Next Article in Journal
Naphthalene Dehydrogenation on Ni(111) in the Presence of Chemisorbed Oxygen and Nickel Oxide
Next Article in Special Issue
Air-Stable and Highly Active Transition Metal Phosphide Catalysts for Reductive Molecular Transformations
Previous Article in Journal
Chemical Looping Technology for Energy Storage and Carbon Emissions Reductions
Previous Article in Special Issue
Acyclic Diene Metathesis (ADMET) Polymerization for the Synthesis of Chemically Recyclable Bio-Based Aliphatic Polyesters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferrocenes

1
Department of Natural Science, Graduate School of Science and Technology, Tokushima University, Tokushima 770-8506, Japan
2
Tokushima International Science Institute, Tokushima University, Tokushima 770-8501, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(2), 123; https://doi.org/10.3390/catal14020123
Submission received: 10 January 2024 / Revised: 1 February 2024 / Accepted: 2 February 2024 / Published: 4 February 2024

Abstract

:
The molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of a series of racemic planar-chiral 1,1′-diallylferrocene derivatives was reinvestigated utilizing the method of generating catalytically active chiral molybdenum-alkylidene species in situ, which allowed us to examine a variety of chiral molybdenum-alkylidene metathesis precatalysts in the present asymmetric reaction. With the catalyst screening experiments conducted in this study, the more practical reaction conditions, including a choice of a proper chiral molybdenum precatalyst, giving planar-chiral ferrocenes of higher enantiomeric purity and better chemoselectivity could be optimized.

Graphical Abstract

1. Introduction

The introduction of two (or more) different substituents in a single η5-cyclopentadienyl ligand in ferrocene breaks the symmetry of the molecule, and so-called planar chirality is induced in it. Planar-chiral ferrocene derivatives have been recognized as being useful chiral scaffolds in organic and organometallic chemistry (Figure 1) [1,2,3,4,5,6,7,8], and various such complexes have been utilized in a wide range of asymmetric reactions as chiral ligands [9,10,11,12,13,14,15,16,17,18,19] or chiral catalysts [19,20,21,22,23,24,25,26]. In spite of their usefulness, enantioselective preparation of planar-chiral ferrocenes is still a challenging problem. Classical methods of obtaining enantiomerically pure (or enantiomerically enriched) planar-chiral ferrocene derivatives are the enantiomeric resolution of preformed racemates [27], including enzymatic resolution [28,29,30], and diastereoselective metalation utilizing chiral ortho-directing groups [31,32,33,34,35,36,37,38,39,40]. To the best of our knowledge, the first catalytic enantioselective synthesis of planar-chiral ferrocenes was reported in 1997 by Schmaltz and Siegel [41]. In 2006, three research groups (O’Brien [42], Moyano [43], and ourselves [44]) independently reported the catalytic asymmetric reactions giving enantiomerically enriched planar-chiral ferrocene derivatives. Since then, nearly 100 research works related to this topic have been published worldwide [45,46,47,48].
Since 2002, our research group has been interested in utilizing the ring-closing metathesis (RCM) reactions for the modulation of various transition-metal complexes [49,50,51,52,53,54,55] using the Schrock’s Mo- [56,57] and the Grubbs’ Ru-alkylidene [58,59,60,61] complexes. The RCM protocols were extended to the asymmetric counterparts to show the excellent enantioselectivity in the asymmetric synthesis of diverse planar-chiral transition-metal complexes either by the kinetic resolution (KR) of the racemic substrates [44,48,62,63,64] or by the desymmetrization of the Cs-symmetric substrates [48,65,66,67].
The chiral catalysts employed in our asymmetric reactions, which produced enantiomerically enriched planar-chiral transition-metal complexes, were chiral Schrock–Hoveyda molybdenum-alkylidene precatalysts [68,69,70,71] (Figure 2). At the beginning of the development of these chiral precatalysts, each chiral molybdenum complex needed to be prepared and isolated one by one prior to the catalytic applications, which made the screening of the chiral precatalysts/diolate ligands tedious and time-consuming. In 2006, Schrock and Hoveyda reported molybdenum complex A, which served as a universal precursor to generate a variety of chiral molybdenum-alkylidene precatalysts in situ by means of a reaction with an appropriate chiral diol (Scheme 1) [72]. The development of this method enabled rapid and operationally simple screening of the various chiral molybdenum-alkylidene precatalysts in asymmetric olefin metathesis reactions.
Whereas our very first contribution to catalytic asymmetric synthesis of planar-chiral transition-metal complexes, which was the enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral 1,1′-diallylferrocene derivatives 1 (Scheme 2), was reported prior to the development of complex A, only one molybdenum-alkylidene species, Mo*/(R)-L1, was examined as a chiral precatalyst in the original publication [44]. In this article, various chiral Mo-alkylidene species were generated in situ, as shown in Scheme 1, and applied in the enantioselective RCM/KR reaction of racemic 1. After the extensive screening of the chiral Schrock–Hoveyda metathesis precatalysts, the more practical conditions giving planar-chiral ferrocenes of higher enantiomeric purity and better chemoselectivity could be determined. Here, we would like to describe the details of our observations.

2. Results and Discussion

2.1. Design and Preparation of Racemic Planar-Chiral 1,1′-Diallylferrocene Substrates 1a-c for the Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution

Our preliminary studies on the enantioselective RCM/KR of racemic planar-chiral 1,1′-diallylferrocene derivatives 1 [44] postulated that a couple of structural factors in the substrates were crucial to achieve the high enantioselectivity in the molybdenum-catalyzed reactions: (i) steric discrimination of the two allylic groups at the 1- and 1’-positions of the ferrocene core with a methyl substituent in the 2-allylic position of one of the two allylic substituents, and (ii) a bulky substituent R’ in the position adjacent to the unsubstituted (i.e., the more reactive) allyl group (Figure 3).
The preparation of the designed unsymmetric ferrocene derivatives rac-1a-c was achieved by the method developed by Manriquez et al., as shown in Scheme 3 [73]. Allylation of lithium 1,3-R’2-cyclopentadienide with allyl bromide, followed by a deprotonation reaction with butyllithium, provided a THF solution of lithium 1-allyl-2,4-R’2-cyclopentadienide (B). The allylation was highly regioselective, and the formation of a regioisomer, lithium 2-allyl-1,3-R’2-cyclopentadienide, was negligible. A reaction of Fe(acac)2 with stoichiometric B at –78 °C generated metastable intermediate C as a THF solution, and a subsequent reaction with sodium methallylcyclopentadienide (D) afforded rac-1a-c in the yields ranging 20–45%. The reaction conditions were not optimized. The reaction giving rac-1a-c was not completely hetero-selective, and the formation of homoleptic ferrocene byproducts, (η5-methallyl-C5H4)2Fe and (η5-1-allyl-2,4-R’2-C5H2)2Fe, could not be eliminated. The target compounds could be separated from the byproducts by the standard column chromatography on alumina.
The ferrocene substrates for this study, rac-1a-c, possess a trisubstituted cyclopentadienide, η5-(C5H2-1-allyl-2,4-R’2), which is responsible for inducing planar chirality in ferrocene compounds, and the monosubstituted η5-cyclopentadienyl ligand η5-(C5H4-methallyl). In the presence of an appropriate chiral (R)-molybdenum-alkylidene metathesis catalyst, one of the two planar-chiral enantiomers in 1 is preferentially cyclized to give enantiomerically enriched bridged ferrocene (ferrocenophane) (R)-2, and antipodal (S)-1 is left intact. The major side reaction of this RCM/KR process is the formation of 3, which is a product of the metathesis dimerization at the unsubstituted allyl group in the trisubstituted cyclopentadienide (Scheme 2).
For the highly enantioselective kinetic resolution of rac-1a-c, the allyl group in a trisubstituted cyclopentadienyl needs to be more reactive than the methallyl (2-methylallyl) substituent in a monosubstituted cyclopentadienyl. In general, less substituted olefins are more reactive than more substituted ones in olefin metathesis. When a chiral molybdenum-alkylidene species, (R)-Mo*, approaches rac-1, an initial reaction takes place preferentially at the allyl group in the planar-chiral trisubstituted Cp to give intermediates (Ra,Rp)- and (Ra,Sp)-E as a diastereomeric mixture (“Ra” represents the absolute configuration of the axially chiral biaryl moiety in Mo*), and the formation of F is unfavorable. While (Ra,Rp)-E is transformed into (R)-2 smoothly via the RCM reaction, the epimeric intermediate, (Ra,Sp)-E, is forced to take an unfavorable conformation due to the steric repulsion between an R’ group in η5-C5H2R’2(allyl) and an R group in the chiral biaryloxide ligand in (R)-Mo* to liberate (S)-1 intact (Scheme 4). On the other hand, the unfavorable formation of intermediate F is crucial for the high enantioselectivity in the RCM/KR process. Whereas the methallyl group in η5-C5H4-methallyl is remote from the planar-chiral η5-C5H2R’2(allyl) moiety, its reaction with (R)-Mo* takes place with low diastereoselectivity. Following the RCM step in F is an intramolecular process, and, thus, both (Ra,Rp)- and (Ra,Sp)-F shall be transformed into the corresponding ferrocenophanes to give 2 with low enantioselection.

2.2. Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferroce Substrates 1a-c: Catalyst Screening Studies

At the outset, the screening of precatalysts/reaction conditions was examined for the enantioselective RCM/KR reaction of rac-1a. Precatalyst Mo*/(R)-L1 (10 mol % to rac-1a), which was generated in situ as outlined in Scheme 1, showed a nearly identical performance (selectivity and reactivity) to the preformed catalyst [44] in the reaction (Table 1, Entry 1). Although the enantioselectivity of this reaction was quite high, with a krel value of 109 (krel = ([reaction rate of the fast-reacting enantiomer]/[reaction rate of the slow-reacting enantiomer]; selectivity factor) [74,75], the observed drawbacks of using Mo*/(R)-L1 were (i) the competition between the RCM reaction giving the desired ferrocenophane 2a and the bimolecular metathesis giving side-product 3a, (ii) the necessity of the high-dilution conditions for minimizing the formation of metathesis dimer 3a, and (iii) the relatively high reaction temperature (50 °C) to retain the reasonable catalytic activity of the molybdenum species under the diluted conditions. It was found that the reaction using Mo*/(R)-L2 (10 mol % to rac-1a) promoted the intramolecular RCM reaction giving (R)-2 preferentially, and the formation of 3 via the intermolecular reaction was not observed (Entry 2). The enantioselectivity was reasonably high (krel = 65) but slightly lower than that in Entry 1. Precatalyst Mo*/(R)-L2 was catalytically more active than Mo*/(R)-L1 in the reaction of rac-1a, and, thus, the reaction could be conducted at a lower temperature. Enantioselectivity was improved to krel = 95 at 25 °C (Entry 3). Since Mo*/(R)-L2 did not drive the bimolecular reaction giving 3, the RCM/KR reaction catalyzed by Mo*/(R)-L2 could be conducted under the more concentrated conditions, which realized the shorter reaction time whilst retaining a high enantioselectivity. Under these conditions, RCM product (R)-2a of 96% ee was obtained in a 45% yield, and unreacted (S)-1a of 75% ee was recovered in 55%, of which krel was 110 (Entry 4). It is worth mentioning that the “concentrated” reaction could be carried out in an NMR sample tube using C6D6 as a solvent, which allowed for the direct monitoring of the reaction progress through 1H-NMR measurements. The catalytic performance of Mo*/(R)-L3 was similar to that of Mo*/(R)-L1 (Entry 5). Precatalyst Mo*/(R)-L4 was the most reactive among the precatalysts examined but far less enantioselective. The reaction of rac-1a catalyzed by Mo*/(R)-L4 at 50 °C for 24 h was leading to the complete consumption of the substrate to provide nearly racemic 2a quantitatively (Entry 6). The reaction at 25 °C realized the kinetic resolution of rac-1a, but its enantioselectivity was far less satisfactory (krel = 5.5; Entry 7).
The trends of the molybdenum-catalyzed enantioselective RCM/KR reactions of rac-1b were similar to those of rac-1a. Precatalysts Mo*/(R)-L1 and Mo*/(R)-L3 showed fairly high enantioselectivity (krel = 117 and 76, respectively) but with unsatisfactory chemoselection, producing a considerable amount of dimeric 3b (Entries 8 and 12). On the other hand, Mo*/(R)-L2 was more reactive in the RCM/KR of rac-1b and did not catalyze the dimerization reaction (Entries 9-11). The reactivity of Mo*/(R)-L4 was too high, showing low enantioselectivity (Entries 13 and 14). The best result in the reaction of rac-1b was obtained using Mo*/(R)-L2 at the lower temperature (25 °C) under the concentrated conditions in C6D6 to give (R)-2b (97% ee, 47% yield) and recover (S)-1b (74% ee, 53%). The krel value of this reaction was estimated to be 146 (Entry 11).
Due to the sterically less-demanding cyclohexyl substituents in 1c (compared to tBu in 1a and SiMe3 in 1b), the RCM/KR reaction of rac-1c was generally less enantioselective (Entries 15–17). The optimized reaction conditions, as in Entries 4 and 11, were applied to the reaction of rac-1c, and the reasonably good enantioselectivity of krel = 9.9 was achieved as well (Entry 17). It should be noted that the reactions of rac-1c under Mo*/(R)-L2 catalysis did not produce the undesirable metathesis dimer 3c (Entries 16 and 17), as expected.
Next, the conditions optimized for the reaction of rac-1a-c were applied in the enantioselective RCM/KR reaction of substrate rac-1d, in which the methallyl group in rac-1a was replaced with a prenyl (3,3-dimethylallyl) substituent. It had been reported previously that the reaction of rac-1d in the presence of catalytic Mo*/(R)-L1 provided dimeric 3d in a 38% yield as a sole metathesis product and that bridged 2d was not detected. The unreacted substrate was recovered in 54%, of which the enantiopurity was as low as (S)-7% ee (Scheme 5, top) [44]. On the other hand, the reaction of rac-1d, as in Entry 4 in Table 1, afforded ferrocenophane (R)-2d of 58% ee in a 49% yield together with recovered (S)-1d of 56% ee in a 50% yield. The krel value for this reaction was determined to be 6.5 (Scheme 5, bottom).

3. Materials and Methods

3.1. General Information

All air- and/or moisture-sensitive reactions were conducted with standard Schlenk techniques under pre-dried nitrogen or with glovebox techniques under pre-purified argon. 1H NMR (at 400 MHz) and 13C NMR (at 100 MHz) chemical shifts were reported in ppm downfield of internal tetramethylsilane. Tetrahydrofuran was distilled from sodium benzophenone-ketyl under dry nitrogen prior to use. Benzene and C6D6 were dried over a Na/K alloy and distilled and deoxygenated under high-vacuum conditions prior to use. Chloroform-d was distilled and deoxygenated from P2O5 under high-vacuum conditions and stored in a glovebox. C5H4tBu2 [76], C5H4(SiMe3)2 [77], C5H4Cy2 [78], (pyrrolyl)2Mo(=CHCMe2Ph)(=N-C6H3-2,6-iPr2) [72], (R)-L1 [69], (S)-L3 [68], (R)-L4 [71], and the Grubbs-II catalyst [79,80] were prepared as reported. All the other chemicals were purchased from commercial suppliers and used without further purification, unless otherwise noted.

3.2. Preparation of Racemic Diallylferrocene Substrates rac-1a-d [44]

A typical procedure is hereby given for the synthesis of rac-1a. To a THF (8 mL) solution of Fe(acac)2 (2.54 g, 10.0 mmol) was added a solution of lithium 1-allyl-2,4-tBu2-cyclopentadienide, which was prepared from C5H3(allyl)tBu2 (2.18 g, 10.0 mmol) and nBuLi (1.60 M hexane solution, 6.3 mL, 10.1 mmol) in THF (25 mL), at –78 °C, and the mixture was stirred at 0 °C for 1h. After cooling the mixture to −78 °C, to this was added a solution of sodium methallylcyclopentadienide, which was prepared from C5H5-methallyl (1.06 g, 10.0 mmol) and NaH (240 mg, 10.0 mmol) in THF (10 mL). The resulting mixture was stirred at room temperature for 3 h. The mixture was diluted with hexane and filtered through a pad of Celite. After removal of the solvent under reduced pressure, the remaining dark-red oil was purified by column chromatography on alumina using hexane as an eluent, and following vacuum transfer gave rac-1a as a dark-red oil. The reaction conditions were not optimized. The characterization data of the diallylferrocene substrates rac-1a-d are given below.

3.3. Characterization Data of Racemic 1,1′-Diallylferrocene Substrates 1a-d [44]

rac-1-Allyl-1′-(2-methylallyl)-2,4-di(tert-butyl)ferrocene (1a). Yield: 45%. 1H NMR (CDCl3): δ 5.88–5.97 (m, 1H), 5.03 (d, J = 4.8 Hz, 1H), 4.99 (s, 1H), 4.62 (s, 1H), 4.59 (s, 1H), 4.13 (br, 1H), 4.08 (br, 1H), 3.93 (br, 2H), 3.76 (br, 1H), 3.72 (br, 1H), 3.22 (dd, J = 15.8 and 6.7 Hz, 1H), 3.07–3.12 (m, 1H), 3.01 (br, 2H), 1.65 (s, 3H), 1.27 (s, 9H), 1.18 (s, 9H). 3C{1H} NMR (CDCl3): δ 146.5, 138.2, 115.0, 110.8, 98.8, 96.8, 86.0, 81.7, 71.2, 70.1, 68.8, 68.1, 67.7, 64.1, 38.5, 34.3, 32.4, 31.9, 31.5, 30.5, 22.2. Anal. Calcd for C25H36Fe: C, 76.52; H, 9.25. Found: C, 76.30; H, 9.03. HRMS Calcd for C25H36Fe: 392.2165. Found: 392.2165.
rac-1-Allyl-1′-(2-methylallyl)-2,4-bis(trimethylsilyl)ferrocene (1b). Yield: 20%. 1H NMR (CDCl3): δ 5.87–5.97 (m, 1H), 5.01 (d, J = 3.9 Hz, 1H), 4.98 (s, 1H), 4.62 (s, 1H), 4.58 (s, 1H), 4.07 (s, 1H), 4.02 (s, 1H), 3.96 (s, 1H), 3.90 (s, 1H), 3.88 (s, 1H), 3.79 (s, 1H), 3.13 (d, J = 6.2 Hz, 2H), 3.00 (br, 2H), 1.64 (s, 3H), 0.26 (s, 9H), 0.22 (s, 9H). 13C{1H} NMR (CDCl3): δ 146.2, 138.2, 115.0, 110.3, 94.4, 86.5, 79.4, 78.2, 74.0, 73.3, 70.8, 69.7, 68.4, 68.1, 38.4, 34.2, 22.2, 0.6, 0.1. Anal. Calcd for C23H36FeSi2: C, 65.07; H, 8.55. Found: C, 65.13; H, 8.43. HRMS Calcd for C23H36FeSi2: 424.1703. Found: 424.1702.
rac-1-Allyl-1′-(2-methylallyl)-2,4-dicyclohexylferrocene (1c). Yield: 30%. 1H NMR (CDCl3): δ 5.91–6.01 (m, 1H), 4.99–5.06 (m, 2H), 4.63 (br, 1H), 4.58 (br, 1H), 3.77–3.97 (m, 6H), 2.94–3.09 (m, 4H), 2.13–2.25 (m, 3H), 1.84-1.93 (m, 3H), 1.66 (s, 3H), 1.61–1.78 (m, 6H), 1.15–1.40 (m, 9H), 0.91 (qd, J = 12.3 and 2.4 Hz, 1H). 13C{1H} NMR (CDCl3): δ 146.5, 138.0, 115.0, 110.1, 92.9, 92.5, 86.2, 82.9, 71.0, 70.5, 69.1, 69.0, 66.9, 64.0, 38.1, 37.4, 36.4, 36.3, 34.4, 34.3, 32.29, 32.26, 27.0, 26.9, 26.8 (2C), 26.6, 26.5, 22.3. Anal. Calcd for C29H40Fe: C, 78.36; H, 9.07. Found: C, 78.47; H, 9.20. HRMS Calcd for C29H40Fe: 444.2477. Found: 444.2484.
rac-1-Allyl-1′-(3-methyl-2-butenyl)-2,4-di(tert-butyl)ferrocene (1d). Yield: 33%. 1H NMR (CDCl3): δ 5.96–5.86 (m, 1H), 5.24–5.21 (m, 1H), 5.01 (d, J = 4.8 Hz, 1H), 4.98 (br, 1H), 4.15 (br, 1H), 4.09 (br, 1H), 3.95 (br, 1H), 3.92 (br, 1H), 3.83 (br, 1H), 3.79 (br, 1H), 3.19 (dd, J = 16.4 and 7.2 Hz, 1H), 3.07–2.99 (m, 3H), 1.68 (s, 3H), 1.65 (s, 3H), 1.25 (s, 9H), 1.16 (s, 9H). 13C{1H} NMR (CDCl3): δ 138.4, 131.1, 124.3, 114.9, 98.7, 96.7, 88.4, 81.7, 70.2, 69.1, 68.4, 67.9, 67.6, 63.8, 34.3, 32.4, 31.9, 31.5, 30.5, 27.8, 25.7, 17.8. Anal. Calcd for C26H38Fe: C, 76.84; H, 9.42. Found: C, 76.70; H, 9.59. HRMS Calcd for C26H38Fe: 406.2321. Found: 406.2328.

3.4. General Procedure for Molybdenum-Catalyzed Enantioselective RCM/Kinetic Resolution of rac-1

The detailed reaction conditions are summarized in Table 1. A mixture of Mo(=NC6H3-2,6-iPr2)(=CHCMe2Ph)(NC4H4)2 (5.4 mg, 10 μmol) and an appropriate chiral ligand L (11 μmol) were placed in a test tube (with a Teflon-sealed screw cap) and dissolved in dry benzene (3.0 mL) in a glovebox under pre-purified argon. The mixture was stirred for 15 min at room temperature, and then to this was added a solution of substrate rac-1 (0.10 mmol) in benzene (17.0 mL). The sealed test tube was taken out of the glovebox and was immersed in an oil bath maintained at 50 °C or 25 °C. After stirring the mixture for 24 h, the reaction was quenched by the addition of acetone (ca. 100 μL). The reaction mixture was passed through a short pad of silica gel (eluent: hexane/Et2O = 9/1). The volatiles were removed under reduced pressure, and the conversion of the reaction was determined by the 1H-NMR measurement of the crude residue. The residue was purified by preparative HPLC [LC-908 recycle HPLC system (Japan Analytical Industry Co., Ltd., Tokyo, Japan) with a GPC column (JAIGEL-H, chloroform, 3.5 mL/min)] to provide RCM product 2 and recovered substrate 1, respectively. Recovered unreacted 1 was treated with Grubbs-II catalyst (5 mol %) in benzene to give the corresponding 2 quantitatively, which was used for a chiral HPLC analysis. The absolute configurations of ferrocenophanes 2 as well as recovered unreacted substrates 1 were determined through the comparison of their signs of the specific rotations with those of the known compounds [44]. The characterization data of the RCM products and the conditions for the chiral HPLC analysis are listed below.

3.5. Characterization Data of Ferrocenophanes 2a-d [44]

1,1′-(3-Methyl-2-buten-1,4-diyl)-2,4-di(tert-butyl)ferrocene (2a). 1H NMR (CDCl3): δ 5.67 (t, J = 7.4 Hz, 1H), 4.14 (s, 1H), 4.02 (s, 1H), 3.99 (s, 1H), 3.89 (s, 1H), 3.82 (s, 1H), 3.73 (s, 1H), 3.28 (dd, J = 14.7 and 7.4 Hz, 1H), 3.12 (d, J = 14.7 Hz, 1H), 2.59-2.67 (m, 2H), 1.93 (s, 3H), 1.28 (s, 9H), 1.14 (s, 9H). 13C{1H} NMR (CDCl3): δ 138.1, 124.1, 99.9, 98.3, 85.3, 83.1, 70.6, 70.3, 66.8, 66.1, 66.0, 63.6, 32.6, 32.1, 31.4, 30.4, 29.4, 26.7, 24.8. Anal. Calcd for C23H32Fe: C, 75.82; H, 8.85. Found: C, 76.08; H, 8.91. HRMS Calcd for C23H32Fe: 364.1852. Found: 364.1856. [α]30D = +22 (c 0.50, CHCl3 for the sample of (R)-96% ee). Chiral HPLC Analysis Conditions: Chiralcel OD-H; eluent: hexane/iPrOH = 2000/1; flow rate: 1.0 mL/min; t1 = 17.7 min (R-isomer), t2 = 20.6 min (S-isomer).
1,1′-(3-Methyl-2-buten-1,4-diyl)-2,4-bis(trimethylsilyl)ferrocene (2b). 1H NMR (CDCl3): δ 5.76 (t, J = 7.8 Hz, 1H), 4.10 (s, 1H), 4.02 (s, 1H), 3.97 (s, 1H), 3.92 (s, 1H), 3.87 (s, 1H), 3.84 (s, 1H), 3.10 (d, J = 14.4 Hz, 1H), 3.02 (dd, J = 14.8 and 7.3 Hz, 1H), 2.75 (dd, J = 14.8 and 8.2 Hz, 1H), 2.69 (d, J = 14.4 Hz, 1H), 1.94 (s, 3H), 0.27 (s, 9H), 0.17 (s, 9H). 13C{1H} NMR (CDCl3): δ 138.1, 124.1, 96.3, 86.1, 78.5, 76.2, 74.8, 74.7, 70.1, 69.7, 66.5, 66.4, 29.4, 26.7, 25.0, 0.9, −0.1. Anal. Calcd for C21H32FeSi2: C, 63.61; H, 8.13. Found: C, 63.46; H, 8.04. HRMS Calcd for C21H32FeSi2: 396.1390. Found: 396.1395. [α]31D = +50 (c 2.0, CHCl3 for the sample of (R)-97% ee). Chiral HPLC Analysis Conditions: Chiralcel OD-H; eluent: hexane/iPrOH = 2000/1; flow rate: 1.0 mL/min; t1 = 15.5 min (R-isomer), t2 = 16.5 min (S-isomer).
1,1′-(3-Methyl-2-buten-1,4-diyl)-2,4-dicyclohexylferrocene (2c). 1H NMR (CDCl3): δ 5.74 (t, J = 8.0 Hz, 1H), 4.09–4.07 (m, 1H), 3.94–3.93 (m, 1H), 3.86–3.85 (m, 1H), 3.85–3.84 (m, 1H), 3.76–3.74 (m, 1H), 3.53-3.52 (m, 1H), 3.04 (d, J = 14.4 Hz, 1H), 2.94 (dd, J = 15.2 and 7.8 Hz, 1H), 2.81 (d, J = 14.4 Hz, 1H), 2.70 (dd, J = 14.8 and 7.8 Hz, 1H), 2.42–2.35 (m, 1H), 2.11-2.02 (m, 2H), 1.98-1.83 (m, 3H), 1.94 (s, 3H), 1.75-1.66 (m, 6H), 1.43–1.08 (m, 9H), 0.96–0.86 (m, 1H). 13C{1H} NMR (CDCl3): δ 138.1, 123.6, 93.8, 93.0, 86.3, 84.9, 71.9, 70.3, 67.6, 66.6, 66.5, 63.2, 37.33, 37.25, 36.3, 34.5, 34.4, 31.7, 31.5, 29.6, 27.1, 26.8, 26.74, 26.65, 23.6, 22.8, 14.3. Anal. Calcd for C27H36Fe: C, 77.88; H, 8.71. Found: C, 77.70; H, 8.92. HRMS Calcd for C27H36Fe: 416.2165. Found: 416.2166. [α]30D = +23 (c 1.3, CHCl3 for the sample of (S)-84% ee derived from recovered (S)-1c). Chiral HPLC Analysis Conditions: Chiralcel OD-H; eluent: hexane; flow rate: 0.5 mL/min; t1 = 78.3 min (R-isomer), t2 = 86.4 min (S-isomer).
1,1′-(2-Buten-1,4-diyl)-2,4-di(tert-butyl)ferrocene (2d). 1H NMR (CDCl3): δ 5.93–6.02 (m, 1H), 4.13 (s, 1H), 3.98–3.99 (m, 2H), 3.93 (s, 1H), 3.80 (s, 1H), 3.79 (s, 1H), 3.48 (dd, J = 15.1 and 6.0 Hz, 1H), 3.02 (dd, J = 15.1 and 6.0 Hz, 1H), 2.88 (dd, J = 14.6 and 7.3 Hz, 1H), 2.74 (dd, J = 14.6 and 7.3 Hz, 1H), 1.31 (s, 9H), 1.15 (s, 9H). 13C{1H} NMR (CDCl3): δ 131.6, 130.2, 99.6, 97.6, 86.2, 82.0, 70.6, 70.1, 66.9, 66.3, 65.7, 63.3, 32.6, 32.0, 31.4, 30.4, 24.3, 24.1. Anal. Calcd for C22H30Fe: C, 75.43; H, 8.63. Found: C, 75.29; H, 8.77. HRMS Calcd for C22H30Fe: 350.1695. Found: 350.1697. [α]31D = −8.2 (c 0.56, CHCl3 for the sample of (R)-58% ee). Chiral HPLC Analysis Conditions: Chiralcel OD-H × 2; eluent: hexane/iPrOH = 3000/1; flow rate: 0.1 mL/min; t1 = 94.2 min (R-isomer), t2 = 101.4 min (S-isomer).

3.6. Calculation of Selectivity Factors “krel” in Table 1 and in Scheme 5

Selectivity factors (krel, krel’) of the first-order KR reaction are calculated by equations 1 or 2 [74,75], where c (0 ≤ c ≤ 1) stands for the conversion of the reaction, and eesub and eepro (0 ≤ ee ≤ 1) are the enantiomeric excesses of recovered (S)-1 and RCM product (R)-2, respectively.
k rel = ln 1 c 1 ee sub ln 1 c 1 + ee sub
k rel = ln 1 c 1 + ee pro ln 1 c 1 ee pro
Among the three variables in Equations (1) and (2) (c, eesub, and eepro), conversion “c”, which was determined by the 1H-NMR measurement of the unpurified reaction mixture, contained up to 5% experimental errors. On the other hand, the eesub and eepro values were much more accurate. The %ee values of recovered substrate (S)-1 and RCM product (R)-2 in Table 1 and in Scheme 5 were determined by chiral HPLC analysis, which is usually reproducible within 1% of errors, if the two enantiomers are clearly separated in the chromatograms. The krel value (determined from eq. 1) and the krel’ value (determined from eq. 2) in a single reaction should be identical, in theory. Accordingly, logical conversion of the reaction could be determined from eesub (%ee of recovered (S)-1) and eepro (%ee of (R)-2) using Equations (1) and (2). The logical conversion values thus obtained showed reasonable agreement with the experimental observations. The krel values in Table 1 and in Scheme 5 were calculated using logical conversion, eesub, and eepro.

4. Conclusions

The molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of planar-chiral 1,1′-diallylferrocene derivatives 1a-d was reinvestigated utilizing the method of generating various catalytically active molybdenum species in situ. Among the molybdenum catalysts screened, Mo*/(R)-L2 showed the best overall performance with good enantioselectivity and excellent chemoselectivity. Since Mo*/(R)-L2 did not drive the undesirable bimolecular reaction giving 3, the RCM/KR reaction could be conducted under the concentrated conditions using this catalyst, which realized a shorter reaction time whilst retaining excellent enantioselectivity.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal14020123/s1: Figures S1–S8: 1H-NMR spectra of substrates 1a-d and RCM products 2a-d; Figures S9–S12: chiral HPLC chromatograms of 2a-d.

Author Contributions

Conceptualization, M.O.; investigation, H.I., K.M., and S.U.; writing—original draft preparation, M.O.; writing—review and editing, M.O.; supervision, M.O.; funding acquisition, M.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Grant-in-Aids for Scientific Research (21H01940) from MEXT, Japan.

Data Availability Statement

All the available data have been made available through the Supplementary Material.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Halterman, R.L. Synthesis and Applications of Chiral Cyclopentadienylmetal Complexes. Chem. Rev. 1992, 92, 965–994. [Google Scholar] [CrossRef]
  2. Wagner, G.; Herrmann, R. Chiral Ferrocene Derivatives. An Introduction. In Ferrocenes; Togni, A., Hayashi, T., Eds.; VCH: Weinheim, Germany, 1995; Chapter 4; pp. 173–218. [Google Scholar]
  3. Togni, A. Planar-chiral Ferrocenes: Synthetic Methods and Applications. Angew. Chem. Int. Ed. 1996, 35, 1475–1477. [Google Scholar] [CrossRef]
  4. Richards, C.J.; Locke, A.J. Recent Advances in the Generation of Non-racemic Ferrocene Derivatives and Their Application to Asymmetric Synthesis. Tetrahedron Asymmetry 1998, 9, 2377–2407. [Google Scholar] [CrossRef]
  5. Štěpnička, P.; Lamač, M. Synthesis and Catalytic Use of Planar Chiral and Polydentate Ferrocene Donors. In Ferrocenes; Štěpnička, P., Ed.; Wiley: Chichester, UK, 2008; Chapter 7; pp. 237–277. [Google Scholar]
  6. Alba, A.-N.R.; Rios, R. Kinetic Resolution: A Powerful Tool for the Synthesis of Planar-chiral Ferrocenes. Molecules 2009, 14, 4747–4757. [Google Scholar] [CrossRef] [PubMed]
  7. Schaarschmidt, D.; Lang, H. Selective Syntheses of Planar-chiral Ferrocenes. Organometallics 2013, 32, 5668–5704. [Google Scholar] [CrossRef]
  8. Arae, S.; Ogasawara, M. Catalytic Asymmetric Synthesis of Planar-chiral Transition-metal Complexes. Tetrahedron Lett. 2015, 56, 1751–1761. [Google Scholar] [CrossRef]
  9. Hayashi, T. Asymmetric Catalysis with Chiral Ferrocenylphosphine Ligands. In Ferrocenes; Togni, A., Hayashi, T., Eds.; VCH: Weinheim, Germany, 1995; Chapter 2; pp. 105–142. [Google Scholar]
  10. Togni, A. New Chiral Ferrocenyl Ligands for Asymmetric Catalysis. In Metallocenes; Togni, A., Halterman, R.L., Eds.; Wiley-VCH: Weinheim, Germany, 1998; Volume 2, Chapter 11; pp. 685–721. [Google Scholar]
  11. Colacot, T.J. A Concise Update on the Applications of Chiral Ferrocenyl Phosphines in Homogeneous Catalysis Leading to Organic Synthesis. Chem. Rev. 2003, 103, 3101–3118. [Google Scholar] [CrossRef] [PubMed]
  12. 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] [PubMed]
  13. Barbaro, P.; Bianchini, C.; Giambastiani, G.; Parisel, S.L. Progress in Stereoselective Catalysis by Metal Complexes with Chiral Ferrocenyl Phosphines. Coord. Chem. Rev. 2004, 248, 2131–3150. [Google Scholar] [CrossRef]
  14. Arrayás, R.G.; Adrio, J.; Carretero, J.C. Recent Applications of Chiral Ferrocene Ligands in Asymmetric Catalysis. Angew. Chem. Int. Ed. 2006, 45, 7674–7715. [Google Scholar] [CrossRef]
  15. Fu, G.C. Applications of Planar-chiral Heterocycles as Ligands in Asymmetric Catalysis. Acc. Chem. Res. 2006, 39, 853–860. [Google Scholar] [CrossRef] [PubMed]
  16. Ganter, C. Planar Chiral Phosphaferrocene-based Ligands. In Phosphorus Ligands in Asymmetric Catalysis; Börner, A., Ed.; Wiley-VCH: Weinheim, Germany, 2008; Chapter 4.3; pp. 393–407. [Google Scholar]
  17. Dai, L.-X.; Hou, X.-L. (Eds.) Chiral Ferrocenes in Asymmetric Catalysis; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  18. Toma, Š.; Csizmadiová, J.; Mečiarová, M.; Šebesta, R. Ferrocene Phosphane-Heteroatom/Carbon Bidentate Ligands in Asymmetric Catalysis. Dalton Trans. 2014, 16557–16579. [Google Scholar] [CrossRef] [PubMed]
  19. 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] [PubMed]
  20. Butsugan, Y.; Araki, S.; Watanabe, M. Enantioselective Addition of Dialkylzinc to Aldehydes Catalyzed by Chiral Ferrocenyl Amino Alcohols. In Ferrocenes; Togni, A., Hayashi, T., Eds.; VCH: Weinheim, Germany, 1995; Chapter 3; pp. 143–169. [Google Scholar]
  21. Fu, G.C. Enantioselective Nucleophilic Catalysis with “Planar-chiral” Heterocycles. Acc. Chem. Res. 2000, 33, 412–420. [Google Scholar] [CrossRef] [PubMed]
  22. Fu, G.C. Asymmetric Catalysis with “Planar-chiral” Derivatives of 4-(Dimethylamino)pyridine. Acc. Chem. Res. 2004, 37, 542–547. [Google Scholar] [CrossRef] [PubMed]
  23. Fu, G.C. Planar-Chiral Heterocycles as Enantioselective Organocatalysts. In Asymmetric Synthesis, 2nd ed.; Christmann, M., Brase, S., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp. 195–199. [Google Scholar]
  24. Marion, N.; Fu, G.C. Applications of Aza- and Phosphaferrocenes and Related Compounds in Asymmetric Catalysis. In Chiral Ferrocenes in Asymmetric Catalysis; Dai, L.-X., Hou, X.-L., Eds.; Wiley-VCH: Weinheim, Germany, 2010; pp. 307–335. [Google Scholar]
  25. Zhu, J.-C.; Cui, D.-X.; Li, Y.-D.; Jiang, R.; Chen, W.-P.; Wang, P.-A. Ferrocene as a Privileged Framework for Chiral Organocatalysts. ChemCatChem 2018, 10, 907–919. [Google Scholar] [CrossRef]
  26. Bernardo, O.; González-Pelayo, S.; López, L.A. Synthesis and Applications of Ferrocene-Fused Nitrogen Heterocycles. Eur. J. Inorg. Chem. 2022, e202100911. [Google Scholar] [CrossRef]
  27. Peluso, P.; Mamane, V. Ferrocene Derivatives with Planar Chirality and Their Enantioseparation by Liquid-Phase Techniques. Electrophoresis 2023, 44, 158–189. [Google Scholar] [CrossRef]
  28. Izumi, T.; Hino, T. Enzymatic Resolution of Planar Chiral Ferrocene Derivatives. J. Chem. Technol. Biotechnol. 1992, 55, 325–331. [Google Scholar] [CrossRef]
  29. Lambusta, D.; Nicolosi, G.; Patti, A.; Piattelli, M. Lipase-Mediated Resolution of Racemic 2-Hydroxymethyl-1-methylthioferrocene. Tetrahedron Lett. 1996, 37, 127–130. [Google Scholar] [CrossRef]
  30. Patti, A.; Lambusta, D.; Piattelli, M.; Nocolosi, G. Lipase-Mediated Resolution of 2-Hydroxymethyl-1-iodoferrocene: Synthesis of Ferrocenes and Biferrocenes with Planar Chirality. Tetrahedron Asymmetry 1998, 9, 3073–3080. [Google Scholar] [CrossRef]
  31. Marquarding, D.; Klusacek, H.; Gokel, G.; Hoffmann, P.; Ugi, I. Correlation of Central and Planar Chirality in Ferrocene Derivatives. J. Am. Chem. Soc. 1970, 92, 5389–5393. [Google Scholar] [CrossRef]
  32. Riant, O.; Samuel, O.; Kagan, H.B. A General Asymmetric Synthesis of Ferrocenes with Planar Chirality. J. Am. Chem. Soc. 1993, 115, 5835–5836. [Google Scholar] [CrossRef]
  33. Rebiere, F.; Riant, O.; Ricard, L.; Kagan, H.B. Asymmetric Synthesis and Highly Diastereoselective ortho-Lithiation of Ferrocenyl Sulfoxides. Application to the Synthesis of Ferrocenyl Derivatives with Planar Chirality. Angew. Chem. Int. Ed. Engl. 1993, 32, 568–570. [Google Scholar] [CrossRef]
  34. Richards, C.J.; Damalidis, T.; Hibbs, D.E.; Hursthouse, M.B. Synthesis of 2-[2-(Diphenylphosphino)ferrocenyl]oxazoline Ligands. Synlett 1995, 1995, 74–76. [Google Scholar] [CrossRef]
  35. Sammakia, T.; Latham, H.A.; Schaad, D.R. Highly Diastereoselective Ortho Lithiations of Chiral Oxazoline-Substituted Ferrocenes. J. Org. Chem. 1995, 60, 10–11. [Google Scholar] [CrossRef]
  36. Riant, O.; Samuel, O.; Flessner, T.; Tauten, S.; Kagan, H.B. An Efficient Asymmetric Synthesis of 2-Substituted Ferrocenecarboxaldehydes. J. Org. Chem. 1997, 62, 6733–6745. [Google Scholar] [CrossRef]
  37. Enders, D.; Peters, R.; Lochtman, R.; Raabe, G. Asymmetric Synthesis of Novel Ferrocenyl Ligands with Planar and Central Chirality. Angew. Chem. Int. Ed. 1999, 38, 2421–2423. [Google Scholar] [CrossRef]
  38. Geisler, F.M.; Helmchen, G. A Straightforward Synthesis of (3S)-4-Methoxybutane-1,3-diol and Its Use as Chiral Auxiliary for the Preparation of (pS)-1-(Diphenylphosphino)-2-formyl-1′,2′,3′,4′,5′-pentamethylferrocene. Synthesis 2006, 2006, 2201–2205. [Google Scholar]
  39. Wölfle, H.; Kopacka, H.; Wurst, K.; Ongania, K.-H.; Görtz, H.-H.; Preishuber-Pflügl, P.; Bildstein, B. Planar Chiral Ferrocene Salen-type Ligands Featuring Additional Central and Axial Chirality. J. Organomet. Chem. 2006, 691, 1197–1215. [Google Scholar] [CrossRef]
  40. Mamane, V. The Diastereoselective Ortho-lithiation of Kagan’s Ferrocenyl Acetal. Generation and Reactivity of Chiral 2-Substituted Ferrocenecarboxaldehydes. Tetrahedron Asymmetry 2010, 21, 1019–1029. [Google Scholar] [CrossRef]
  41. Siegel, S.; Schmalz, H.-G. Insertion of Carbenoids into Cp-H Bonds of Ferrocenes: An Enantioselective-Catalytic Entry to Planar-Chiral Ferrocenes. Angew. Chem. Int. Ed. Engl. 1997, 36, 2456–2458. [Google Scholar] [CrossRef]
  42. Genet, C.; Canipa, S.J.; O’Brien, P.; Taylor, S.J. Catalytic Asymmetric Synthesis of Ferrocenes and P-Stereogenic Bisphosphines. J. Am. Chem. Soc. 2006, 128, 9336–9337. [Google Scholar] [CrossRef] [PubMed]
  43. Bueno, A.; Rosol, M.; García, J.; Moyano, A. Asymmetric Dihydroxylation of 2-Substituted 1-Vinylferrocenes: The First Non-Enzymatic Kinetic Resolution of Planar-Chiral Ferrocenes. Adv. Synth. Catal. 2006, 348, 2590–2596. [Google Scholar] [CrossRef]
  44. Ogasawara, M.; Watanabe, S.; Fan, L.; Nakajima, K.; Takahashi, T. Kinetic Resolution of Planar-Chiral Ferrocenes by Mo-Catalyzed Enantioselective Metathesis. Organometallics 2006, 25, 5201–5203. [Google Scholar] [CrossRef]
  45. Zhu, D.-Y.; Chen, P.; Xia, J.-B. Synthesis of Planar Chiral Ferrocenes by Transition-Metal-Catalyzed Enantioselective C-H Activation. ChemCatChem 2016, 8, 68–73. [Google Scholar] [CrossRef]
  46. Gao, D.-W.; Gu, Q.; Zheng, C.; You, S.-L. Synthesis of Planar Chiral Ferrocenes via Transition-Metal-Catalyzed Direct C−H Bond Functionalization. Acc. Chem. Res. 2017, 50, 351–365. [Google Scholar] [CrossRef]
  47. Liu, C.-X.; Gu, Q.; You, S.-L. Asymmetric C-H Bond Functionalization of Ferrocenes: New Opportunities and Challenges. Trends Chem. 2020, 2, 737–749. [Google Scholar] [CrossRef]
  48. Ogasawara, M. Enantioselective Preparation of Planar-Chiral Transition Metal Complexes by Asymmetric Olefin-Metathesis Reactions in Metal Coordination Spheres. Chem. Rec. 2021, 21, 3509–3519. [Google Scholar] [CrossRef] [PubMed]
  49. Bauer, E.B.; Gladysz, J.A. Metal-Catalyzed Olefin Metathesis in Metal Coordination Spheres. In Handbook of Metathesis; Grubbs, R.H., Ed.; Wiley-VCH: Weinheim, Germany, 2003; Volume 2, Chapter 2.11; pp. 403–431. [Google Scholar]
  50. Fiedler, T.; Gladysz, J.A. Multifold Ring-Closing Olefin Metatheses in Syntheses of Organometallic Molecules with Unusual Connectivities. In Olefin Metathesis; Grela, K., Ed.; Wiley-VCH: Weinheim, Germany, 2014; pp. 311–328. [Google Scholar]
  51. Ogasawara, M.; Nagano, T.; Hayashi, T. Metathesis Route to Bridged Metallocenes. J. Am. Chem. Soc. 2002, 124, 9068–9069, Erratum in J. Am. Chem. Soc. 2002, 124, 12626. [Google Scholar] [CrossRef]
  52. Ogasawara, M.; Wu, W.-Y.; Arae, S.; Nakajima, K.; Takahashi, T. Inter- versus Intraannular Ring-Closing Metathesis on Polyallylferrocenes: Five-Fold RCM within a Single Molecule. Organometallics 2013, 32, 6593–6598. [Google Scholar] [CrossRef]
  53. Locke, A.J.; Jones, C.; Richards, C.J. A Rapid Approach to Ferrocenophanes via Ring-Closing Metathesis. J. Organomet. Chem. 2001, 637-639, 669–676. [Google Scholar] [CrossRef]
  54. Hüerländere, D.; Kleigrewe, N.; Kehr, G.; Erker, G.; Fröhlich, R. Synthesis, Structural and Chemical Characterization of Unsaturated C4- and C10-Bridged Group-4 ansa-Metallocenes Obtained Through a Ring-Closing Olefin Metathesis Reaction. Eur. J. Inorg. Chem. 2002, 2002, 2633–2642. [Google Scholar] [CrossRef]
  55. Buchowicz, W.; Jerzykiewicz, L.B.; Krasińska, A.; Losi, S.; Pietzykowski, A.; Zanello, P. ansa-Nickelocenes by the Ring-Closing Metathesis Route:  Syntheses, X-ray Crystal Structures, and Physical Properties. Organometallics 2006, 25, 5076–5082. [Google Scholar] [CrossRef]
  56. Schrock, R.R.; Murdzek, J.S.; Bazan, G.C.; Robbins, J.; DiMare, M.; O’Regan, M. Synthesis of Molybdenum Imido Alkylidene Complexes and Some Reactions Involving Acyclic Olefins. J. Am. Chem. Soc. 1990, 112, 3875–3886. [Google Scholar] [CrossRef]
  57. Schrock, R.R. The Discovery and Development of High Oxidation State Mo and W Imido Alkylidene Complexes for Alkene Metathesis. In Handbook of Metathesis; Grubbs, R.H., Ed.; Wiley-VCH: Weinheim, Germany, 2003; Volume 1, Chapter 1.3; pp. 8–32. [Google Scholar]
  58. Nguyen, S.T.; Johnson, L.K.; Grubbs, R.H.; Ziller, J.W. Ring-Opening Metathesis Polymerization (ROMP) of Norbornene by a Group VIII Carbene Complex in Protic Media. J. Am. Chem. Soc. 1992, 114, 3974–3975. [Google Scholar] [CrossRef]
  59. Schwab, P.; Grubbs, R.H.; Ziller, J.W. Synthesis and Applications of RuCl2(=CHR’)(PR3)2: The Influence of the Alkylidene Moiety on Metathesis Activity. J. Am. Chem. Soc. 1996, 118, 100–110. [Google Scholar] [CrossRef]
  60. Scholl, M.; Ding, S.; Lee, C.W.; Grubbs, R.H. Synthesis and Activity of a New Generation of Ruthenium-Based Olefin Metathesis Catalysts Coordinated with 1,3-Dimesityl-4,5-dihydroimidazol-2-ylidene Ligands. Org. Lett. 1999, 1, 953–956. [Google Scholar] [CrossRef] [PubMed]
  61. Nguyen, S.T.; Trnka, T.M. The Discovery and Development of Well-Defined, Ruthenium-Based Olefin Metathesis Catalysts. In Handbook of Metathesis; Grubbs, R.H., Ed.; Wiley-VCH: Weinheim, Germany, 2003; Volume 1, Chapter 1.6; pp. 61–85. [Google Scholar]
  62. Ogasawara, M.; Wu, W.-Y.; Arae, S.; Watanabe, S.; Morita, T.; Takahashi, T.; Kamikawa, K. Kinetic Resolution of Planar-Chiral (h6-Arene)chromium Complexes by Molybdenum-Catalyzed Asymmetric Ring-Closing Metathesis. Angew. Chem. Int. Ed. 2012, 51, 2951–2955. [Google Scholar] [CrossRef]
  63. Ogasawara, M.; Arae, S.; Watanabe, S.; Nakajima, K.; Takahashi, T. Kinetic Resolution of Planar-Chiral 1,2-Disubstituted Ferrocenes by Molybdenum-Catalyzed Asymmetric Intraannular Ring-Closing Metathesis. Chem. Eur. J. 2013, 19, 4151–4154. [Google Scholar] [CrossRef]
  64. Ogasawara, M.; Arae, S.; Watanabe, S.; Nakajima, K.; Takahashi, T. Kinetic Resolution of Planar-Chiral Ferrocenylphosphine Derivatives by Molybdenum-Catalyzed Asymmetric Ring-Closing Metathesis and Their Application in Asymmetric Catalysis. ACS Catal. 2016, 6, 1308–1315. [Google Scholar] [CrossRef]
  65. Ogasawara, M.; Watanabe, S.; Nakajima, K.; Takahashi, T. Enantioselective Synthesis of Planar-Chiral Phosphaferrocenes by Molybdenum-Catalyzed Asymmetric Interannular Ring-Closing Metathesis. J. Am. Chem. Soc. 2010, 132, 2136–2137. [Google Scholar] [CrossRef]
  66. Kamikawa, K.; Arae, S.; Wu, W.-Y.; Nakamura, C.; Takahashi, T.; Ogasawara, M. Simultaneous Induction of Axial and Planar Chirality in Arene-Chromium Complexes by Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis. Chem. Eur. J. 2015, 21, 4954–4957. [Google Scholar] [CrossRef]
  67. Ogasawara, M.; Tseng, Y.-Y.; Uryu, M.; Ohya, N.; Chang, N.; Ishimoto, H.; Arae, S.; Takahashi, T.; Kamikawa, K. Molybdenum-Catalyzed Enantioselective Synthesis of Planar-Chiral (h5-Phosphacyclopentadienyl)manganese(I) Complexes and Application in Asymmetric Catalysis. Organometallics 2017, 36, 4061–4069. [Google Scholar] [CrossRef]
  68. Alexander, J.B.; La, D.S.; Cefalo, D.R.; Hoveyda, A.H.; Schrock, R.R. Catalytic Enantioselective Ring-Closing Metathesis by a Chiral Biphen-Mo Complex. J. Am. Chem. Soc. 1998, 120, 4041–4042. [Google Scholar] [CrossRef]
  69. Aeilts, S.L.; Cefalo, D.R.; Bonitatebus, P.J., Jr.; Houser, J.H.; Hoveyda, A.H.; Schrock, R.R. A Readily Available and User-Friendly Chiral Catalyst for Efficient Enantioselective Olefin Metathesis. Angew. Chem. Int. Ed. 2001, 40, 1452–1456. [Google Scholar] [CrossRef]
  70. Schrock, R.R.; Jamieson, J.Y.; Dolman, S.J.; Miller, S.A.; Bonitatebus, P.J., Jr.; Hoveyda, A.H. Synthesis of Enantiomerically Pure Molybdenum Imido Alkylidene Catalysts for Asymmetric Olefin Metathesis that Contain Diolate Ligands Based on 3,3′-Disubstituted Octahydrobinaphtholate and 2,6-Dichlorophenylimido Combinations. Organometallics 2002, 21, 409–417. [Google Scholar] [CrossRef]
  71. Singh, R.; Czekelius, C.; Schrock, R.R.; Müller, P.; Hoveyda, A.H. Molybdenum Imido Alkylidene Metathesis Catalysts that Contain Electron Withdrawing Biphenoxides or Biphenolates. Organometallics 2007, 26, 2528–2539. [Google Scholar] [CrossRef] [PubMed]
  72. Hock, A.S.; Schrock, R.R.; Hoveyda, A.H. Dipyrrolyl Precursors to Bisalkoxide Molybdenum Olefin Metathesis Catalysts. J. Am. Chem. Soc. 2006, 128, 16373–16375. [Google Scholar] [CrossRef] [PubMed]
  73. Bunel, E.; Valle, L.; Manriquez, J.M. Pentamethylcyclopentadienyl Acetylacetonate Complexes of Iron(II), Cobalt(II), and Nickel(II). Convenient Synthetic Entries to Mono-h5-C5Me5 Derivatives. Organometallics 1985, 4, 1680–1682. [Google Scholar] [CrossRef]
  74. Kagan, H.B.; Fiaud, J.C. Kinetic Resolution. In Topics in Stereochemistry; Eliel, E.L., Wilen, S.H., Eds.; John Wiley & Sons: New York, NY, USA, 1988; Volume 18, pp. 249–330. [Google Scholar]
  75. Vedejs, E.; Jure, M. Efficiency in Nonenzymatic Kinetic Resolution. Angew. Chem. Int. Ed. 2005, 44, 3974–4001. [Google Scholar] [CrossRef] [PubMed]
  76. Venier, C.G.; Casserly, E.W. Di-tert-butylcyclopentadiene and Tri-tert-butylcyclopentadiene. J. Am. Chem. Soc. 1990, 112, 2808–2809. [Google Scholar] [CrossRef]
  77. Ustynyuk, Y.A.; Kisin, A.V.; Pribytkova, I.M.; Zenkin, A.A.; Antonova, N.D. Nuclear Magnetic Resonance Spectroscopy of Metal Cyclopentadienyls X. Proton Magnetic Resonance Spectra of, and Dynamic Behaviour in, Bis(trimethylsilyl)cyclopentadiene. J. Organomet. Chem. 1972, 42, 47–63. [Google Scholar] [CrossRef]
  78. Clark, T.J.; Killian, C.M.; Luthra, S.; Nile, T.A. Synthesis and Properties of Sterically Congested Cyclopentadienes and Their Transition Metal Complexes. J. Organomet. Chem. 1993, 462, 247–257. [Google Scholar] [CrossRef]
  79. Trnka, T.M.; Morgan, J.P.; Sanford, M.S.; Wilhelm, T.E.; Scholl, M.; Choi, T.-L.; Ding, S.; Day, M.D.; Grubbs, R.H. Synthesis and Activity of Ruthenium Alkylidene Complexes Coordinated with Phosphine and N-Heterocyclic Carbene Ligands. J. Am. Chem. Soc. 2003, 125, 2546–2558. [Google Scholar] [CrossRef]
  80. Garber, S.B.; Kingsbury, J.S.; Gray, B.L.; Hoveyda, A.H. Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122, 8168–8179. [Google Scholar] [CrossRef]
Figure 1. Enantiomeric pair of a planar-chiral ferrocene.
Figure 1. Enantiomeric pair of a planar-chiral ferrocene.
Catalysts 14 00123 g001
Figure 2. Representative chiral Schrock–Hoveyda molybdenum-alkylidene precatalysts [68,69,70,71].
Figure 2. Representative chiral Schrock–Hoveyda molybdenum-alkylidene precatalysts [68,69,70,71].
Catalysts 14 00123 g002
Scheme 1. Reaction generating chiral molybdenum-alkylidene precatalysts in situ from complex A and proligand (R)-L [72].
Scheme 1. Reaction generating chiral molybdenum-alkylidene precatalysts in situ from complex A and proligand (R)-L [72].
Catalysts 14 00123 sch001
Scheme 2. Mo*/(R)-L1-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocene rac-1a [44].
Scheme 2. Mo*/(R)-L1-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocene rac-1a [44].
Catalysts 14 00123 sch002
Figure 3. Structural requirements for ferrocene substrates showing high selectivity in molybdenum-catalyzed enantioselective RCM/KR reaction.
Figure 3. Structural requirements for ferrocene substrates showing high selectivity in molybdenum-catalyzed enantioselective RCM/KR reaction.
Catalysts 14 00123 g003
Scheme 3. Preparation of racemic planar-chiral ferrocene substrates rac-1a-c.
Scheme 3. Preparation of racemic planar-chiral ferrocene substrates rac-1a-c.
Catalysts 14 00123 sch003
Scheme 4. Plausible stereochemical pathways of the molybdenum-catalyzed enantioselective RCM KR of rac-1.
Scheme 4. Plausible stereochemical pathways of the molybdenum-catalyzed enantioselective RCM KR of rac-1.
Catalysts 14 00123 sch004
Scheme 5. Molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocene rac-1d.
Scheme 5. Molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocene rac-1d.
Catalysts 14 00123 sch005
Table 1. Molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocenes rac-1a-c a.
Table 1. Molybdenum-catalyzed enantioselective ring-closing metathesis/kinetic resolution of racemic planar-chiral ferrocenes rac-1a-c a.
Catalysts 14 00123 i001
EntrySubstrate bLigandConditionsYields (%) of 1/2/3 c% ee of (S)-1/(R)-2 d,ekrel f
11a (0.005)(R)-L150 °C, 24 h46/52/299/91109
21a (0.005)(R)-L250 °C, 24 h45/55/097/8865
31a (0.005)(R)-L225 °C, 24 h53/47/067/9695
4 g1a (0.05)(R)-L225 °C, 1 h55/45/075/96110
51a (0.005)(R)-L350 °C, 24 h35/45/2094/9175
61a (0.005)(R)-L450 °C, 24 h0/100/0--- h/--- h--- h
71a (0.005)(R)-L425 °C, 24 h12/88/099/185.5
81b (0.005)(R)-L150 °C, 24 h51/38/1189/95117
91b (0.005)(R)-L250 °C, 24 h45/55/092/8851
101b (0.005)(R)-L225 °C, 24 h45/55/097/9189
11 g1b (0.05)(R)-L225 °C, 2 h53/47/074/97146
121b (0.005)(R)-L350 °C, 24 h44/52/491/9276
131b (0.005)(R)-L450 °C, 24 h0/100/0--- h/--- h--- h
141b (0.005)(R)-L425 °C, 24 h20/80/090/264.4
151c (0.005)(R)-L150 °C, 24 h28/52/2098/5415
161c (0.005)(R)-L250 °C, 24 h1/99/0--- h/5--- h
17 g1c (0.05)(R)-L225 °C, 1 h45/55/084/599.9
a The reaction was carried out with rac-1 (0.10 mmol) in benzene using a molybdenum catalyst generated in situ (10 mol %), unless otherwise noted. b Initial concentration of substrate 1 in parentheses. c Determined via the 1H-NMR analysis of the crude reaction mixture. d Determined by chiral HPLC analysis (see Materials and Methods Section for detail). e Enantiomeric excess of recovered 1a-c was determined after converting them into the corresponding 2a-c via the RCM reaction using the Grubbs-II catalyst. f Calculated based on a first-order equation [74,75]. g The reaction was carried out in C6D6. h Not determined.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Imazu, H.; Masaoka, K.; Uike, S.; Ogasawara, M. Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferrocenes. Catalysts 2024, 14, 123. https://doi.org/10.3390/catal14020123

AMA Style

Imazu H, Masaoka K, Uike S, Ogasawara M. Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferrocenes. Catalysts. 2024; 14(2):123. https://doi.org/10.3390/catal14020123

Chicago/Turabian Style

Imazu, Haruna, Kakeru Masaoka, Saki Uike, and Masamichi Ogasawara. 2024. "Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferrocenes" Catalysts 14, no. 2: 123. https://doi.org/10.3390/catal14020123

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop