.-P.; Hou, X.-L. Asymmetric Catalysis with Chiral Ferrocene Ligands. Acc. Chem. Res. 2003, 36, 659–667. [Google Scholar] [CrossRef]
  32. Vineyard, B.D.; Knowles, W.S.; Sabacky, M.J. Catalytic Asymmetric Hydrogenation. J. Chem. Soc. Chem. Commun. 1972, 10–11. [Google Scholar] [CrossRef]
  33. Vineyard, B.D.; Knowles, W.S.; Sabacky, M.J.; Bachman, G.L.; Weinkauff, D.J. Asymmetric Hydrogenation. Rhodium Chiral Bisphosphine Catalyst. J. Am. Chem. Soc. 1977, 99, 5946. [Google Scholar] [CrossRef]
  34. Kagan, H.; Dang, T.-P. Asymmetric Catalytic Reduction with Transition Metal Complexes. I. A Catalytic System of Rh (I) with (−)-2,3-O−Isopropylidene-2,3-dihydroxy-1,4-bis(diphenylphosphino)butane, a New Chiral Diphosphine. J. Am. Chem. Soc. 1972, 94, 6429. [Google Scholar] [CrossRef]
  35. Burk, M.J. C2-Symmetric Bis(phospholanes) and Their Use in Highly Enantioselective Hydrogenation Reactions. J. Am. Chem. Soc. 1991, 113, 8518. [Google Scholar] [CrossRef]
  36. Burk, M.J.; Wang, Y.M.; Lee, J.R. A Convenient Asymmetric Synthesis of α-1-Arylalkylamines Through the Enantioselective Hydrogenation of Enamides. J. Am. Chem. Soc. 1996, 118, 5142. [Google Scholar] [CrossRef]
  37. Mohara, B.; Stephan, M. Practical Enantioselective Hydrogenation of α-Aryl- and α-Carboxyamidoethylenes by Rhodium(I)-{1,2-Bis[(o-tert-butoxyphenyl)(phenyl)phosphino]ethane}. Adv. Synth. Catal. 2013, 355, 594–600. [Google Scholar] [CrossRef]
  38. Li, X.; You, C.; Li, S.; Lv, H.; Zhang, Z. Nickel-Catalyzed Enantioselective Hydrogenation of β-(Acylamino) acrylates: Synthesis of Chiral β-Amino Acid Derivatives. J. Org. Chem. 2004, 69, 8157–8160. [Google Scholar] [CrossRef]
  39. Tang, W.; Capacci, A.G.; White, A.; Ma, S.; Rodriguez, S.; Qu, B.; Savoie, J.; Patel, N.D.; Wei, X.; Haddad, N.; et al. Novel and Efficient Chiral Bisphosphorus Ligands for Rhodium-Catalyzed Asymmetric Hydrogenation. Org. Lett. 2010, 12, 1104–1107. [Google Scholar] [CrossRef]
  40. Kohara, T.; Hashimoto, Y.; Saigo, K. Design, synthesis, and optical resolution of a novel non-natural chiral auxiliary, 1-(2,5-dimethoxyphenyl)ethylamine. Application to diastereoselective alkylation of aldimines. Tetrahedron 1999, 55, 6453–6464. [Google Scholar] [CrossRef]
Figure 1. Examples of privileged P,P ligands.
Figure 1. Examples of privileged P,P ligands.
Molecules 27 06078 g001
Figure 2. Selected ferrocene-pyrrolidine containing ligands.
Figure 2. Selected ferrocene-pyrrolidine containing ligands.
Molecules 27 06078 g002
Figure 3. Ferrocenylphosphine-aminophosphine and ferrocenyphosphine-phosphoramidite ligands.
Figure 3. Ferrocenylphosphine-aminophosphine and ferrocenyphosphine-phosphoramidite ligands.
Molecules 27 06078 g003
Scheme 1. Enantioselective synthesis of (R)-13.
Scheme 1. Enantioselective synthesis of (R)-13.
Molecules 27 06078 sch001
Scheme 2. Enantioselective synthesis of L1L8.
Scheme 2. Enantioselective synthesis of L1L8.
Molecules 27 06078 sch002
Figure 4. Series of novel ligands L1L8 prepared.
Figure 4. Series of novel ligands L1L8 prepared.
Molecules 27 06078 g004
Figure 5. Two-dimensional 31P{1H}–31P{1H} spectra (31P COSY) of L1.
Figure 5. Two-dimensional 31P{1H}–31P{1H} spectra (31P COSY) of L1.
Molecules 27 06078 g005
Figure 6. X-ray crystallographic analysis of L5 (15 % oxidation of P1 was observed).
Figure 6. X-ray crystallographic analysis of L5 (15 % oxidation of P1 was observed).
Molecules 27 06078 g006
Scheme 3. Substrate scope of dehydroamino acid esters 16a–k. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 12 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 4 h at 0 °C. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 12 h at room temperature. See the Supporting Information for further details.
Scheme 3. Substrate scope of dehydroamino acid esters 16a–k. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 12 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 1 bar H2, in THF for 4 h at 0 °C. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 12 h at room temperature. See the Supporting Information for further details.
Molecules 27 06078 sch003
Scheme 4. Substrate scope of αaryl enamides 18af. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 40 bar H2, in CH2Cl2 for 24 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 20 bar H2, in MeOH for 1 h at room temperature. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 1 h at room temperature except for substrate 18f, which was subjected to Rh(COD)2OTf (1.0 mol %), L7 (1.1 mol %), 60 bar H2, in THF for 2 h at room temperature. See the Supporting Information for further details.
Scheme 4. Substrate scope of αaryl enamides 18af. Reactions with L1 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 40 bar H2, in CH2Cl2 for 24 h at room temperature. Reactions with L4 (1.1 mol %) were performed using Rh(COD)2OTf (1.0 mol %), 20 bar H2, in MeOH for 1 h at room temperature. Reactions with L7 (0.22 mol %) were performed using Rh(COD)2OTf (0.2 mol %), 10 bar H2, in THF for 1 h at room temperature except for substrate 18f, which was subjected to Rh(COD)2OTf (1.0 mol %), L7 (1.1 mol %), 60 bar H2, in THF for 2 h at room temperature. See the Supporting Information for further details.
Molecules 27 06078 sch004
Table 1. Magnitude of the tsJPP coupling constants in L1L8.
Table 1. Magnitude of the tsJPP coupling constants in L1L8.
LigandtsJPP (Hz)
L116.4
L226.0
L319.5
L426.2
L556.2
L629.3
L7-
L8-
Table 2. Screening of asymmetric hydrogenation conditions with substrate 16a.
Table 2. Screening of asymmetric hydrogenation conditions with substrate 16a.
Molecules 27 06078 i001
EntryLigandRh SourceSolventH2 Pressure (bar)Time (h)Conv. (%) [a]Ee (%) [b]
1(±)-BINAPRh(COD)2OTfTHF2.315>99-
2L1Rh(COD)2OTfTHF112>9989.0
3L1Rh(COD)2BF4THF11278.790.0
4L1Rh(COD)2OTfTHF202>9976.3
5L1Rh(COD)2OTfTHF401>9971.2
6L1Rh(COD)2OTfDCM112>9985.3
7L1Rh(COD)2OTfMeOH112>9985.7
8L1Rh(COD)2OTfDMF11242.580.2
9L1Rh(COD)2OTf1,4-Dioxane1126682.8
10L1Rh(COD)2OTfToluene1125080.3
11L1Rh(COD)2BF4Toluene11238.381.8
12L2Rh(COD)2OTfTHF11270.627.3
13L2Rh(COD)2OTfDCM11064.317.6
14L3Rh(COD)2OTfTHF112>9988.9
15L4Rh(COD)2OTfTHF12>9992.2
16 [c]L4Rh(COD)2OTf THF13>9995.5
17 [c]L4Rh(COD)2BF4 THF149795.4
18L5Rh(COD)2OTfTHF112>9997.9
19L5Rh(COD)2BF4THF112>9997.7
20L5Rh(COD)2OTfTHF102>9995.3
21L5Rh(COD)2OTfDCM112>9997.7
22L5Rh(COD)2OTfMeOH112>9992.3
23L6Rh(COD)2OTfTHF112>9986.7
24L7Rh(COD)2OTfTHF11>9999.5
25 [d]L7Rh(COD)2OTfTHF204>9997.7
26L8Rh(COD)2OTfTHF112>9946.7 (R)
27 [5](R,S)-JosiphosRh(COD)2OTfMeOH10.33>9996
28 [35]DuPhosRh(COD)2OTfMeOH2 (atm)1>9985
29 [16]BoPhozRh(COD)2OTfTHF119699.5
Reactions were performed on a 0.5-mmol 16a scale; see the Supporting Information for further details. [a] Determined by 1H NMR spectroscopy of the crude product. [b] Determined by high performance liquid chromatography using a chiral stationary phase. [c] Reaction performed at 0 °C. [d] 0.02 mol % Rh(COD)2OTf and 0.022 mol % ligand used, reaction performed on 2.5 mmol 16a scale.
Table 3. Screening of asymmetric hydrogenation conditions with substrate 18a.
Table 3. Screening of asymmetric hydrogenation conditions with substrate 18a.
Molecules 27 06078 i002
EntryLigandSolventH2 Pressure (bar)Time (h)Conv. (%) [a]Ee (%) [b]
1(±)-BINAPTHF4018>99-
2L1THF402>9933.6
3L1THF202>9933.4
4L1CH2Cl2202>9929.5
5L1MeOH201>9946.0
6L2MeOH201>9912.5
7L3MeOH201>9945.6
8L4MeOH201>9991.3
9L5THF102>9992.0
10L5THF134289.2
11L7THF12.5>9996.4
12L7THF101>9996.0
13 [c]L7THF101>9991.0
14L8THF125931.7 (R)
15 [36]Me-BPEMeOH4 (atm)12>9995.2
16 [22]BoPhozCH2Cl210199.561.8
Reactions were performed on a 0.5 mmol 18a scale, see the Supporting Information for further details. [a] Determined by 1H NMR spectroscopy of the crude product. [b] Determined by high performance liquid chromatography using a chiral stationary phase. [c] 0.2 mol % Rh(COD)2OTf, 0.21 mol % ligand used.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Brennan, T.B.; Kingston, C.; Ortin, Y.; Guiry, P.J. Design and Synthesis of Pyrrolidinyl Ferrocene-Containing Ligands and Their Application in Highly Enantioselective Rhodium-Catalyzed Olefin Hydrogenation. Molecules 2022, 27, 6078. https://doi.org/10.3390/molecules27186078

AMA Style

Li X, Brennan TB, Kingston C, Ortin Y, Guiry PJ. Design and Synthesis of Pyrrolidinyl Ferrocene-Containing Ligands and Their Application in Highly Enantioselective Rhodium-Catalyzed Olefin Hydrogenation. Molecules. 2022; 27(18):6078. https://doi.org/10.3390/molecules27186078

Chicago/Turabian Style

Li, Xin, Therese B. Brennan, Cian Kingston, Yannick Ortin, and Patrick J. Guiry. 2022. "Design and Synthesis of Pyrrolidinyl Ferrocene-Containing Ligands and Their Application in Highly Enantioselective Rhodium-Catalyzed Olefin Hydrogenation" Molecules 27, no. 18: 6078. https://doi.org/10.3390/molecules27186078

Article Metrics

Back to TopTop