Next Article in Journal
Tuning the Inter-Chromophore Electronic Coupling in Perylene Diimide Dimers with Rigid Covalent Linkers
Previous Article in Journal
A New Fluorescence Band of Anthocyanins as a Simple Oxidation Biomarker of Food Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enantiomerically Pure ansa-η5-Complexes of Transition Metals as an Effective Tool for Chirality Transfer

by
Pavel V. Kovyazin
,
Leonard M. Khalilov
and
Lyudmila V. Parfenova
*
Institute of Petrochemistry and Catalysis, Ufa Federal Research Center, Russian Academy of Sciences, 141 Prospekt Oktyabrya, 450075 Ufa, Russia
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(12), 2511; https://doi.org/10.3390/molecules30122511
Submission received: 28 April 2025 / Revised: 2 June 2025 / Accepted: 4 June 2025 / Published: 8 June 2025
(This article belongs to the Special Issue Advances in Metallocene Chemistry)

Abstract

Chiral ansa-η5-complexes of transition metals have shown remarkable efficacy in organometallic synthesis and catalysis. Additionally, enantiomerically pure ansa-complexes hold promise for the development of novel chiral materials and pharmaceuticals. The discovery and synthesis of a diverse range of group IVB and IIIB metal complexes represents a significant milestone in the advancement of stereoselective catalytic methods for constructing metal-C, C-C, C-H, and C-heteroatom bonds. The synthesis of enantiomerically pure metallocenes can be accomplished through several strategies: utilizing optically active precursors of η5-ligands, separation of diastereomers of complexes with enantiomerically pure agents, and synthesis via the stereocontrolled reactions of enantiomerically pure σ-complexes with prochiral anions of η5-ligands. This review focuses on the analysis of various nuances of the synthesis of enantiomerically pure ansa-η5-complexes of titanium and lanthanum families. Their applicability as effective catalysts in asymmetric carbomagnesiation, carbo- and cycloalumination, oligo- and polymerization, Diels–Alder cycloaddition, reactions of zirconaaziridines, cyclization, hydrosilylation, hydrogenation, hydroamination, and other processes are highlighted as well.

1. Introduction

The discovery of metallocene compounds in the mid-20th century stimulated the development of several significant areas within organometallic chemistry [1,2]. Metallocenes have exhibited outstanding catalytic properties in the formation of metal–carbon (M-C), carbon–carbon (C-C), and carbon–hydrogen (C-H) bonds. Among these, single-site alkene polymerization using catalytic systems composed of IV subgroup metal η5-complexes and organoaluminum or -boron compounds has emerged as a particularly promising avenue [3,4,5,6,7,8,9,10,11]. Furthermore, various catalytic methods of alkene and acetylene transformations, such as di- and oligomerization [12,13,14,15,16,17,18,19], hydrometalation [20,21,22,23], carbometalation [24,25,26,27], and cyclometalation [21,24,28,29,30], are well known. Particular interest is drawn to the application of metallocenes as biosensors, antitumor agents, and other pharmaceuticals in medicine [31,32,33,34,35].
The high activity, chemo-, and stereoselectivity of metallocene complexes can be attributed to their unique structural characteristics. These compounds effectively ensure the stability of the electronic and steric environment of the transition metal center due to the high bond energy of the metal–ligand interactions. Additionally, they allow for variations in electrophilicity and geometry of the catalytically active sites through extensive structural modifications of the π-ligands. The introduction of bridging groups has led to the development of ansa-complexes with distinct properties. The global literature describes the synthesis of numerous ansa-metallocene complexes [36,37,38,39,40], which have found diverse applications in organic and organometallic chemistry. Notably, the binding of η5-ligands increases electrophilicity and accessibility of the catalytically active center by altering the coordination geometries of the π-ligands on the metal, while also fixing the geometry of the complex. This, combined with planar chirality, endows these compounds with significant potential for catalytic stereo- and enantioselective construction of C-C, C-H, C-N, and C-O bonds.
The generation of stereoisomers of ansa-complexes is facilitated by the presence of both enantiotopic and diastereotopic faces in the bridging ligands, whose reaction with metal salts or complexes is accompanied by a stereodifferentiating effect [39,41]. Stereoisomers can be synthesized through various pathways. First, when using nonchiral metal salts or complexes, additional separation of the racemic mixture of the resulting ansa-derivatives with an enantiomerically pure reagent is necessary (Scheme 1, path A). Second, both diastereoselective and enantioselective synthesis of ansa-complexes can be achieved from enantiomerically pure ligands that contain stereogenic centers in both the substituents at the η5-ligands and the bridging group (Scheme 1, path B). Third, the use of metal complexes with chiral elements in σ-substituents can determine stereoselectivity at the π-coordination stage of prochiral ligands (Scheme 1, path C).
Enantiomerically pure ansa-complexes of III and IV subgroup metals, characterized by fixed geometries of η5-ligands, have demonstrated high efficiency in generating stereogenic centers on carbon atoms in prochiral substrates such as alkenes, ketones, imines, etc. Several reviews in the literature have focused on the synthesis of optically active metallocenes and their applications in organic chemistry. Reviews [39,41] detail methods for synthesizing chiral metallocenes, including those with ansa-bound η5-ligands. Review [36] discusses the synthesis of enantiopure planar chiral organometallic complexes via diastereoselective complexation methodologies. Additionally, the mini-review [40] presents various approaches for the synthesis of chiral ansa-metallocene complexes of Ti, Zr, and Hf that incorporate carbon bridging groups. Insights into the use of chiral C2-symmetric complexes of Zr and Ti as catalysts for the asymmetric construction of C-C and C-heteroatom bonds are provided in review articles [25,42].
This review considers and systematizes information on the approaches to the synthesis of enantiomerically pure ansa-η5-complexes of group IIIB and IVB transition metals, as well as various aspects of their application for asymmetric induction in the construction of C-C, M-C, C-H, and C-heteroatom bonds.

2. Synthesis of Enantiomerically Pure ansa-bis-η5-Complexes of Group III and IV Transition Metals

2.1. Path A: Resolution of Racemic Complexes into Enantiomers via Derivatization

The primary method for achieving enantiomerically enriched ansa-metallocenes is the kinetic resolution of rac-isomers of complexes through the use of derivatizing reagents (Scheme 1, path A).
In 1979, Brintzinger et al. first synthesized the optical isomer p-S, p-S-propandienyl-bis-(tert-butylcyclopentadienyl)titanium binaphtholate (S,p-S,p-S-3a) by the reaction of dichloride rac-complex 2a with an equivalent amount of S-binaphthol (S-BINOL) in the presence of sodium amide in toluene (Scheme 2). The diastereomerically pure binaphtholate complex 3a was isolated by column chromatography [43] and subsequently converted to the corresponding dichloride. Later, Collins et al. successfully resolved rac-ethylene-bis-tert-butylcyclopentadienyltitanium dichloride 2b into its enantiomers using S-binaphthol [44].
A combination of photoisomerization and treatment with lithium R-binaphtholate was employed to obtain enantiomerically pure Zr ansa-cyclopentadienyl complexes featuring Si-bound ligands [45] (Scheme 3). This approach yielded compounds 7 and 8 with yields reaching up to 88% and enantiomeric purity exceeding 95%, as determined by 1H NMR spectroscopy of the carboxylate complex R,R,p-S-10.
The enantiomers of the bis-tetrahydroindenyl complex rac-11 were derivatized with S-binaphthol [46] (Scheme 4). This reaction produced two enantiomers, S,p-S,p-S-12 and p-R,p-R-11, with yields of 77% and 70%, respectively, while the enantiomeric purity of p-R,p-R-11 was 75%. Following isolation, the binaphtholate complex S,p-S,p-S-12 was converted into a dimethyl analog, which was treated with hydrochloric acid to yield a red precipitate of titanocene dichloride p-S,p-S-11.
Optically active acids present an alternative to enantiomerically pure binaphthols for the resolution of chiral ansa-metallocenes [47]. The corresponding carboxylate complexes were synthesized through various methods: (i) converting metallocene dichloride complexes into dimethyl analogs followed by their reaction with optically active carboxylic acids; (ii) reacting metallocene dichlorides with lithium or sodium salts of acids; and (iii) treating metallocene dichlorides with acids in the presence of triethylamine. It was demonstrated that metallocenes react almost completely with optically active acids, such as R-acetylmandelic acid in the presence of Et3N (Scheme 5), yielding diastereomeric carboxylate complexes 14 and 15. These complexes can be efficiently separated by fractional crystallization, achieving yields of 77–88% and minimizing the loss of the desired diastereomerically pure complexes. The conversion of carboxylates back to dichloride complexes p-S,p-S-11 and p-S,p-S-13 was accomplished via dimethyl derivatives followed by treatment with hydrochloric acid in ether (yield > 50%).
The derivatization method for ansa-complexes described above remains the most widely utilized approach. Subsequent research has focused on refining this method to increase the yield and enantiomeric purity of the target metal complexes.
To minimize the loss of enantiomers of complex 11, the racemic mixture was subjected to reflux with sodium and S-binaphthol (Scheme 6) [48,49]. This process increased the yield of enantiomerically pure complex S,p-S,p-S-12 to 78%, while dichloride p-R,p-R-11 was isolated at a yield of 40%. Following treatment of S,p-S,p-S-12 with a solution of MeLi in hexane and subsequently HCl in ether, p-S,p-S-11 was obtained with a yield of 68% and an enantiomeric purity of 93%, as confirmed by 1H NMR spectroscopy of the carboxylate complex R,R,p-S,p-S-14.
The enantiomerically pure Zr complex p-S,p-S-17 was synthesized through several stages, beginning with racemic 13, which underwent a reaction with dilithium binaphtholate, followed by treatment with aluminum powder in dichloromethane (Scheme 7), resulting in the formation of complex S,p-S,p-S-16. Further alkylation of S,p-S,p-S-16 with MeLi yielded the dimethyl complex p-S,p-S-17 with a yield of 63% and an enantiomeric purity of 98%ee [50]. This approach also led to the synthesis of p-S,p-S-18 with a yield of 14% [51].
The resolution of racemic ansa-cyclopentadienyl Ti (20a) and Zr (21a) complexes containing biphenyl bridging groups was successfully achieved using R-binaphtholate, while acetylmandelic acid was successfully applied only to ansa-L2TiCl2 (Scheme 8) [52,53]. This process resulted in the acquisition of enantiomerically pure complexes R,R,R-20b and R,R,R-21b. The reaction of titanocene 20a with R-acetylmandelic acid produced a mixture of diastereomers 24 (the yield was determined by 1H NMR of the reaction mixture), which could not be separated via column chromatography.
Complexes rac-11 (Ti) and rac-13 (Zr) were derivatized using a mixture of R-binaphthol and p-aminobenzoic acid in toluene under reflux with an excess of triethylamine (Scheme 9) [54,55]. This reaction yielded binaphtholate complexes R,p-R,p-R-12 (34%) and R,p-R,p-R-16 (37%), along with aminobenzoic acid derivatives p-S,p-S-25 and p-S,p-S-26, which were separated based on their differing solubilities in non-polar solvents. The titanocene binaphtholates were converted into enantiomerically pure dichloride complexes p-R,p-R-11 (97%) and p-S,p-S-11 (33%) through treatment with hydrochloric acid in dichloromethane, while the Zr p-aminobenzoate was transformed into biphenolate S,p-S,p-S-27 (27%, 97%ee) via reaction with 2,2′-biphenol.
Bidentate phenolates of metallocenes rac-13 (Zr) and rac-28 (Hf) (Scheme 10) were quantitatively resolved into enantiomers using HPLC on cellulose (3,5-dimethylphenylcarbamate) as the chiral stationary phase, with a hexane/ethanol (9:1 v/v) mixture serving as the eluent [56]. The specific rotation exhibited a positive value for the first peak (S) and a negative value for the second (R). However, the dichlorides 13a and 13b were not separated into enantiomers under these conditions. The method was further extended by employing new commercially available bidentate ligands (c–h) for the derivatization of rac-13 [57]. In all cases, the (+)-isomers were eluted first, with the highest separation factor observed for complex rac-13d. Optically active biphenolate hafnocenes R,p-R,p-R-28d25435 = −1100°) and S,p-S,p-S-28d25435 = +1100°) were also isolated.
Enantiomerically pure triflate derivatives of ansa-metallocenes Ti (28) and Zr (30) were synthesized starting from titanocene p-S,p-S-11 and zirconocene binaphtholate S,p-S,p-S-16, respectively, as depicted in Scheme 11 [58,59,60]. Additionally, the cationic Zr compound p-S,p-S-29 was obtained.

2.2. Path B: Stereoselective Synthesis of ansa-Metallocenes via Metalation of Chiral η5-Ligand Precursors

Optically active C2- and C1-symmetric ansa-complexes can be synthesized from enantiomerically pure η5-ligand precursors (path B). This method typically results in high diastereo- and enantioselectivity in the synthesis of metallocenes.
For example, optically active C2-symmetric titanium (p-S,p-S-34) and zirconium (p-S,p-S-35) complexes were synthesized via the metalation of ligand S-33, which features a biphenyl bridge, achieving yields of 46% and 20%, respectively (Scheme 12) [61]. The ligand was obtained by separating the dicarboxylic acid R,S-31 into enantiomers R-32 and S-32 using (+)-brucine. The reaction of enantiomer S-32 with a dimagnesium Grignard reagent, followed by protonolysis, yielded enantiomerically pure ligand S-33 with a 60% yield.
A similar approach was employed for the synthesis of a C2-symmetric optically active titanocene 38, which contains a homotopic bis-tetrahydroindenyl ligand. This complex was synthesized from enantiomerically pure Br-substituted binaphthalene S-(+)-36, yielding only one stereoisomer of complex 38 with a yield of 68% (Scheme 13) [62]. The configuration of the binaphthyl fragment controls the structure of the complex, thereby excluding the formation of diastereomeric pairs during transmetalation.
Halterman and co-workers proposed a method for the synthesis of enantiomerically enriched complexes 41, 42, and 44 (Scheme 14) based on optically active ligands obtained from Br-substituted binaphthyl structures 39 [62,63]. The ligand R-(+)-39 was converted into (R)-(+)-2,2′-bis(1-indenylmethyl)-1,1′-binaphthyl (40), which upon coordination with titanium (41) and zirconium (42) formed only one C2-symmetric stereoisomer (Scheme 14). This method was later adapted for substituted indene [64]. For example, the reaction of 4,7-dimethylindenyllithium with S-(−)-39 produced the ligand S-(−)-43 with a 69% yield. The subsequent reaction of the anion of ligand 43 with TiCl3 produced complex 44 with a 24% yield. The developed method provided hard-to-reach enantiomerically pure bis-indenyl complexes; in this process, the substitution of Ti with Zr or introducing methyl groups into the indenyl ligand significantly reduced the yield of the target metallocenes.
Enantiomerically pure bis-indenyl ligand S,S-46 was synthesized from the mesylate of (R,R)-2,3-butanediol 45 (Scheme 15) [65]. Deprotonation of S,S-46 followed by transmetalation to a transition metal and hydrogenation resulted in a mixture of diastereomeric complexes, with the content of isomer C increasing when Ti was replaced with Zr. The optically active Ti complexes 47 were subjected to photolysis, after which the mixture reacted with R-(+)-binaphthol and was separated on silanized silica gel. This process yielded enantiomerically pure titanocenes p-R,p-R-47A (45%) and p-R,p-S-47C (14%). For zirconium, the yield of the target metallocenes was significantly lower, at 9%, for the isomers p-R,p-R-48A and p-S,p-S-48C, with a ratio of 5.9:1.
The reaction of the bis(methanesulfonate) ester of 2,4-diisopropyl-1,4-cyclohexanediol (50) with IndLi in THF produced ligand (+)-(1R,2R,4R,5R)-51, which upon metalation with BuLi and TiCl3, followed by oxidative workup (HCl, air, chloroform), selectively yielded enantiomerically pure titanocene 52 with a high yield (Scheme 16) [66]. The corresponding bis(indenyl)zirconium dichloride was not obtained from this route. Complex 52 was hydrogenated over PtO2 to give 1R,2R,4R,5R-53 in good yield.
High stereoselectivity in the synthesis of titanocene complex p-S,p-S-58 was achieved when the optically active ligand (1R,2R)-trans-1,2-bis-(1-indenylmethyl)cyclopentane 55, obtained from di-(1R,2S,5R)-menthyl-(1S,2S)-trans-1,2-cyclopentanedicarboxyl 54, was reacted with BuLi and TiCl4·2THF (Scheme 17) [67]. Subsequent hydrogenation of p-S,p-S-58 yielded p-S,p-S-59.
Using the less conformationally rigid C4-bridged (4S,5S)-trans-4,5-bis(1H-inden-3-ylmethyl)-2,2-dimethyl-1,3-dioxolane ligand 60, obtained from enantiomerically pure tosylate 59 as a chiral substrate, resulted in a loss of racemo-selectivity and the formation of optically active meso-isomers of complexes 62ac and 63ac (Scheme 18) [68].
The formation of the meso-isomer (66) was also observed in the case of (−)-dimethyl-bis-[3-(neomenthyl)cyclopentadienyl]silane 65, derived from neomenthol 1S,2S,5R-64. In the 1H NMR spectra of the reaction mixture, signals for three possible isomers were observed with a ratio of 5:2:3. However, after recrystallization from pentane, only one optically active metallocene meso-66 was isolated with a yield of up to 34% (Scheme 19) [69].
(+)-(1S,2R,5S)-Menthol (1S,2R,5S-64) was introduced as a chiral element in the bridging group of the bis-indenyl ligand, leading to the synthesis of enantiomerically pure di[(1S,2R,5S)-menthoxy]silylene-bis[η5-1-(R,R)-(+)-indenyl] zirconium dichloride (p-R,p-R-68) with a yield of up to 19% (Scheme 20) [70].
Various isomers of ethylene-bridged cyclopentadienyl titanium complexes 7375 were synthesized from the sterically hindered enantiomerically pure (−)-8,10-dimethyltricyclo[5.2.2.02,6]-2,5-undecadiene 69 (Scheme 21) [71]. It was observed that similar complexes with silicon or methylene bridges could not be synthesized. The metalation stage of ligand 70 occurred with varying selectivity, as this precursor represented a difficult-to-separate mixture of isomers differing by the position of the bridge. Consequently, complexes 7375 were obtained with yields of up to 84%, with a ratio of 73/74/75 = 71:7:22.
To restrict the conformational mobility of the cyclohexane fragment of the tetrahydroindenyl ligand and to stabilize the exo-, exo-configuration (Scheme 22), ligand 80 based on (R)-myrtenal 76 was synthesized. Transmetalation of the ligand with Zr(NMe2)4 produced a single C2-symmetric stereoisomer of complex 81 with an exo-, exo-configuration of the cyclohexyl rings, achieving a yield of 35% [72].
The reaction of optically active bis-amide R,R-82 with cyclohexenylmagnesium bromide lithium chloride yielded the corresponding bis-enone 83 without loss of enantiomeric purity (>99.9%ee) (Scheme 23) [73]. Through the stages of methylenation of 83 and bis-dibromocyclopropanation of compound 84, the ligand precursor 85 was obtained. The rearrangement into a chiral C2-symmetric ligand was performed using MeLi, followed by deprotonation with sodium bis(trimethylsilyl)amide and reaction with TiCl3·3THF without isolation. Three isomers—(S,S,R,R)-47B, the pseudoenantiomeric complex (R,R,R,R)-47A, and the pseudomeso-complex (R,S,R,R)-47C—were isolated in a ratio of B:A:C = 3.6:1:2.8. This method allows for the rapid synthesis of bis-tetrahydroindenyl complexes with good rac-selectivity from alkenyl Grignard reagents and bis-amides.
Another strategy for obtaining optically active ansa-metallocenes via Path B involves the synthesis of complexes with nonequivalent ligands (C1-symmetric complexes). In this scenario, the chirality of the complex is introduced by incorporating a substituent with stereogenic centers into one of the η5-ligands.
For example, a method for synthesizing optically active ansa-metallocenes 8890a,b from Si-bridged bis-cyclopentadienyl ligands 86a and 86b, where one of the Cp fragments contains an enantiomerically pure menthyl or neomenthyl substituent, was developed (Scheme 24) [74,75]. The reaction with TiCl3 yielded a mixture of diastereomers R- and S-88a with a ratio of 3:1. The titanium complex R-88a was crystallized from hot hexane, while the S-enantiomer 88a was obtained after UV irradiation of a mixture of complexes in toluene, followed by passing the solution through silylated silica gel and crystallization from hot pentane. Zr (89a,b) and Hf (90b) complexes were synthesized similarly. The content of the R-enantiomers of ansa-metallocenes increased with the growing atomic radius of the transition metal when using the menthyl-substituted ligand 86a. In all cases, replacing the menthyl group with a neomenthyl group led to a loss of diastereoselectivity.
Following a similar approach to bis-cyclopentadienyl complexes, optically active mixed menthyl-substituted cyclopentadienyl-octahydrofluorenyl ansa-complexes 93 and 94 were synthesized (Scheme 25) [76]. For this purpose, ligand 92 was metalated via reaction with Zr(NMe2)4 (route A, Scheme 25) in toluene at 120 °C. Subsequent treatment with Me2NH·HCl produced a mixture of diastereomers of complex 93 with a yield of 82% and a ratio S/R = 85:15. Deprotonation of 92 with butyllithium (route B, Scheme 25) and transmetalation with ZrCl4 yielded stereoisomers of complex 93 with a low yield.
Neomenthyl-substituted ansa-complexes with an isopropylidene bridge (98) were synthesized (Scheme 26) [77]. In the initial stages of synthesis, the menthyl-substituted fulvene 96 reacted with lithium fluorenide to yield precursor 97. The metalation of ligand 97 resulted in a mixture of diastereomeric complexes 98 in a ratio of 60:40. Thus, introducing the neomenthyl group and isopropylidene bridge into the ligand decreased the stereoselectivity of complex synthesis.
High diastereoselectivity was achieved in the synthesis of the enantiomerically pure (>99%ee) indenyl-fluorenyl ansa-zirconocene with a naphthyl bridge p-S-103 (Scheme 27) [78]. The ligand 102 used in this synthesis was obtained by the reaction of (1R)-menthyl-(S)-1-(p-tolylsulfinyl)naphthalene-2-carboxylate (100) with 1-methylfluorenyllithium at 0 °C. The resulting ester reacted with a Grignard dimagnesium reagent to afford the R-enantiomer of the ansa-ligand 102, which upon metalation selectively produced complex p-S-103.
The reaction of R-(−)-epoxystyrene 104 with fluorenyllithium and subsequent transformations through various stages provided enantiomerically pure ansa-ligands S-107 and S-108 with a stereogenic center in the bridge (Scheme 28) [79,80]. Deprotonation of S-107 with ZrCl4 yielded diastereomers R,p-R-109 and R,p-S-109 in a 1:3 ratio. Hydrogenation of 109 with H2/PtO2 produced corresponding R,p-R-110 and R,p-S-110 in a 1:1 ratio. The metalation of bis-fluorenyl ligand 108 afforded Zr and Hf complexes R-111a,b with yields of 28% and 63%, respectively. Octahydrofluorenyl complexes R-112a,b were obtained following the hydrogenation of R-111a,b.
Enantiomerically pure C1-symmetric ansa-complexes S-116ac were synthesized from optically active cyclopentadiene with a branched hydrocarbon substituent (115) (Scheme 29) [81]. To enhance the rigidity of the complex molecule, the ligands were linked by two Me2Si-bridges.
The presence of two different substituents in one of the cyclopentadienyl fragments of ligand 121 led to the formation of two isomers of complexes S-122 in a 1:1 ratio when using ZrCl4 (Scheme 30) [81]. Utilizing Zr(NMe2)4 as a metalation agent significantly increased the content of one of the possible diastereomers of complex S-123.
The literature also provides examples of preparing enantiomerically pure complexes of group IIIB metals [82]. The synthetic approaches to chiral organolanthanide ansa-η5-complexes have been discussed, with most work focusing on the synthesis of C1-symmetric compounds with stereogenic substituents in one of the cyclopentadienyl rings (Table 1) [83,84]. A series of enantiomerically pure La, Nd, Sm, Y, and Lu complexes 127131ac were obtained via metalation of optically active cyclopentadienyl ligands with neomenthyl, menthyl, and phenylmenthyl substituents 124ac (Scheme 31). The study also demonstrated the potential for the epimerization of chloro lanthanide complexes to enhance the yield and isolation of their diastereomeric antipodes. It was shown that the degree of epimerization depends on the nature of the donor solvent, the lanthanoid ion, the structure of the substituent R*, and the temperature (Table 1). The epimerization of chloride complexes 127131ac can be completely inhibited by adding ligands like TMEDA or 12-crown-4. To obtain stable lanthanide complexes effective in asymmetric catalysis, which do not undergo epimerization in solution, complexes 127131ac were converted into amides 132135ac and alkyls 136139ac (Scheme 31).
Chiral C1-symmetric organolanthanide complexes S-140ac (M = Sm, Y, Lu), with ansa-bridged η5-octahydrofluorenyl and cyclopentadienyl ligands substituted with a menthyl group, were synthesized analogously to complexes 127131ac (Scheme 32) [85]. The highest diastereoselectivity, reaching a ratio of 9:1, was attained when the reaction was performed in dimethoxyethane (DME). The reaction of the diastereomerically enriched complexes S-140ac with KN(TMS)2 proceeded to furnish S-141ac in quantitative yield. Remarkably, at ambient temperature, the process preserved the diastereomeric ratio almost entirely, mirroring that of the initial chloro complexes. The resulting amido complexes exhibited notable thermal stability and were isolated via extraction with boiling hexane. A single recrystallization from hot hexane afforded exclusively the diastereomerically pure amido complexes S-141ac.
Bercaw and colleagues [86] synthesized enantiomerically pure C2-symmetric yttrocenes (R,S)-144, characterized by stereogenic centers within the bridging moiety. This was accomplished through the preliminary preparation of enantiomerically pure ansa-ligands 143 via the reaction of R- or S-binaphthol with tert-butyl-substituted bis(cyclopentadienyl)chlorosilane (142) (Scheme 33). Consequently, coordination to yttrium proceeded with complete diastereoselectivity: the ligand derived from R-binaphthol yielded (R,S)-144, whereas the ligand obtained from S-binaphthol directed the reaction exclusively toward the formation of (S,R)-144.

2.3. Path C: Stereoselective Synthesis of ansa-Metallocenes via Enantiomerically Pure σ-Complexes

Enantiomerically pure σ-complexes can serve as effective stereodifferentiating precursors. For example, the synthesis of the R,R-enantiomer of amide complex 148, which achieved a yield of 84%, was carried out starting from R,R-diphenylpropanediamine (147) (Scheme 34) [87]. Subsequent reactions of amide 148 with lithium salts of ansa-ligands containing ethanediyl or Me2Si bridges provided enantiomerically pure amide complexes 149 and 150 with yields of 97% and 95%, respectively. Compound 149 was transformed into the bis-indenyl complex p-S,p-S-151, attaining a yield of 91% and enantioselectivity exceeding 99%ee.
The diastereo- and enantioselective synthesis of the ethylene-bis-tetrahydroindenyl zirconium complex 13 was proposed via the association of ZrCl4 (R- or S-153) with enantiomerically pure sterically hindered S- or R-binaphthol ethers (Scheme 35) [88]. A higher enantioselectivity (95–96%ee) was observed in the reaction of binaphthol dimethyl ether complexes with the ansa-ligand dianion.
The growing interest in enantiomerically pure metallocenes is primarily motivated by their potential application as catalysts in various asymmetric syntheses. Another significant aspect is the investigation of stereoinduction and chiral recognition mechanisms, which facilitate the understanding of the structural features and dynamics of catalytically active centers, as well as the processes of substrate coordination and transformation by metal centers. In this regard, the subsequent chapters of this review will explore the applications of enantiomerically pure group IVB and IIIB transition metal ansa-η5-complexes, along with some mechanistic aspects of their asymmetric action across various types of reactions.

3. Application of Enantiomerically Pure ansa-η5-Complexes in Asymmetric Catalysis

3.1. Carbomagnesation of Unsaturated Compounds

The complexes of group IVB transition metals are widely acknowledged as effective catalysts for the formation of metal–carbon, carbon–carbon, and carbon–hydrogen bonds. The synthesis of Brintzinger’s ansa-titanocene and zirconocene complexes subsequently facilitated the development of highly active and stereoselective catalytic systems for alkene polymerization [3,4,5,6,7,8,9,10,11,89,90,91,92,93,94,95,96,97,98,99,100,101,102]. These complexes are also well known for their capability to catalyze di- and oligomerization reactions, as well as carbometalation and cyclometalation of alkenes and acetylenes [12,13,14,15,16,17,18,19,20,21,24,25,26,27,28,29,30,42,103,104,105,106,107,108]. A pivotal step in these reactions is the carbometalation of the unsaturated substrate, which generates a highly reactive metal–carbon bond. Consequently, the introduction of enantioselective catalysts into these reactions has enabled the development of efficient methodologies for the functionalization of alkenes, dienes, and acetylenes, yielding a wide range of optically active synthons that are in demand in organic synthesis [24,25,26,27,42,107,108,109,110,111,112].
The first report detailing the application of the enantiomerically pure zirconocene complex p-R,p-R-13 as a stereoselective catalyst for the carbomagnesiation of alkenes was presented in 1993 [113]. Ethylmagnesiation of the allylic R- and S-enantiomers of alcohols 154a,c and esters 154b,d,e (Scheme 36) exhibited varying diastereoselectivity contingent upon the type of starting enantiomer. It was demonstrated that high diastereoselectivity could be attained when employing S-enantiomers of allylic alcohols (syn:anti = 92:8) and R-enantiomers of esters (syn:anti = 7:93). The reaction with R-norbornene alcohol 154f proceeded with high regioselectivity, favoring the formation of the 2R,5S-enantiomer of the ethylmagnesiation product 155f.
In further developments of the method, the reaction of alkyl magnesium halides with readily available heterocyclic alkenes, catalyzed by the enantiomerically pure complexes p-R,p-R-13 and R,p-R,p-R-16, was investigated (Scheme 37) [114,115]. Specifically, the reaction of cyclic olefins 156ad with EtMgCl or n-PrMgCl in the presence of p-R,p-R-13 yielded carbomagnesation products 157ag with good yields (40–77%) and excellent enantioselectivity (over 90%ee). In the carbomagnesation reaction of cyclic amines 156b,e,j,i, the carbometalation proceeded with product yields of 54–81% and enantioselectivity exceeding 95%ee.
Utilizing rac-5,6-dihydropyrans 158ai as substrates, carbomagnesation products 159ai were obtained with high enantiomeric purity. Moreover, enantiomerically enriched 158ai were isolated as a result of the kinetic resolution of the racemic mixture of the starting compounds, due to the participation of only the S-enantiomer in the reaction (Scheme 38) [115,116]. Thus, asymmetric catalytic carbomagnesation proves to be an effective method for the kinetic resolution of enantiomers of unsaturated pyrans, which are important synthons for medicinal chemistry [116].
To investigate the kinetic effect of p-R,p-R-13 in the ethylmagnesiation reaction, enantiomerically pure substituted pyrans were employed. It was found that the reaction was most favorable for S-enantiomers of olefins (S-158d,k), which subsequently provided R-enantiomers of 160d, 159k with good yields and enantioselectivity (Scheme 39) [115].
Substituted dihydrofurans 162ad also undergo kinetic resolution in the carbomagnesation reaction [117]. Unlike pyrans, both enantiomers of substituted furans fully participate in the reaction with the Grignard reagent (Scheme 40). In this case, the reaction proceeded with high regio-, enantio- (>98%ee), and diastereoselectivity (95%de) concerning both enantiomers of the substrates, resulting in the formation of different products: the S-isomer yielded only primary alcohol S-162ad, while the R-enantiomer produced the secondary alcohol R,S-163ad.
The application of the binaphtholate complex p-R,p-R-16 provided a high degree of enantiomer separation for seven-membered cycloolefins 164ai, predominantly resulting in the isolation of R-enantiomers 165 (Scheme 41) [115,118]. The use of the biphenol complex p-R,p-R-27 yielded nearly identical results. A decrease in enantioselectivity to 70%ee was observed for the cyclohexyl-substituted substrate 165c.
During the carbomagnesiation reaction catalyzed by R,p-R,p-R-16 or S,p-S,p-S-16, cyclic alkoxy-substituted olefins 166av were also subjected to kinetic resolution, leading to the enrichment of the mixture with S-enantiomers exhibiting purities ranging from 50% to 99%ee (Scheme 42) [119,120,121]. In this case, the highest degree of separation was achieved in reactions involving seven-membered cyclic substrates 166k,l and 166q,r,u,v.
Thus, cyclic olefins in the presence of enantiomerically pure ansa-tetrahydroindenyl Zr complexes and Grignard reagents undergo kinetic resolution, during which one of the enantiomers can participate in an ethylmagnesiation reaction with ring opening. The structure of the starting olefins plays a significant role in these processes. For example, the enantioselectivity of the reaction is influenced by the size of the olefinic cycle (5-, 6-, 7-, or 8-membered) and the structure of the alkyl group in the R-O moiety. It is noted that increased steric hindrance of the substrate, both from the larger cycle and the substituent, leads to a loss of enantioselectivity in the process.
The ethylmagnesiation of inactivated alkenes by MgEt2 in the presence of the rigid ansa-complex R,p-R,p-R-16 proceeded with the formation of products 168ac with low yield and enantioselectivity (Scheme 43) [122,123].
Optically active titanocene complexes p-S,p-S-4b [44] and p-S,p-S-11 [48] were evaluated as catalysts for enantioselective reactions aimed at producing homoallylic alcohols [124,125]. The utilization of various metal complexes in the alkylation of ketones and aldehydes allows for the modulation of stereoselectivity in this process. For example, the reaction of crotyl titanocenes with different substituents in the cyclopentadienyl ring (R = H, Me, i-Pr) with primary, secondary, and tertiary aldehydes was described (Scheme 44) [125]. The products of the reaction, homoallylic alcohols 169, were formed with good to excellent selectivity, consistently favoring the anti-diastereomer. The degree of diastereoselectivity correlates with the size of the substituent in the cyclopentadienyl ring of the Ti complex and in the starting aldehyde. Specifically, for aliphatic and aromatic aldehydes, an increase in diastereoselectivity was observed with greater steric hindrance at the metal center.
The reaction of allyl-(171) and crotyltitanocenes (172), derived from ansa-titanocene p-S,p-S-11, with aldehydes afforded homoallylic alcohols 170ah with excellent yields exceeding 80% and a diastereoselectivity of anti/syn = (2.5–40)/1 (Scheme 45). It was suggested that allyltitanocene additions proceed via six-membered chair- or boat-like transition states, which regulate the reaction stereoselectivity [48,59,125].
The substitution of the tetrahydroindenyl catalyst p-S,p-S-11 with ansa-ethylene-bis-(tert-butyl-cyclopentadienyl)titanium dichloride p-S,p-S-4b resulted in a change in absolute configuration from R- to S- in the reaction products—anti-homoallylic alcohols (Scheme 46) [44]. The diastereoselectivity of the process increased with the size of the substituent in the aldehyde (anti:syn = >99:1). In the reaction with pivaldehyde, the predominance of the anti-R enantiomer 169d was observed. Furthermore, the syn-R-stereoisomer of homoallylic alcohol 169a was obtained with greater enantioselectivity than the anti-isomer.
Titanocenes 52, 58, and 59 reacted with 2-alkyl-1,3-butadienes 174a,b and i-PrMgCl, yielding allylic complexes 175, which were then subjected to interaction with excess CO2 or R’CHO, resulting in the formation of S-enantiomers of carboxylic acids 176a,b or S-threo/erythro-homoallylic alcohols 177, respectively (Scheme 47) [126]. The highest enantiomeric excess of >91%ee was observed in the reaction, catalyzed by complex 52, with a predominance of S-threo isomers of the alcohols. In this case, the yield and enantiomeric excess of erythro- isomers of the alcohols were low—12–44%ee.
Enantiomerically pure titanocene p-S,p-S-11, when employed as a reagent in the Barbier-type carbonyl compound propargylation or allylation, yielded homo-propargylic (179) or allylic (180) alcohols, respectively, demonstrating the fundamental possibility of conducting this reaction in an asymmetric manner (Scheme 48) [127,128,129].

3.2. Carbo- and Cycloalumination of Unsaturated Compounds

The asymmetric Zr-catalyzed enantioselective carboalumination of alkenes was first demonstrated by Negishi in 1995 [130,131,132], exemplified by the asymmetric carboalumination of alkenes using AlR3 (R=Me, Et) in the presence of the conformationally labile complex p-S,p-S-bis-neomenthylindenylzirconium dichloride (p-S,p-S-186), achieving enantioselectivities of up to 85%ee for AlMe3 and 92%ee for AlEt3, respectively.
Studies on the enantioselectivity of ansa-complexes p-S,p-S-13, p-S,p-S-16, p-S,p-S-18, p-S,p-S-19 [51], p-R,p-R-13, p-R,p-R-16 [130], p-S,p-S-151, p-S,p-S-187ac, p-S,p-S-188 [132], and p-R,p-S-189 [133] indicated the dependence of the stereodifferentiating effect on the structure of the substrate and the catalyst (Table 2, Scheme 49).
It has been demonstrated [132] that the carboalumination of styrene and allylbenzene by AlMe3 in the presence of 10–13 mol% of p-S,p-S-151 or p-S,p-S-187ac proceeded with conversions of the starting alkenes in the range of 72–89% over 3–6 h. The highest enantioselectivity of 80%ee was achieved in the reaction with styrene, then the content of the p-S,p-S-188 complex, activated by [Ph3C][B(C6F5)4], was increased to 26 mol%. The methylalumination of allylbenzene, followed by oxidation of the reaction mixture, afforded 2-methyl-3-phenyl-1-propanol with an enantiomeric excess of 25–33%ee. The transition from an ethylene bridge to a dimethylsilylene bridge in the p-S,p-S-151 and p-S,p-S-187a complexes, along with the introduction of methyl groups at the 4,7- or 3- positions of the indene ligands in the p-S,p-S-187b and p-S,p-S-187c complexes, did not lead to significant changes in the enantioselectivity of the reaction.
To elucidate this phenomenon, the mechanism of alkene methylalumination in the presence of enantiomerically pure ansa-zirconocene complexes was proposed (Scheme 50) [132]. According to Scheme 50, the starting ansa-complex in the presence of MAO provides an activated state—the cationic complex S,S-1a.1. Subsequent coordination and insertion of the olefin yield a new alkylzirconium cation S,S-1a.3, whose association with an AlMe3 molecule leads to the heterobimetallic complex S,S-1a.4. The target OAC is formed as a result of alkyl exchange in complex S,S-1a.4 and its dissociation. It is suggested that the higher enantioselectivity of the reaction is due to the coordination stage of the olefin to the activated complex in a “side-on” manner, while lower selectivity is associated with a “head-on” approach. It is noted that, apparently, in complexes p-S,p-S-187b and p-S,p-S-187c, the methyl substituents do not possess sufficient volume to obstruct the “head-on” direction. The increase in enantioselectivity of the methylalumination when utilizing styrene is attributed to the enhanced interaction between the ligand framework of the complex and the phenyl fragment of styrene.
The reaction of allylbenzene or styrene with AlEt3 catalyzed with (+)-[η5:η1-indene dimethylsilyl(α-methylbenzyl)amido]titanium dichloride (p-R,S-189) yielded the cycloalumination product 184 with moderate yields of 40–60% and an enantiomeric excess ranging from 19 to 28%ee [133]. Remarkably high enantioselectivities of 96–99%ee were observed in the reaction of 2,5-dihydrofurans with AlEt3, catalyzed by the complexes p-R,p-R-13 or R,p-R,p-R-16 (Scheme 51) [137].
The methylalumination of terminal alkenes with AlMe3 catalyzed by the binaphtholate complex bearing Si-bridged ligands p-S,p-S-18, activated by MAO, afforded the corresponding products with a yield of 66% and an enantiomeric excess of 65%ee (R) [51]. The dichloride complexes p-S,p-S-13 and p-S,p-S-18 exhibited the highest enantioselectivities in carboalumination, achieving 50–51%ee (R). The enantiomeric purity of cyclic organoaluminum compounds OACs (185) formed in reactions catalyzed by complexes S,p-S-13, p-S,p-S-18, S,p-S,p-S-16, and S,p-S,p-S-19 ranged between 12 and 26%ee (S). It is noteworthy that the fixation of the geometry of the zirconocene complex provided methyl- and ethylalumination products with identical configurations at the β-stereogenic center but with diminished stereoselectivity compared to the conformationally flexible complex p-S,p-S-bis-neomenthylindenylzirconium dichloride (p-S,p-S-186) [135,136].

3.3. Asymmetric Oligo- and Polymerization of Alkenes and Dienes

The potential of chiral metallocenes of group IVB metals as catalysts for the stereoselective synthesis of oligomers and polymers from terminal alkenes has been convincingly demonstrated [99,100]. For example, the polymerization of propylene utilizing a catalytic system based on p-R,p-R-17 and methylaluminoxane (MAO) afforded a mixture comprising hydrogenated isotactic polypropylenes 194 and liquid hydrogenated oligomers 195 (Scheme 52) [138,139].
Infrared and 13C NMR spectroscopic analyses of the oligomer fractions revealed the presence of terminal n-butyl, n-propyl, and isobutyl groups, with a predominance of n-propyl moieties. Notably, the polymers were devoid of double bonds, thus characterizing the process as “hydrooligomerization.” The hydrooligomers exhibited isotactic configurations, with the first asymmetric carbon atom in the growing oligomer chain possessing an S-configuration. Comparison of mass-spectrometry data, as well as 1H and 13C NMR spectra of the obtained low-molecular-weight hydrooligomers (isolated by distillation of 4,6-dimethyldecane and 2,4,6,8-tetramethylnonane), with the spectra of rac-4,6-dimethyldecane and meso- and rac-2,4,6,8-tetramethylnonane, led to the conclusion that absolute configurations of the stereogenic centers at positions 4 and 6 differ. As a result, 4S,6R-4,6-dimethyldecane and 4R,6S-2,4,6-trimethylnonane were identified.
The polymerization of 1-pentene in the presence of R,p-R,p-R-16 or p-R,p-R-17 and MAO, followed by treatment with D2, yielded optically active deuterated pentanes of general formula C5H12−xDx and deuterated oligomers (Scheme 53) [140]. Fractional distillation and successive solvent extractions (methanol, acetone, ethyl acetate) facilitated the isolation of these oligomers. It was proposed that the elongation of the oligomer chain enhances the stereoselectivity of the process, as evidenced by the superior optical purity of the dimers relative to the deuteropentanes. This observation aligns with the hypothesis that stereoselectivity in α-olefin polymerization is governed both by the chirality of the catalytic center and the spatial orientation of the last monomeric unit of the growing chain towards the metal atom [138].
The enantiomerically pure acetylmandelate complex R,R,p-S,p-S-15, combined with MAO, was evaluated for propene and 1-butene oligomerization (Scheme 54) [141]. Under increased catalyst concentration and elevated monomer content, isotactic oligomers bearing terminal double bonds were obtained, indicative of an accelerated chain termination rate relative to polymer chain propagation.
Initially, methylaluminoxane mediates the substitution of the mandelate ligand with a methyl group in the complex, generating a Zr–Me bond that facilitates substrate incorporation (Scheme 54). Subsequent olefin coordination promotes chain elongation. β-Hydride transfer to the metal center terminates the chain, yielding a complex bearing a free Zr–H bond capable of coordinating another alkene to regenerate the metal alkyl species. Gas chromatographic analysis revealed the presence of oligomers with degrees of polymerization (n) ranging from 2 to 7. Under analogous conditions, propylene predominantly formed pentamers, whereas 1-butene yielded trimers and tetramers. The higher oligomers exhibited isomeric mixtures primarily composed of configurational isomers. For example, the tetramer fraction contained a diastereomeric mixture of (R,R)/(S,S)- and (R,S)/(S,R)-2,4,6-trimethyldecene in a 5:4 ratio, while the pentamer fraction comprised four pairs of enantiomers of 2,4,6,8-tetraethyl-1-dodecene. Gas chromatographic separation of the enantiomers of the 1-butene trimer mixture was achieved via modified cyclodextrin chromatography [142,143], with an enantiomeric purity of 27%ee. Consistent with the hydrooligomerization of propylene, the reaction favored the formation of R-enantiomers. The stereoselectivity of the second alkene insertion step was notably enhanced, as the stereogenic centers in the growing oligomer chain enhanced the stereodifferentiating effect (growing chain control).
The catalytic system p-S,p-S-13-MAO exhibited a high degree of isotacticity in the polymerization of propylene, 1-butene, 1-pentene, and 1-hexene [144]. The elongation of the alkyl substituent exerted minimal influence on isotacticity, which remained between 95 and 99%mmmm at temperatures below 20 °C. The optical rotation of higher oligomers diminished with increasing reaction temperature, whereas the enantiomeric excess of the oligomers increased from 10% to 90% as the temperature decreased from 60 °C to 0 °C.
Comparative studies of propene polymerization stereoselectivity in the presence of R,R,p-S,p-S-15 and a mixture of diastereomers R/S-15 revealed that the exchange of polymer chains between metal centers of opposing chirality significantly undermines stereochemical control [145]. In most experiments, the precursor was activated with an optimal quantity of AlBui3 and stoichiometric amounts of [Ph3C][B(C6F5)4]. Both racemic and enantiopure catalysts displayed comparable activities; however, the racemic catalyst exhibited diminished stereo- and regioselectivity. The stereoselectivity of R/S-15 improved markedly upon lowering its concentration, reducing the reaction temperature, or immobilization on silica supports. It was concluded that the racemate forms two catalytically active centers with different configurations capable of exchanging polymer alkyl chains, thereby reducing stereoselectivity and increasing regioerrors.
Furthermore, the catalytic activity and stereoselectivity of chiral homogeneous and heterogeneous catalytic systems were compared [146]. The optically active heterogeneous catalyst, prepared by depositing TiCl3 on MgCl2 in the presence of di[(S)-2-methylbutyl]phthalate as an electron donor, exhibited moderate activity in racemic 4-methyl-1-hexene (4MH) polymerization (2.6·103 g polymer/(mol Ti·mol 4MH·h)). The (4S)-4-methylhexene enantiomer was incorporated more favorably than the (R)-enantiomer, with a polymerization rate ratio R(S)/R(R) of 1.02 and a polymer rotation angle [α]D25 of +4.8°. Consequently, the residual monomer was enriched in the R-enantiomer. In contrast, the homogeneous catalytic system R,R,p-S,p-S-15-MAO exhibited a higher rate ratio of 1.4 and a polymer rotation angle of +17 to +19°, with an activity of 112·103 g polymer/(mol Zr·mol 4MH·h). The stereoregularity index (SI) was lower in heterogeneous catalysis (maximum 85%) than in the homogeneous system (maximum 97%). Fractionation of polymer in various solvents (acetone, ethyl acetate, diethyl ether, and cyclohexane) yielded fractions with distinct rotation angles. For instance, fractions obtained in the system R,R,p-S,p-S-15-MAO displayed [α]D25 values of –7.8° (acetone), +22° (ethyl acetate), and +31° (cyclohexane), whereas the TiCl3-MgCl2-Et3Al-di[(S)-2-methylbutyl]phthalate system yielded positive rotations of +5.4°, +10°, +15°, and +9.9° in the respective solvents. This disparity, including the inversion of rotation sign, was attributed to stereoerrors, i.e., the incorporation of both S- and R-enantiomers under specific conformations of the catalytically active centers in homogeneous systems—a phenomenon more prevalent at elevated temperatures due to increased conformational flexibility. Conversely, heterogeneous catalysis features fixed geometries of active centers, albeit with greater diversity in Lewis acidity and steric environments.
Cyclopolymerization of 1,5-hexadiene catalyzed by achiral precatalysts Cp2MX2 (M = Ti, Zr; X = Cl, Me) and MAO at ambient temperature afforded poly(methylene-1,3-cyclopentane) (PMCP) 196, which can exhibit four possible microstructural variants (Scheme 55) [147]. Among these, only the trans-isotactic microstructure is chiral. According to 13C NMR data, approximately 80% of PMCP rings adopted the trans-configuration. Lowering the polymerization temperature enhanced trans-selectivity, reaching 90% trans rings at –78 °C. In contrast, sterically hindered complexes (CpMe5)2MX2 (Cp* = pentamethylcyclopentadienyl) predominantly yielded cis-PMCP, with 86% cis content at –25 °C. Cyclooligomerization of 1,5-hexadiene catalyzed by chiral catalysts rac-13 and rac-151 afforded PMCP with 70% trans content (Scheme 55). Here, the cis/trans stereochemistry was less sensitive to polymerization temperature compared to achiral metallocenes. Enantiomerically pure precatalysts R,p-R,p-R-16 and S,p-S,p-S-16 produced optically active trans-isotactic PMCP 196 with molar optical rotations [Φ]40528 = +51.0° and [Φ]40528 = −51.2°, respectively.
Enantiomerically enriched styrene hydrooligomers were synthesized via catalysis by complex p-R,p-R-16, activated with methylaluminoxane under a hydrogen atmosphere [148]. The oligomers included R-1,3-diphenylbutane (197) with an enantiomeric excess of 77%ee (Scheme 56). Subsequent functionalization without loss of chirality underscores the potential of this catalytic system for preparing optically active synthons for organic synthesis.
Polymerization of racemic α-olefins bearing bulky substituents 199af catalyzed by enantiomerically pure S-116ad in the presence of MAO achieved kinetic resolution, yielding the R-enantiomers of 199a–f with enantiomeric excesses of 7–59%ee and chiral, highly isotactic polymers 200af enriched in the S-enantiomer (Scheme 57, Table 3) [81,86,149,150,151]. The selectivity factor s (rate of the fast-reacting enantiomer)/(rate of the slow-reacting enantiomer) for most olefins ranged from approximately 1.1 to 3.2, indicating partial kinetic resolution; however, polymerization of 3,4-dimethyl-1-pentene (199c) exhibited pronounced resolution (s > 15) [150,151]. To rationalize isotactic polymer formation, the authors proposed epimerization of the catalytically active center, whereby the polymer chain alternates sides relative to the metallocene cation after each monomer insertion. A high rate of site epimerization relative to olefin coordination ensures that the olefin encounters a consistent chiral pocket, accounting for the high stereoselectivity. The stereogenic centers in the growing polymer chain may further influence the stereoselectivity of the process.
In the presence of catalysts 116a and 116d, the stereoselectivity of propene polymerization was markedly influenced by substrate concentration ([mmmm] = 2.8–62.9%, [rrrr] = 6.8–49.6%) (Scheme 58) [150]. Elevated isotacticity was attained at diminished monomer concentrations ([mmmm] = 62%), whereas conducting the reaction in neat propylene yielded a syndiotactic polymer ([rrrr] = 44–49%). At intermediate concentrations, catalyst 116d favored a higher proportion of isotactic polymer relative to 116a. Moreover, 116d exhibited enhanced selectivity during the kinetic resolution of 3-substituted racemic α-olefins.
To rationalize these observations, it is noted that the polymerization of racemic α-olefins likely proceeds via the involvement of a pair of diastereomeric pairs at each catalytic site (even more if chain end control is taken into account), instead of involving just a pair of diastereomers as observed in simple α-olefin polymerizations [150]. Consequently, the selectivity in racemic polymerization is governed by matching and mismatching pairs in a double diastereoselective manner. The proposed enantiomorphic site control model (Scheme 58) incorporates a double diastereoselective transition state—for instance, the (S)-metallocene/si-olefin face/(S)-olefin diastereomer is favored over the (S)/si/(R) diastereomer. Nonetheless, the scenario is more complex, especially when considering potential insertions from both faces of the zirconocene, chain end control, and/or poor enantiofacial selectivity. Additionally, the rate of anion reorganization during racemic α-olefin polymerization may also contribute significantly.
Subsequently, a one-pot method for the synthesis of the deoxypropionate scaffold was developed based on the stereocontrolled oligomerization of propylene (Scheme 59) [152]. It was demonstrated that the reaction of propylene with ZnEt2 in the presence of optically active zirconocene p-S,p-S-13 and MAO afforded oligomers. Under identical conditions, the employment of AlMe3 failed to yield the target products. Following oxidation and hydrolysis, branched primary alcohols 201 were obtained with high enantiomeric purity (99%ee), albeit with modest yields due to the formation of polymerization side products. The CS-symmetric zirconocene 202 catalyzed syndiospecific oligomerization under analogous conditions, delivering rel-(2R,4S,6R,8S)-2,4,6,8-tetramethylundecanol-1 [153], a constituent of Antitrogus parvulus (Scheme 59).
Oligomerization of 1-hexene catalyzed by enantiomerically pure Zr ansa-complexes p-S,p-S-13, p-S,p-S-18, S,p-S,p-S-16, and S,p-S,p-S-19 in the presence of AlR3 (R = Me, Et) and MAO yielded optically active, functionally substituted oligomers 204a,b with degrees of oligomerization up to six and yields ranging from 10 to 54% (Scheme 60) [51].
In developing this methodology, stereoselective oligomerization of 1-hexene was achieved by employing p-R,p-R-13, AlMe3, and activators such as MMAO-12 or [Ph3C][B(C6F5)4]. This approach yielded organometallic intermediates 203a bearing an initial methyl group, which upon oxidation and hydrolysis furnished alcohols 204a (Scheme 60) [154]. In the presence of MMAO-12, the oligomer fraction (204a, n = 2–6) constituted 63–81%, whereas the use of [Ph3C][B(C6F5)4] as an activator enhanced oligomer yields to 94–98%. The functionally substituted oligomers 204a exhibited enantiomeric enrichment and diastereomeric purity (99%dr). An increase in oligomer chain length correlated with augmented enantioselectivity, irrespective of activator type. Nevertheless, the nature of the activator significantly influenced enantioselectivity. Specifically, reactions with AlMe3 and MMAO-12 proceeded with moderate enantioselectivity—60–64%ee for oligomers containing two monomeric units, and 78–84%ee for oligomers with n = 3,4—reaching 93%ee for n = 5, 6. Substitution of MMAO-12 with [Ph3C][B(C6F5)4] diminished enantioselectivity to an average of 10–20%ee. Therefore, stereoselectivity in these catalytic systems is governed by both the metal center and the growing polymer chain, which is consistent with the Corradini model [155]. The experimental data further underscore the pivotal role of counterion effects during alkene insertion into catalytically active centers of the type [L2Zr+Alk…SLA] (SLA—strong Lewis acid).
Despite substantial progress in stereoselective olefin polymerization, enantioselective oligomerization and polymerization of alkenes remain comparatively underexplored. Significant advancements have been realized employing enantiomerically pure ansa-zirconocenes. Mechanistic descriptions frequently invoke multiple conformations of catalytically active centers during both alkene coordination and chain propagation, implying that reaction stereoselectivity is largely dictated by the ligand environment surrounding the transition metal and the growing polymer chain. Recently, an increasing imperative has emerged to scrutinize the influence of activator structure on the organization of catalytically active centers, and consequently, on the stereochemical outcome of these transformations.

3.4. The Diels–Alder Cycloaddition

Enantiomerically pure zirconium bis-alkoxy or triflate ansa-complexes serve as catalysts for Diels–Alder cycloadditions between cyclopentadiene and various dienophiles, exemplified by 205ag (Scheme 61) [58]. Reactions involving dienophiles bearing ethenyl substituents (205a,c) preferentially yield the endo-product. The endo/exo product ratio depends on the substituent’s nature at the double bond. In all cases, enantioselectivity did not exceed 50%ee. The authors attribute the low enantioselectivity of the reactions to both the steric factors of the substrate (endo- and exoselectivity) and the steric environment of the transition metal center.
The cycloaddition of enones 205eg with cyclopentadiene catalyzed by chiral triflate complexes Ti (p-S,p-S-28) and Zr (p-S,p-S-30) predominantly afforded R-endo isomers 206eg, with enantiomeric excess heavily influenced by solvent polarity (Scheme 61) [60]. In dichloromethane, enantioselectivity was below 69%ee (R), whereas in nitromethane or 2-nitropropane, stereoselectivity reached 90–95%ee (R, endo/exo = (6–15)/1) [59]. The degree of asymmetric induction in this reaction was ascribed to the conformations of the intermediates—coordination products of the ansa-complex and the starting oxazolidinone substrate 205eg (Scheme 61) [60]. Spectroscopic (EXSY, NOESY) and kinetic analyses revealed the formation of two isomeric complexes upon substrate–catalyst binding. The relative reactivities of these intermediates determine the diastereo- and enantioselectivity of the reaction. Thus, solvent-dependent stereoselectivity likely arises from the differential stability of complexes 207A and 207B. It is plausible that the minor isomer 207B reacts with cyclopentadiene more rapidly and selectively, favoring the formation of endo enantiomers 206eg.

3.5. Reactions of Imine ansa-Zirconocenes—Zirconaaziridines

Metal–carbon bonds exhibit pronounced reactivity toward organic substrates and serve as pivotal intermediates in many catalytic processes. Since 1991, asymmetric catalysis employing imine complexes of ansa-zirconocenes—termed zirconaaziridines—has been reported. The synthesis of zirconaaziridines is typically conducted in situ (Scheme 62). For example, the reaction of the dimethyl complex p-S,p-S-17 with trifluoromethanesulfonic acid and lithium amide provided zirconocene methyl amide 209. Subsequent methane elimination at 80 °C yielded zirconaaziridine 210 [50,156].
A diverse array of heteroatom-containing compounds has been synthesized via reactions of zirconaaziridines with unsaturated substrates. Notably, asymmetric induction in the reaction of amines with alkynes, alkenes, or aldehydes in the presence of zirconaaziridines 210 has been demonstrated, yielding allylic amines 212ai and phenylalkyl amines 212j,k with respectable yields (3872%) and high enantioselectivity (Scheme 63) [50].
Reactions of complexes 214 with various alkenes furnished metallocycles 215ad in yields ranging from 67% to 94% (Scheme 64) [157]. Diastereoselectivity in reactions with aliphatic alkenes was enhanced using the bis-indenyl catalyst rac-213, achieving up to 83%de, whereas the tetrahydroindenyl analog rac-17 exhibited diminished stereoselectivity (66%de). Employing styrene or vinyltrimethylsilane as substrates altered regioselectivity and augmented diastereoselectivity to 98%de, independent of catalyst structure. Reaction of 1-alkenes with excess picoline in the presence of optically active dimethylzirconocene p-S,p-S-17 and [HNBu3][BPh4] afforded substituted pyridine 216 with 58%ee (R) (Scheme 64).
Reactions of alkyl- or aryl-substituted zirconaziridine 210 with ethylene carbonate or isocyanates produced amino acid esters 218ah with enantiomeric purities up to 98%ee, and phenylglycinamides 220ad with 80-99%ee, respectively (Scheme 65) [156]. In these transformations, the ansa-complex functions both as catalyst and substrate, necessitating stoichiometric amounts of the initial complex.
In discussing the plausible mechanism of stereoinduction, a hypothesis was proposed correlating enantioselectivity with the equilibrium between epimers of zirconaziridine 210. This ratio can be modulated by varying reagent quantities (ethylene carbonate or isocyanate), reaction parameters (temperature, solvent), or substrate substituents (Scheme 66).
Furthermore, it was demonstrated that the reaction of enantiomerically enriched imido-complex p-S,p-S-222 with allenes proceeds via kinetic resolution, yielding a single R-enantiomer of allene 225 with 50–61% yield and 7898%ee (Scheme 67) [158]. The more reactive allene enantiomer was subsequently regenerated from the metallacycle by treatment with the parent allene, 1,2-propadiene. Thus, the kinetically favored S-allene enantiomer was recovered from metallacycle R,p-S,p-S-223 via reaction with 1,2-propadiene.

3.6. Cyclization Reactions

Complexes of Group IVB metals, similar to cobalt complexes, facilitate the Pauson–Khand cyclo-condensation of enynes with carbon monoxide, yielding cyclopentenones. The first examples of this reaction being carried out in an asymmetric manner are presented in Refs. [159,160] (Scheme 68). It was found that p-S,p-S-226 serves as a selective catalyst for transforming various 1,6-enynes 227ah in the presence of CO into the corresponding bicyclic cyclopentenones 228ah, achieving yields of 70–94% and enantioselectivity of 72–96%ee.
Subsequently, a method for the asymmetric cyclocarbonylation of nitrogen-containing enynes was developed (Scheme 69) [161]. Studies of the effects of substituents at the nitrogen atom and catalyst concentration on the enantioselectivity of this type of cyclization revealed that enynes 229aj undergo cyclization in the presence of CO and catalytic amounts of p-S,p-S-226, producing azabicyclo[3.3.0]oct-5-en-7-ones 230aj with yields exceeding 62% and enantioselectivity up to 95%ee. It was demonstrated that reactions involving enynes with N-octyl, N-allyl, or N-benzyl groups exhibited high enantioselectivity. In contrast, the presence of an anilino, amide, or sulfonamide moiety in the starting molecules resulted in products with lower ee values. When nitrogen was part of a carbamate group, moderate enantioselectivity was observed. It is suggested that substrates with a relatively electron-rich nitrogen center cyclize to enones with higher enantioselectivity compared to those with an electron-deficient nitrogen atom.
The intramolecular asymmetric reductive coupling of ketones with nitriles, catalyzed by the optically active titanocene p-R,p-R-11, was developed. This method yields substituted α-hydroxyketones (acyloins)—structural fragments of numerous natural compounds. Cyclic hydroxy-substituted cyclopentanones 232ao were synthesized with commendable yields ranging from 47% to 94%, with enantioselectivity exceeding 65%ee (Scheme 70) [162].

3.7. Asymmetric Hydrosilylation of Ketones

The asymmetric hydrosilylation [163] of ketones 233av using polymeric methylhydrosiloxane Me3SiO[MeSi(H)O]nSiMe3 in the presence of chiral titanocene binaphtholates R,p-R,p-R-12 proceeded with high enantioselectivity (82–96%ee) (Scheme 71) [164]. Among these, ketones lacking a double bond at the α-position (234s,u) exhibited poor enantioselectivity. Furthermore, it was demonstrated that employing substrates with bulky alkyl groups in the alkylaryl ketone structure enhanced enantiomeric excess. Generally, conducting the reaction with branched substrates necessitated an additional catalyst amount (up to 10 mol%) and extended reaction times.
Hydrosilylation of ketones was performed using silanes (EtO)3SiH, Me(EtO)2SiH, [MeSi(H)O]4, Me3SiO[MeSi(H)O]nSiMe3, and MeSiH3, alongside catalysts prepared by alkylation of S,p-S,p-S-12 with MeLi or n-BuLi (Scheme 72) [165]. The enantiomeric purity of the acetophenone hydrosilylation product 234a, obtained in the presence of the S,p-S,p-S-12-n-BuLi system, was significantly higher (approximately 99%ee) than that achieved with the MeLi-alkylated catalyst (approximately 40–50%ee). The hydride complex [(EBTHI)TiH]2, synthesized in the reaction of S,p-S,p-S-12 with MeLi and PhMeSiH2, exhibited essentially the same enantioselectivity as the MeLi-based catalyst. Since the titanocene hydride demonstrated low enantioselectivity in the acetophenone hydrosilylation, the authors proposed two potential reaction mechanisms involving TiIII and TiIV species (Scheme 72). In these mechanisms, key intermediates were suggested to be species with Ti-Si bonds, whose coordination with the substrates determines both the reaction rate and stereoselectivity. The observed differences in enantiomeric excess with MeLi and n-BuLi were attributed to the rapid decomposition of L2TiBu2 alongside Ti reduction, in contrast to the L2TiMe2 complexes, as well as the direct involvement of Ti alkyl complexes as intermediates.
Optically active titanocene difluoride p-R,p-R-235 was found to be an effective catalyst for the hydrosilylation of ketones by PhSiH3 in the presence of alcohol additives (Scheme 73) [166,167]. It has been shown that primary alcohols were more effective as additives than secondary alcohols or primary amines. For example, the gradual addition of methanol to the reaction mixture resulted in an increased conversion of the starting ketone (99%) and the enantiomeric purity of the synthesized secondary alcohols (99%ee) [166].
In the proposed mechanism illustrated in Scheme 74, the stereodifferentiating and rate-limiting steps of the reaction are distinct. The smaller the size of the alkyl group in the alcohol, the more rapid the σ-bond metathesis occurs in the rate-limiting step. Consequently, the added alcohol does not participate in the stereodifferentiating step; rather, it accelerates the reaction, thereby diminishing the probability of non-selective pathways during the hydrosilylation of ketones.
The catalytic activity of η5-titanium complexes S-88a, R-88a, p-S,p-S-38, and 1R,2R,4R,5R-53 has been studied in the reduction reactions of aryl-substituted ketones using phenylsilane (Scheme 75) [74,168,169]. These catalysts exhibited varying activity (0–100% substrate conversion) and relatively low enantioselectivity (with a maximum of 82%ee in the case of acetophenone). Consequently, their practical significance is considerably lower compared to, for example, complexes with ethylene bridges [164,166]. In this case, the steric bulk of the complexes increases the enantioselectivity of the process; however, it significantly reduces their catalytic activity.

3.8. Asymmetric Hydrogenation of N-Substituted Compounds

With the growing practical significance of enantiomerically pure nitrogen-containing compounds in the pharmaceutical and agrochemical industries, there is a pressing need to develop synthetic methods that exhibit high chemo-, regio-, and stereoselectivity. Asymmetric catalytic reduction of imines stands out as an effective method for the synthesis of chiral amines. Enantiomerically pure titanocene complexes R,p-R,p-R-12 and R,R,R-22 showed high activity and enantioselectivity in imine hydrogenation (Scheme 76) [53,170,171,172,173]. The catalytically active center in this reaction is presumably a titanocene (III) hydride, generated in situ through the interaction of titanocene complexes with an excess of triphenylsilane. Subsequently, phenylsilane was replaced with gaseous hydrogen, which proved to be a more effective stoichiometric reductant. The hydrogenation of imines 236av, catalyzed by R,p-R,p-R-12, was carried out at hydrogen pressures of 80–2000 psi for durations ranging from 8 to 48 h [170,171,172,173] (Scheme 76). As a result, both cyclic and acyclic amines 237av were formed with varying enantiomeric purity. High enantioselectivity (97–99%ee) was achieved with cyclic substrates, particularly ketimines, whereas acyclic compounds were reduced with moderate enantioselectivity (59–93%ee). The observed decrease in enantiomeric purity of the products is likely attributable to the fact that acyclic imines exist as mixtures of anti- and syn-isomers, which can interconvert during the reaction. Utilizing the same approach, 1-phenylpyrroline (236w) and N-(R-methylbenzylidene)benzylamine (236x) were reduced in the presence of the titanium complex R,R,R-22 [53]. Under conditions of 150 bar H2 and 80 °C for 12 h, the corresponding secondary amines 237w and 237x were obtained with yields of 96% and enantioselectivities of 76–98%ee.
The binaphtholate Ti complex S,p-S,p-S-12 can be effectively employed as a catalyst for the hydrogenation of 1,1-disubstituted enamines [174]. This reaction yielded enantiomerically enriched tertiary amines 239ai (>91%ee) with good yields (73–92%) (Scheme 77). Sterically hindered substrates reacted under elevated pressures (80 psi) and temperatures. Bulkier or bromine-containing substrates did not participate in the reaction.
The catalytic system based on titanocene difluoride p-S,p-S-235 demonstrated high stereoselectivity in the hydrosilylation of imines, owing to the rapid conversion of the Ti-F bond to Ti-H under the influence of PhSiH3 or polymethylhydrosiloxane (PMHS) in a MeOH/pyrrolidine medium (Scheme 78) [175,176,177,178,179]. Consequently, the asymmetric hydrosilylation of N-aryl-substituted imines 240aad using PhSiH3 or PMHS was developed, providing chiral amines 241aad with yields of 55–97% and substantial enantiomeric excesses.
Hydrosilylation of rac-2,5-disubstituted pyrroles 242af in the presence of the binaphtholate complex S,p-S,p-S-12 was accompanied by the kinetic resolution of the substrates, resulting in cis-pyrrolidines 243ah with yields of 34–44% and enantiomeric purity exceeding 95%ee (Scheme 79) [180]. The unreacted pyrrole was isolated with yields of 33–43%, also exhibiting a high enantiomeric excess of >95%ee. The reaction catalyzed by the difluoride complex p-R,p-R-235 [181] resulted in a highly enantioselective (>94%ee) kinetic resolution of the starting substrates 242a,d,g, yielding pyrrolidines 243a,d,g, along with isomers of the starting pyrrole iso-242a,d,g.
Optically active indanones or tetralones are also of significant interest as synthons for the synthesis of biologically active compounds. The kinetic resolution of N-alkyl (R = Me, i-Pr, n-Pr) imines derived from 3-substituted indanones 244af and 4-substituted tetralones 244gj can be achieved through hydrosilylation in the presence of catalytic amounts of titanocene difluoride p-R,p-R-235 or p-S,p-S-235 (Scheme 80) [182]. Following the hydrolysis of the reaction mixture, ketones 245aj were obtained with enantiomeric purity of 68–99%ee, as well as amines 246aj exhibiting high diastereomeric (cis > 82%) and enantiomeric purity (47–97%ee). This method has been successfully applied in the diastereo- and enantioselective synthesis of the antidepressant sertraline 1S,4S-246j.
An example of hydrosilylation of alkenes was demonstrated using phenylbutene 247a, which, in the reaction with PhSiH3 in the presence of samarium complexes R- or S-137a, provided silane 248 with yields up to 98% and enantioselectivity of 65–68%ee (Scheme 81) [183].

3.9. Enantioselective Hydrogenation of Olefins

Asymmetric hydrogenation of olefins has been the subject of extensive research [184,185,186,187]. Enantiomerically pure ansa-complexes have also been evaluated as catalysts for this process. Pino et al. [188] (Scheme 82) demonstrated that the hydrogenation of 2-phenylbut-1-ene (247a), catalyzed by zirconocene binaphtholate R,p-R,p-R-16 and MAO, yielded 2-phenylbutane 249a with a good yield (over 89%) but low enantioselectivity (36%ee, R), while rac-13 exhibited no catalytic activity. The hydrogenation of 247a in the presence of enantiomerically pure zirconocene p-S,p-S-17 and [PhMe2NH][Co(C2B9H11)2] produced phenylbutane with an enantioselectivity of 23%ee, S (Scheme 82) [189]. It is suggested that the low enantioselectivity in these systems arises from the minimal energy differences between states with different enantiofaces of the olefin at the catalytic site.
Upon activation with n-BuLi, diastereomers of complexes 88a,b catalyzed the hydrogenation of 2-phenylbut-1-ene, with a moderate enantioselectivity of 35–60%ee [168].
Enantiomerically pure C1-symmetric Sm complexes S- and R-137a,b exhibited higher enantioselectivity (up to 96%ee) in this reaction; the enantioselectivity was regulated by varying the reaction conditions and the ratio of enantiomers of the catalyst (Scheme 82) [83,183].
The polymerization of pentene-1 in the presence of D2 and R,p-R,p-R-16 or p-R,p-R-17 and MAO resulted in the formation of deuterated products with the general formula C5H12−xDx (Scheme 53) [140,188]. Through fractional distillation, 1,2-dideuteriopentane (249b) was isolated with an enantiomeric excess of 23–35%ee, and the R-configuration of the product was determined by optical rotation. As a result of styrene deuteration, the R-enantiomer of 249b was obtained with enantiomeric excesses of 65% and 43%ee when using R,p-R,p-R-16 [188] and R-137b [83], respectively (Scheme 82). The mixture of S-137b/R-137b (70/30) allowed for the attainment of a relatively high enantiomeric excess of the S-product at the level of 72%ee.
Trisubstituted olefins 250ah were hydrogenated under the 2000 psi of H2 in the presence of Ti complex S,p-S,p-S-12, activated with n-BuLi at 1 atm of H2 and stabilized by phenylsilane. The reaction afforded products 251ah with high yields of 70–94% and enantioselectivity up to 99%ee (Scheme 83) [190].
The hydrogenations of tetrasubstituted olefins were conducted at ambient temperature in aromatic hydrocarbons utilizing either p-R,p-R- or p-S,p-S- dimethylzirconocene complex 17 and activator [PhMe2NH]+[B(C6F5)4], applying hydrogen pressures of 80 or 1000–2000 psig (Scheme 84) [191]. Products 253ah exhibiting high diastereomeric (>95%, cis) and enantiomeric purity exceeding 90% were achievable at elevated pressures and catalyst concentrations. The reaction mechanism suggests the formation of highly reactive cationic species bearing a Zr-H bond, which engage in reactions with re face of alkene.

3.10. Hydroamination and Cyclization of Amino- and Phosphinoalkenes

The binaphtholate complex S,p-S,p-S-16 was employed in the synthesis of enantiomerically enriched N-heterocycles 255 and 256 (Scheme 85) [192]. In this reaction, the substituent at the double bond of the alkenyl group of the amine significantly influenced the course of the reaction. For example, the reaction of Bu2Mg with diallylbenzylamine in the presence of R,p-R,p-R-16 produced heterocycle 255a with low yield and enantiomeric excess (up to 13%). The reaction of substituted alkenes 254bj with BuMgCl, catalyzed by S,p-S,p-S-16, afforded five- and six-membered cyclic products 255aj and 256c,d with yields of up to 84% (combined cis- and trans-) and enantioselectivity of up to 95%ee.
The successful application of enantiomerically enriched η5-complexes of lanthanide metals has been demonstrated in the reaction of asymmetric hydroamination/cyclization of amino- and phosphinoalkenes 257ae [85,193,194,195,196] (Scheme 86). With a substrate-to-catalyst ratio of (40–200)/1, nearly 100% conversion of N- or P-substituted alkenes was achieved with over 95% regioselectivity. Thus, C1-symmetric complexes of Sm and La with menthyl- or neomenthyl-substituted ansa-cyclopentadienyl ligands catalyzed the hydroamination/cyclization of aminoalkenes 257ac, providing the corresponding pyrrolidines and piperidines 258ac with substrate conversion of up to 100% and enantiomeric purity < 74%ee [193,194,195]. The high diastereoselectivity of the reaction was demonstrated using 2-aminohexene 257d, which was transformed into trans-2,5-dimethylpyrrolidine 258d with >95%de in the presence of complexes 133a and 137a.
Furthermore, complexes with octahydrofluorenyl ligand S-141ac (Sm, Y, Lu) exhibited activity and stereoselectivity in the hydroamination/cyclization of aminoalkenes 257ac,e (Scheme 87) [85]. However, these complexes proved to be less active and stereoselective compared to cyclopentadienyl complexes 133 and 137. Enantioselectivity at the level of 67%ee was achieved only in the case of the sterically hindered substrate 257c (Scheme 87). It was noted that the decrease in turnover frequency (Nt) and selectivity for aminoalkenes 257a,b correlated with the decrease in atomic radii of the lanthanide centers Sm3+—Y3+—Lu3+. The hydrophosphination/cyclization of phosphinoalkenes 259a,b catalyzed by S-141ac under the same conditions produced phospholanes 260a,b with high diastereoselectivity of up to 96%de (Scheme 87).
The cyclization of aminodienes 261ac in the presence of complexes S-141a, S-133b, or S-134b yielded enantiomerically enriched substituted pyrrolidines and piperidines 262ac (Scheme 88) [196,197,198,199]. The natural alkaloid (+)-coniine·HCl (264) was obtained from the prochiral aminodiene 261b via a two-step sequence in high isolated yield (S-141a was used as a catalyst). Catalytic hydroamination/cyclization of aminoalkenes 261df with an internal double bond was also investigated (Scheme 88) [197,198]. The highest enantioselectivity was observed in the reaction catalyzed by Y complex S-141b.

3.11. Other Applications of Enantiomerically Enriched ansa-Complexes in Asymmetric Catalysis

Epoxides are valuable starting materials in organic synthesis due to their easy availability and high reactivity of the three-membered ring. Asymmetric catalytic epoxidation of alkyl- and aryl-substituted alkenes 265ag using t-BuOOH in toluene at 40–80 °C was conducted in the presence of ansa complexes 38 and R,p-R,p-R-41 (Scheme 89). The enantiomeric purity of epoxides 266ag did not exceed 22%ee, with substrate conversion of 6–25% [62,200].
Enantiomerically pure ansa-titanocene dichloride p-S,p-S-11 catalyzed the ring-opening of epoxides 267ac, producing the corresponding alcohols 268ac with moderate enantioselectivity (56%ee), compared to conformationally labile neomenthylcyclopentadienyl titanium complexes 269ac (more than 90%ee) (Scheme 90) [201,202,203].
Chiral bis-(indenyl)-(52) and bis-(tetrahydroindenyl)-(53) titanocenes were tested as catalysts in the homocoupling of benzaldehyde 270. As a result, pinacol-type compounds 271 were obtained (Scheme 91) [66]. It was found that the steric hindrance in tetrahydroindenyl complex 53 provided a lower yield but a higher enantiomeric excess of rac-1,2-diphenylethane-1,2-diol (32%ee).
The catalytic properties of the Ti complex p-R,p-R-11 were demonstrated in the synthesis of γ-lactol (+)-274. The product was isolated with a yield of 49% and enantiomeric purity of 33%ee (Scheme 92) [204].
The potential application of chiral titanium ansa-complexes 52 or 73 in enantioselective isomerization of trans- or cis-4-tert-butyl-1-vinylcyclohexanes 275 to 4-tert-butyl-1-(ethylidene)cyclohexanes 276 was studied (Scheme 93) [71,205]. Complex 52 showed higher activity and enantioselectivity compared to the catalyst 73. The reaction rate and the enantiomeric purity of the product were highly influenced by the temperature of the reaction. The decrease in enantiomeric purity observed at higher temperatures (180 °C) is believed to be caused by racemization through product equilibration, a phenomenon that is likely to be overcome with longer reaction times (12–120 h).

4. Conclusions

Chiral metal complexes of transition metals have occupied significant positions in organometallic chemistry and catalysis. Moreover, the potential applications for enantiomerically pure complexes include their use as sensors, antitumor agents, and other medicinal drugs, as well as templates for constructing 2D and 3D supramolecular structures. The introduction of ansa-η5-ligands significantly alters the electronic and spatial structure of the complexes, as well as their dynamics. Enantiomerically pure ansa-η5-complexes can be obtained through several approaches: synthesis from optically active precursors of η5-ligands, separation of diastereomers of the complexes utilizing enantiomerically pure agents, and the reaction of chiral σ-complexes with prochiral anions of η5-ligands. In organometallic synthesis, kinetic resolution of racemic complexes through interaction with derivatizing enantiomerically pure reagents is most commonly employed. However, the approach involving the fixation of ligands with a relatively rigid chiral bridge provides excellent results, allowing for the immediate production of enantiomers of metallocenes, including those that are less stable, such as indenyl complexes, which typically do not withstand separation using derivatizing reagents. The yield of the reaction and stereoselectivity in the synthesis of metal complexes is predetermined by a combination of numerous factors: the nature of the transition metal, the structure of the ligand, and the conditions of the synthesis (solvent, temperature, reagent ratio, etc.). As literature analysis shows, there is currently no universal solution to the problem of selecting these factors in the synthesis of desired enantiomerically pure metallocene structures.
The ansa-complexes of IVB and IIIB group metals demonstrate exceptional activity and stereoselectivity in the formation of metal–carbon, carbon–carbon, carbon–hydrogen, and carbon–heteroatom bonds. These complexes have proven to be effective catalysts for enantioselective alkene carbometalation, oligomerization, and polymerization. Notably, the stability of the spatial structure of ansa-complexes enhances stereoselectivity as the oligomeric chain grows (chain control). In contrast, chiral conformationally labile complexes can achieve high stereoselectivity at the initial stages of alkene insertion into catalytically active sites (site control); however, they struggle to maintain adequate control over stereoinduction as the chain lengthens. This area of research holds significant promise, particularly in advancing diastereo- and enantioselective oligomerization and polymerization of alkenes using conformationally stable complexes, facilitating the stereoselective synthesis of biologically active compounds and optically active polymers, which are in demand for the development of new materials.
Examples of asymmetric cycloaddition, cyclization, hydrosilylation, hydroamination, hydrogenation, etc., catalyzed by enantiomerically pure ansa-complexes of Group IIIB and IVB metals demonstrate the potential of these complexes as effective catalysts for practically important processes. Further advancement in this area necessitates an understanding of the stereoinduction and chiral recognition mechanisms that enable comprehension of the structure and dynamics of catalytically active centers, as well as the processes of substrate coordination and transformations. These studies are highly relevant, as their realization will facilitate a more optimal approach to the development of catalytic systems and processes based on them.

Author Contributions

Conceptualization, P.V.K., L.V.P. and L.M.K.; writing—original draft preparation, P.V.K. and L.V.P.; writing—review and editing, P.V.K., L.V.P. and L.M.K.; visualization, P.V.K. and L.V.P.; supervision, L.V.P. and L.M.K.; project administration, L.M.K.; funding acquisition, L.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Russian Science Foundation, grant number 23-73-00024.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Marek, I. Titanium and Zirconium in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2002; p. 537. [Google Scholar]
  2. Takahashi, T. Metallocenes in Regio- and Stereoselective Synthesis; Springer: Berlin/Heidelberg, Germany, 2005; p. 244. [Google Scholar]
  3. Zohuri, G.H.; Albahily, K.; Schwerdtfeger, E.D.; Miller, S.A. 3.21—Metallocene Alkene Polymerization Catalysts. In Polymer Science: A Comprehensive Reference; Moeller, M., Matyjaszewski, K., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2012; Volume 3, pp. 673–697. [Google Scholar] [CrossRef]
  4. Vaughan, A.; Davis, D.S.; Hagadorn, J.R.; Salwani, M.S. Industrial Catalysts for Alkene Polymerization. Ref. Modul. Mater. Sci. Mater. Eng. 2017, 1–18. [Google Scholar] [CrossRef]
  5. Kaminsky, W. Discovery of Methylaluminoxane as Cocatalyst for Olefin Polymerization. Macromolecules 2012, 45, 3289–3297. [Google Scholar] [CrossRef]
  6. Chen, E.Y.X.; Marks, T.J. Cocatalysts for Metal-Catalyzed Olefin Polymerization:  Activators, Activation Processes, and Structure−Activity Relationships. Chem. Rev. 2000, 100, 1391–1434. [Google Scholar] [CrossRef]
  7. Coates, G.W. Precise Control of Polyolefin Stereochemistry Using Single-Site Metal Catalysts. Chem. Rev. 2000, 100, 1223–1252. [Google Scholar] [CrossRef]
  8. Atiqullah, M.; Tinkl, M.; Pfaendner, R.; Akhtar, M.N.; Hussain, I. Synthesis of Functional Polyolefins using Metallocenes: A Comprehensive Review. Polym. Rev. 2010, 50, 178–230. [Google Scholar] [CrossRef]
  9. Collins, R.A.; Russell, A.F.; Mountford, P. Group 4 metal complexes for homogeneous olefin polymerisation: A short tutorial review. Appl. Petrochem. Res. 2015, 5, 153–171. [Google Scholar] [CrossRef]
  10. Brintzinger, H.H.; Fischer, D.; Mülhaupt, R.; Rieger, B.; Waymouth, R.M. Stereospecific Olefin Polymerization with Chiral Metallocene Catalysts. Angew. Chem. Inter. Ed. Engl. 1995, 34, 1143–1170. [Google Scholar] [CrossRef]
  11. Brintzinger, H.H.; Fischer, D. Development of ansa-Metallocene Catalysts for Isotactic Olefin Polymerization. Adv. Polym. Sci. 2013, 258, 29–42. [Google Scholar] [CrossRef]
  12. Janiak, C. Metallocene and related catalysts for olefin, alkyne and silane dimerization and oligomerization. Coord. Chem. Rev. 2006, 250, 66–94. [Google Scholar] [CrossRef]
  13. Janiak, C.; Blank, F. Metallocene Catalysts for Olefin Oligomerization. Macromol. Symp. 2006, 236, 14–22. [Google Scholar] [CrossRef]
  14. Kaminsky, W.; Arndt, M. Chapter 2.3: Reactions of Unsaturated Compounds Polymerization, Oligomerization, and Copolymerization of Olefins Chemical Background. In Applied Homogeneous Catalysis with Organometallic Compounds; Cornils, B., Herrmann, W.A., Eds.; Wiley-VCH: Weinheim, Germany, 1996; pp. 213–385. [Google Scholar] [CrossRef]
  15. Skupinska, J. Oligomerization of α-olefins to higher oligomers. Chem. Rev. 1991, 91, 613–648. [Google Scholar] [CrossRef]
  16. Parfenova, L.V.; Bikmeeva, A.K.; Kovyazin, P.V.; Khalilov, L.M. The Dimerization and Oligomerization of Alkenes Catalyzed with Transition Metal Complexes: Catalytic Systems and Reaction Mechanisms. Molecules 2024, 29, 502. [Google Scholar] [CrossRef] [PubMed]
  17. Nicholas, C.P. Applications of light olefin oligomerization to the production of fuels and chemicals. Appl. Catal. A Gen. 2017, 543, 82–97. [Google Scholar] [CrossRef]
  18. Nifant’ev, I.; Ivchenko, P.; Tavtorkin, A.; Vinogradov, A.; Vinogradov, A. Non-traditional Ziegler-Natta catalysis in α-olefin transformations: Reaction mechanisms and product design. Pure Appl. Chem. 2017, 89, 1017–1032. [Google Scholar] [CrossRef]
  19. Nifant’ev, I.; Ivchenko, P. Fair look at coordination oligomerization of higher α-olefins. Polymers 2020, 12, 1082. [Google Scholar] [CrossRef]
  20. Schwartz, J.; Labinger, J.A. Hydrozirconation: A New Transition Metal Reagent for Organic. Angew. Chem. Inter. Ed. Engl. 1976, 15, 333–340. [Google Scholar] [CrossRef]
  21. Dzhemilev, U.M.; Ibragimov, A.G. Metal complex catalysis in the synthesis of organoaluminium compounds. Russ. Chem. Rev. 2000, 69, 121–135. [Google Scholar] [CrossRef]
  22. Guiry, P.J.; Coyne, A.G.; Carroll, A.M. C-E bond formation through hydroboration and hydroalumination. In Comprehensive Organometallic Chemistry III; Michael, D., Mingos, P., Crabtree, R.H., Eds.; Elsevier: Amsterdam, The Netherlands, 2007; Volume 10, pp. 839–869. [Google Scholar] [CrossRef]
  23. Sato, F. Organic Syntheses via Hydroalumination of Olefins Catalyzed by Titanium or Zirconium Compounds. J. Synth. Org. Chem. 1980, 38, 234–243. [Google Scholar] [CrossRef]
  24. Parfenova, L.V.; Khalilov, L.M.; Dzhemilev, U.M. Mechanisms of reactions of organoaluminium compounds with alkenes and alkynes catalyzed by Zr complexes. Russ. Chem. Rev. 2012, 81, 524–548. [Google Scholar] [CrossRef]
  25. Hoveyda, A.H.; Morken, J.P. Enantioselective C-C and C-H Bond Formation Mediated or Catalyzed by Chiral ebthi Complexes of Titanium and Zirconium. Angew. Chem. Inter. Ed. Engl. 1996, 35, 1262–1284. [Google Scholar] [CrossRef]
  26. Negishi, E.I. Discovery of ZACA reaction: Zr-catalyzed asymmetric carboalumination of Alkenes. Arkivoc 2011, 2011, 34–53. [Google Scholar] [CrossRef]
  27. Tanaka, K. Chapter 18: Asymmetric carbometallations including carbocyclizations. In Catalytic Asymmetric Synthesis, 4th ed.; Akiyama, T., Ojima, I., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2022; pp. 705–726. [Google Scholar] [CrossRef]
  28. D’yakonov, V.A. Dzhemilev reaction for the synthesis of cyclic and acyclic organometallic compounds. In Organometallic Compounds: Preparation, Structure and Properties; Nova Science Publishers: Hauppauge, NY, USA, 2011; p. 96. [Google Scholar]
  29. Beweries, T.; Haehnel, M.; Rosenthal, U. Recent advances in the chemistry of heterometallacycles of group 4 metallocenes. Catal. Sci. Technol. 2013, 3, 18–28. [Google Scholar] [CrossRef]
  30. Dzhemilev, U.M.; Ibragimov, A.G. Catalytic cyclometalation reaction of unsaturated compounds in synthesis of magnesa- and aluminacarbocycles. J. Organomet. Chem. 2010, 695, 1085–1110. [Google Scholar] [CrossRef]
  31. Caruso, F.; Rossi, M.; Opazo, C.; Pettinari, C. Structural features of antitumor titanium agents and related compounds. Bioinorg. Chem. Appl. 2005, 2005, 317–329. [Google Scholar] [CrossRef]
  32. Pradines, B.; Orlandi-Pradines, E.; Henry, M.; Bogreau, H.; Fusai, T.; Mosnier, J.; Baret, E.; Durand, C.; Bouchiba, H.; Penhoat, K.; et al. Métallocènes et paludisme: Une nouvelle approche thérapeutiqueMetallocenes and malaria: A new therapeutic approach. Ann. Pharm. Fr. 2005, 63, 284–294. [Google Scholar] [CrossRef]
  33. Medici, S.; Peana, M.; Nurchi, V.M.; Lachowicz, J.I.; Crisponi, G.; Zoroddu, M.A. Noble metals in medicine: Latest advances. Coord. Chem. Rev. 2015, 284, 329–350. [Google Scholar] [CrossRef]
  34. Meléndez, E. Metallocenes as target specific drugs for cancer treatment. Inorg. Chim. Acta 2012, 393, 36–52. [Google Scholar] [CrossRef]
  35. Erker, G. Group 4 metallocenes in bioorganometallic chemistry. J. Organomet. Chem. 2007, 692, 1187–1197. [Google Scholar] [CrossRef]
  36. Paley, R.S. Enantiomerically Pure Planar Chiral Organometallic Complexes via Facially Selective π-Complexation. Chem. Rev. 2002, 102, 1493–1524. [Google Scholar] [CrossRef]
  37. Shapiro, P.J. The evolution of the ansa-bridge and its effect on the scope of metallocene chemistry. Coord. Chem. Rev. 2002, 231, 67–81. [Google Scholar] [CrossRef]
  38. Wang, B. Ansa-metallocene polymerization catalysts: Effects of the bridges on the catalytic activities. Coord. Chem. Rev. 2006, 250, 242–258. [Google Scholar] [CrossRef]
  39. Halteman, R.L. Chapter 8: Synthesis of Chiral Titanocene and Zirconocene Dichlorides. In Metallocenes: Synthesis Reactivity Applications; Togni, A., Halterman, R.L., Eds.; Wiley-VCH: Weinheim, Germany, 1998; pp. 455–544. [Google Scholar] [CrossRef]
  40. Bau, M.A.; Wiesler, S.; Younas, S.L.; Streuff, J. Strategies for the Synthesis of Chiral Carbon-Bridged Group IV ansa-Metallocenes. Chem. Eur. J. 2019, 25, 10531–10545. [Google Scholar] [CrossRef] [PubMed]
  41. Halterman, R.L. Synthesis and applications of chiral cyclopentadienylmetal complexes. Chem. Rev. 1992, 92, 965–994. [Google Scholar] [CrossRef]
  42. Hoveyda, A.H. Chapter 6: Chiral Zirconium Catalysts for Enantioselective Synthesis. In Titanium and Zirconium in Organic Synthesis; Marek, I., Ed.; Wiley-VCH: Weinheim, Germany, 2002; pp. 180–229. [Google Scholar] [CrossRef]
  43. Schnutenhaus, H.; Brintzinger, H.H. 1,1′-Trimethylenebis(η5-3-tert-butylcyclopentadienyl)-titanium(IV)Dichloride, a Chiral ansa-Titanocene Derivative. Angew. Chem. Inter. Ed. Engl. 1979, 18, 777–778. [Google Scholar] [CrossRef]
  44. Kuntz, B.A.; Ramachandran, R.; Taylor, N.J.; Guan, J.; Collins, S. Addition of a sterically hindered, chiral crotyl ansa-titanocene complex to aldehydes. J. Organomet. Chem. 1995, 497, 133–142. [Google Scholar] [CrossRef]
  45. Schmidt, K.; Reinmuth, A.; Rief, U.; Diebold, J.; Brintzinger, H.H. Photochemical Isomerization of Me2Si-Bridged Zirconocene Complexes. Application to Stereoselective Syntheses of ansa-Zirconocene Binaphtholate Stereoisomers. Organometallics 1997, 16, 1724–1728. [Google Scholar] [CrossRef]
  46. Wild, F.R.W.P.; Zsolnai, L.; Huttner, G.; Brintzinger, H.H. ansa-Metallocene derivatives IV. Synthesis and molecular structures of chiral ansa-titanocene derivatives with bridged tetrahydroindenyl ligands. J. Organomet. Chem. 1982, 232, 233–247. [Google Scholar] [CrossRef]
  47. Schäfer, A.; Karl, E.; Zsolnai, L.; Huttner, G.; Brintzinger, H.H. ansa-Metallocene derivatives: XII. Diastereomeric derivatisation and enantiomer separation of ethylenebis(tetrahydronindenyl)-titanium and -zirconium dichlorides. J. Organomet. Chem. 1987, 328, 87–99. [Google Scholar] [CrossRef]
  48. Collins, S.; Kuntz, B.A.; Hong, Y. Additions of chiral allyltitanocenes to aldehydes: Diastereoselective synthesis of homoallylic alcohols with a recyclable chiral transition metal reagent. J. Org. Chem. 1989, 54, 4154–4158. [Google Scholar] [CrossRef]
  49. Collins, S.; Hong, Y.; Taylor, N.J. General synthetic routes to chiral, ethylene-bridged ansa-titanocene dichlorides. Organometallics 1990, 9, 2695–2703. [Google Scholar] [CrossRef]
  50. Grossman, R.B.; Davis, W.M.; Buchwald, S.L. Enantioselective, zirconium-mediated synthesis of allylic amines. J. Am. Chem. Soc. 1991, 113, 2321–2322. [Google Scholar] [CrossRef]
  51. Parfenova, L.V.; Zakirova, I.V.; Kovyazin, P.V.; Karchevsky, S.G.; Istomina, G.P.; Khalilov, L.M.; Dzhemilev, U.M. Intramolecular mobility of η5-ligands in chiral zirconocene complexes and the enantioselectivity of alkene functionalization by organoaluminum compounds. Dalt. Trans. 2016, 45, 12814–12826. [Google Scholar] [CrossRef] [PubMed]
  52. Huttenloch, M.E.; Dorer, B.; Rief, U.; Prosenc, M.H.; Schmidt, K.; Brintzinger, H.H. ansa-metallocene derivatives XXXIX biphenyl-bridged metallocene complexes of titanium, zirconium, and vanadium: Syntheses, crystal structures and enantioseparation. J. Organomet. Chem. 1997, 541, 219–232. [Google Scholar] [CrossRef]
  53. Ringwald, M.; Stürmer, R.; Brintzinger, H.H. Asymmetric Thermal Transformation, a New Way to Enantiopure Biphenyl-Bridged Titanocene and Zirconocene Complexes:  Efficient Catalysts for Asymmetric Imine Hydrogenation. J. Am. Chem. Soc. 1999, 121, 1524–1527, Erratum in J. Am. Chem. Soc. 1999, 121, 1524–1527. https://doi.org/10.1021/ja9955161. [Google Scholar] [CrossRef]
  54. Chin, B.; Buchwald, S.L. An Improved Procedure for the Resolution of (rac)-Ethylenebis(tetrahydroindenyl)Titanium Derivatives. J. Org. Chem. 1996, 61, 5650–5651. [Google Scholar] [CrossRef]
  55. Chin, B.; Buchwald, S.L. Improved Procedure for the Preparation of Enantiomerically Pure Ethylenebis(tetrahydroindenyl)zirconium Derivatives. J. Org. Chem. 1997, 62, 2267–2268. [Google Scholar] [CrossRef]
  56. Habaue, S.; Sakamoto, H.; Okamoto, Y. Optical Resolution of Chiral Ethylenebis(4,5,6,7-tetrahydro-1-indenyl)zirconium Derivatives by High-Performance Liquid Chromatography. Chem. Lett. 1996, 25, 383–384. [Google Scholar] [CrossRef]
  57. Habaue, S.; Sakamoto, H.; Baraki, H.; Okamoto, Y. Chiral ethylenebis(4,5,6,7-tetrahydro-1-indenyl) complexes of zirconium and hafnium: Separation and application to asymmetric polymerization. Macromol. Rapid. Commun. 1997, 18, 707–713. [Google Scholar] [CrossRef]
  58. Hong, Y.; Kuntz, B.A.; Collins, S. Asymmetric induction in the Diels-Alder reaction using chiral metallocene catalysts. Organometallics 1993, 12, 964–969. [Google Scholar] [CrossRef]
  59. Jaquith, J.B.; Guan, J.; Wang, S.; Collins, S. Asymmetric Induction in the Diels-Alder Reaction Catalyzed by Chiral Metallocene Triflate Complexes: Dramatic Effect of Solvent Polarity. Organometallics 1995, 14, 1079–1081. [Google Scholar] [CrossRef]
  60. Jaquith, J.B.; Levy, C.J.; Bondar, G.V.; Wang, S.; Collins, S. Diels−Alder Reactions of Oxazolidinone Dienophiles Catalyzed by Zirconocene Bis(triflate) Catalysts:  Mechanism for Asymmetric Induction. Organometallics 1998, 17, 914–925. [Google Scholar] [CrossRef]
  61. Ellis, W.W.; Hollis, T.K.; Odenkirk, W.; Whelan, J.; Ostrander, R.; Rheingold, A.L.; Bosnich, B. Synthesis, structure, and properties of chiral titanium and zirconium complexes bearing biaryl strapped substituted cyclopentadienyl ligands. Organometallics 1993, 12, 4391–4401. [Google Scholar] [CrossRef]
  62. Halterman, R.L.; Ramsey, T.M. Asymmetric synthesis of a sterically rigid binaphthyl-bridged chiral metallocene: Asymmetric catalytic epoxidation of unfunctionalized alkenes. Organometallics 1993, 12, 2879–2880. [Google Scholar] [CrossRef]
  63. Burk, M.J.; Colletti, S.L.; Halterman, R.L. C2-symmetric 2,2′-dimethyl-1,1′-binaphthyl-bridged ansa-bis(1-indenyl)metal complexes. Organometallics 1991, 10, 2998–3000. [Google Scholar] [CrossRef]
  64. Halterman, R.L.; Combs, D.; Kihega, J.G.; Khan, M.A. Synthesis of ansa-2,2′-bis[(4,7-dimethyl-inden-1-yl)methyl]-1,1′-binaphthyl and ansa-2,2′-bis[(4,5,6,7-tetrahydroinden1-yl)methyl]-1,1′-binaphthyltitanium and -zirconium dichlorides. J. Organomet. Chem. 1996, 520, 163–170. [Google Scholar] [CrossRef]
  65. Rheingold, A.L.; Robinson, N.P.; Whelan, J.; Bosnich, B. Preparation and properties of chiral titanocene and zirconocene dichloride complexes of a chiral ligand. Organometallics 1992, 11, 1869–1876. [Google Scholar] [CrossRef]
  66. Halterman, R.L.; Zhu, C.; Chen, Z.; Dunlap, M.S.; Khan, M.A.; Nicholas, K.M. Preparation of [2,5-Diisopropylcyclohexane-1,4- bis(indenyl)]titanium Dichloride and [2,5-Diisopropylcyclohexane-1,4-bis(tetrahydroindenyl)]- titanium Dichloride and Their Comparison as Catalysts for the Enantioselective Pinacol Coupling of Benzaldehyde. Organometallics 2000, 19, 3824–3829. [Google Scholar] [CrossRef]
  67. Hollis, T.K.; Robinson, N.P.; Whelan, J.; Rhelngold, A.L.; Bosnich, B. Preparation and properties of (S,S)-[Ti((R,R)-cyclacene)Cl2], a chiral strapped bent metallocene. Organometallics 1992, 11, 2812–2816. [Google Scholar] [CrossRef]
  68. Bandy, J.A.; Green, M.L.H.; Gardiner, I.M.; Prout, K. Polymerisation of ethylene and propene using new chiral zirconium derivatives. Crystal structure of [ZrL1Cl2][H2L1 = (4S,5S)-trans-4,5-bis(1H-inden-1-ylmethyl)-2,2-dimethyl-1,3-dioxolane]. J. Chem. Soc., Dalton Trans. 1991, 8, 2207–2216. [Google Scholar] [CrossRef]
  69. Thiele, S.; Erker, G.; Fritze, C.; Psiorz, C.; Fröhlich, R. Synthesis of a Chiral Meso-Like Ansa-Zirconocene Complex and its Use for the Catalytic Formation of Low Molecular Weight Polypropylene. Z. Naturforschung Sect. B J. Chem. Sci. 1995, 50, 982–989. [Google Scholar] [CrossRef]
  70. Chen, Y.-X.; Rausch, M.D.; Chien, J.C.W. Heptane-soluble homogeneous zirconocene catalyst: Synthesis of a single diastereomer, polymerization catalysis, and effect of silica supports. J. Polym. Sci. A Polym. Chem. 1995, 33, 2093–2108. [Google Scholar] [CrossRef]
  71. Halterman, R.L.; Chen, Z.; Khan, M.A. Synthesis, Structure Determination, and Reactivity of C2-Symmetrical Ethylene-Bridged ansa-Bis(DiMeBCOCp)titanium Dichlorides. Organometallics 1996, 15, 3957–3967. [Google Scholar] [CrossRef]
  72. Taber, D.F.; Balijepalli, B.; Liu, K.K.; Kong, S.; Rheingold, A.L.; Askham, F.R. Designing the Chiral Ligand Space around an Early Transition Metal:  Myrtanyl Zirconocene. J. Org. Chem. 1999, 64, 4525–4527. [Google Scholar] [CrossRef]
  73. Wiesler, S.; Bau, M.A.; Younas, S.L.; Luu, H.T.; Kratzert, D.; Streuff, J. A Unified Approach to Customized Chiral Carbon-Bridged ansa-Metallocenes. Chem. Eur. J. 2018, 24, 16532–16536. [Google Scholar] [CrossRef]
  74. Beagley, P.; Davies, P.; Adams, H.; White, C. Chiral metallocenes: The synthesis and X-ray crystal structures of TiCl2η55-C5Me4SiMe2C5H3R*) (R* = menthyl or neomenthyl) and related compounds. Can. J. Chem. 2001, 79, 731–741. [Google Scholar] [CrossRef]
  75. Giardello, M.A.; Eisen, M.S.; Stern, C.L.; Marks, T.J. Chiral C1-Symmetric Group 4 Metallocenes as Catalysts for Stereoregular.alpha.-Olefin Polymerization. Metal, Ancillary Ligand, and Counteranion Effects. J. Am. Chem. Soc. 1995, 117, 12114–12129. [Google Scholar] [CrossRef]
  76. Obora, Y.; Stern, C.L.; Marks, T.J.; Nickias, P.N. Ancillary Ligand Effects in Chiral C1-Symmetric ansa-Metallocene Catalysts for Stereoregular α-Olefin Polymerization. “Wingspan” Modification with Octahydrofluorene. Organometallics 1997, 16, 2503–2505. [Google Scholar] [CrossRef]
  77. Halterman, R.L.; Fahey, D.R.; Marin, V.P.; Dockter, D.W.; Khan, M.A. Synthesis, characterization and polymerization properties of isopropylidene(η5-3-neomenthylcyclopentadienyl)(η5-fluorenyl)zirconium dichloride. J. Organomet. Chem. 2001, 625, 154–159. [Google Scholar] [CrossRef]
  78. Baker, R.W.; Foulkes, M.A.; Turner, P. Enantiospecific synthesis of a planar chiral fluorenylzirconium ansa-metallocene: Synthesis and structures of [1-(9-fluorenyl)- and [1-(1-methyl-9-fluorenyl)-2-(2-indenyl)naphthalene]zirconium dichloride. J. Chem. Soc. Dalton Trans. 2000, 4, 431–433. [Google Scholar] [CrossRef]
  79. Rieger, B.; Jany, G. A Convenient Synthesis of Enantiomerically Pure Ethylene-Bridged Metallocene Complexes Bearing Fluorenyl- and Octahydrofluorenyl Ligands. Chem. Ber. 1994, 127, 2417–2419. [Google Scholar] [CrossRef]
  80. Jany, G.; Fawzi, R.; Steimann, M.; Rieger, B. Synthesis of Enantiomerically Pure Ethylene-Bridged ansa-Zirconocene and -Hafnocene Complexes Bearing Fluorenyl, Indenyl, Octahydrofluorenyl, and Tetrahydroindenyl Ligands. Organometallics 1997, 16, 544–550. [Google Scholar] [CrossRef]
  81. Baar, C.R.; Levy, C.J.; Min, E.Y.J.; Henling, L.M.; Day, M.W.; Bercaw, J.E. Kinetic Resolution of Chiral α-Olefins Using Optically Active ansa-Zirconocene Polymerization Catalysts. J. Am. Chem. Soc. 2004, 126, 8216–8231. [Google Scholar] [CrossRef] [PubMed]
  82. Aspinall, H.C. Chiral Lanthanide Complexes: Coordination Chemistry and Applications. Chem. Rev. 2002, 102, 1807–1850. [Google Scholar] [CrossRef] [PubMed]
  83. Conticello, V.P.; Brard, L.; Giardello, M.A.; Tsuji, Y.; Sabat, M.; Stern, C.L.; Marks, T.J. Chiral organolanthanide complexes for enantioselective olefin hydrogenation. J. Am. Chem. Soc. 1992, 114, 2761–2762. [Google Scholar] [CrossRef]
  84. Giardello, M.A.; Conticello, V.P.; Brard, L.; Sabat, M.; Stern, C.L.; Marks, T.J.; Rheingold, A.L. Chiral Organolanthanides Designed for Asymmetric Catalysis. Synthesis, Characterization, and Configurational Interconversions of Chiral, C1-Symmetric Organolanthanide Halides, Amides, and Hydrocarbyls. J. Am. Chem. Soc. 1994, 116, 10212–10240. [Google Scholar] [CrossRef]
  85. Douglass, M.R.; Ogasawara, M.; Hong, S.; Metz, M.V.; Marks, T.J. “Widening the Roof”: Synthesis and Characterization of New Chiral C1-Symmetric Octahydrofluorenyl Organolanthanide Catalysts and Their Implementation in the Stereoselective Cyclizations of Aminoalkenes and Phosphinoalkenes. Organometallics 2002, 21, 283–292. [Google Scholar] [CrossRef]
  86. Mitchell, J.P.; Hajela, S.; Brookhart, S.K.; Hardcastle, K.I.; Henling, L.M.; Bercaw, J.E. Preparation and Structural Characterization of an Enantiomerically Pure, C2-Symmetric, Single-Component Ziegler−Natta α-Olefin Polymerization Catalyst. J. Am. Chem. Soc. 1996, 118, 1045–1053. [Google Scholar] [CrossRef]
  87. LoCoco, M.D.; Jordan, R.F. Enantioselective Synthesis of ansa-Zirconocenes. J. Am. Chem. Soc. 2004, 126, 13918–13919. [Google Scholar] [CrossRef]
  88. Xu, L.; Mikami, K. Diastereo- and enantioselective syntheses of ansa-metallocenes from metal halide complexes with tropos biphenol and atropos binaphthol ethers. Tetrahedron Lett. 2004, 45, 9215–9217. [Google Scholar] [CrossRef]
  89. Vaughan, A.; Davis, D.S.; Hagadorn, J.R. Industrial Catalysts for Alkene Polymerization. In Polymer Science: A Comprehensive Reference; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; Volume 3, pp. 657–672. [Google Scholar] [CrossRef]
  90. Kaminsky, W. Highly active metallocene catalysts for olefin polymerization. J. Chem. Soc. Dalton Trans. 1998, 9, 1413–1418. [Google Scholar] [CrossRef]
  91. Kaminsky, W. The discovery of metallocene catalysts and their present state of the art. J. Polym. Sci. A Polym. Chem. 2004, 42, 3911–3921. [Google Scholar] [CrossRef]
  92. Kaminsky, W.; Funck, A.; Hähnsen, H. New application for metallocenecatalysts in olefinpolymerization. Dalton Trans. 2009, 41, 8803–8810. [Google Scholar] [CrossRef] [PubMed]
  93. Shamiri, A.; Chakrabarti, M.H.; Jahan, S.; Hussain, M.A.; Kaminsky, W.; Aravind, P.V.; Yehye, W.A. The Influence of Ziegler-Natta and Metallocene Catalysts on Polyolefin Structure, Properties, and Processing Ability. Materials 2014, 7, 5069–5108. [Google Scholar] [CrossRef]
  94. Hlatky, G.G. Heterogeneous Single-Site Catalysts for Olefin Polymerization. Chem. Rev. 2000, 100, 1347–1376. [Google Scholar] [CrossRef]
  95. Alt, H.G.; Köppl, A. Effect of the Nature of Metallocene Complexes of Group IV Metals on Their Performance in Catalytic Ethylene and Propylene Polymerization. Chem. Rev. 2000, 100, 1205–1222. [Google Scholar] [CrossRef]
  96. Resconi, L.; Cavallo, L.; Fait, A.; Piemontesi, F. Selectivity in Propene Polymerization with Metallocene Catalysts. Chem. Rev. 2000, 100, 1253–1346. [Google Scholar] [CrossRef]
  97. Suzuki, N. Stereospecific Olefin Polymerization Catalyzed by Metallocene Complexes. In Metallocenes in Regio- and Stereoselective Synthesis, 1st ed.; Takahashi, T., Ed.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 177–215. [Google Scholar] [CrossRef]
  98. Scheirs, J.; Kaminsky, W. Metallocene-Based Polyolefins—Preparation, Properties and Technology; Wiley: New York, NY, USA, 2000; p. 560. [Google Scholar]
  99. Guerra, G.; Cavallo, L.; Corradini, P. Chirality of Catalysts for Stereospecific Polymerizations. In Materials-Chirality, 1st ed.; Green, M.M., Nolte, R.J.M., Meijer, E.W., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2003; Volume 24, pp. 1–69. [Google Scholar] [CrossRef]
  100. Nakayama, Y.; Shiono, T. Molecules Developments of Chiral Metallocenes as Polymerization Catalysts. Molecules 2005, 10, 620–633. [Google Scholar] [CrossRef]
  101. Resconi, L.; Chadwick, J.C.; Cavallo, L. Olefin Polymerizations with Group IV Metal Catalysts. In Comprehensive Organometallic Chemistry III; Mingos, D.M.P., Crabtree, R.H., Eds.; Elsevier: Amsterdam, The Netherlands, 2007; Volume 4, pp. 1005–1166. [Google Scholar] [CrossRef]
  102. Zaccaria, F.; Sian, L.; Zuccaccia, C.; Macchioni, A. Chapter One—Ion pairing in transition metal catalyzed olefin polymerization. In Advances in Organometallic Chemistry; Pérez, P.J., Ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2020; pp. 1–78. [Google Scholar] [CrossRef]
  103. Osakada, K. Organometallic Reactions and Polymerization. In Lecture Notes in Chemistry 85; Springer: Berlin/Heidelberg, Germany, 2014; p. 301. [Google Scholar]
  104. Corma, A.; Iborra, S. Oligomerization of Alkenes. In Microporous and Mesoporous Solid Catalysts; Derouane, E.G., Roberts, S.M., Eds.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006; pp. 125–140. [Google Scholar]
  105. Bryliakov, K.P.; Talsi, E.P. Frontiers of mechanistic studies of coordination polymerization and oligomerization of α-olefins. Coord. Chem. Rev. 2012, 256, 2994–3007. [Google Scholar] [CrossRef]
  106. Christoffers, J.; Bergman, R.G. Zirconocene-alumoxane (1:1)—A catalyst for the selective dimerization of α-olefins. Inorg. Chim. Acta 1998, 270, 20–27. [Google Scholar] [CrossRef]
  107. Negishi, E.I.; Huo, S. Synthesis and Reactivity of Zirconocene Derivatives. In Titanium and Zirconium in Organic Synthesis; Marek, I., Ed.; Wiley: Hoboken, NJ, USA, 2002; pp. 1–49. [Google Scholar] [CrossRef]
  108. Negishi, E.I. Asymmetric Carbometallations. In Catalytic Asymmetric Synthesis; Ojima, I., Ed.; Wiley-VCH: Weinheim, Germany, 2000; pp. 165–189. [Google Scholar] [CrossRef]
  109. Xu, S.; Oda, A.; Kamada, H.; Negishi, E.I. Highly enantioselective synthesis of γ-, δ-, and ɛ-chiral 1-alkanols via Zr-catalyzed asymmetric carboalumination of alkenes (ZACA)–Cu- or Pd-catalyzed cross-coupling. Proc. Natl. Acad. Sci. USA 2014, 111, 8368–8373. [Google Scholar] [CrossRef]
  110. Xu, S.; Oda, A.; Bobinski, T.; Li, H.; Matsueda, Y.; Negishi, E.I. Highly Efficient, Convergent, and Enantioselective Synthesis of Phthioceranic Acid. Angew. Chem. Inter. Ed. 2015, 54, 9319–9322. [Google Scholar] [CrossRef] [PubMed]
  111. Xu, S.; Lee, C.T.; Wang, G.; Negishi, E.I. Widely Applicable Synthesis of Enantiomerically Pure Tertiary Alkyl-Containing 1-Alkanols by Zirconium-Catalyzed Asymmetric Carboalumination of Alkenes and Palladium- or Copper-Catalyzed Cross-Coupling. Chem. Asian J. 2013, 8, 1829–1835. [Google Scholar] [CrossRef] [PubMed]
  112. Xu, S.; Oda, A.; Negishi, E.I. Enantioselective Synthesis of Chiral Isotopomers of 1-Alkanols bya ZACA–Cu-Catalyzed Cross-Coupling Protocol. Chem. Eur. J. 2014, 20, 16060–16064. [Google Scholar] [CrossRef] [PubMed]
  113. Hoveyda, A.H.; Morken, J.P. Regio- and stereoselective carbon-carbon bond formation through transition metal catalysis. The influence of catalyst chirality on selective ethylmagnesiation of chiral, nonracemic alcohols and ethers. J. Org. Chem. 1993, 58, 4237–4244. [Google Scholar] [CrossRef]
  114. Morken, J.P.; Didiuk, M.T.; Hoveyda, A.H. Zirconium-catalyzed asymmetric carbomagnesation. J. Am. Chem. Soc. 1993, 115, 6997–6998. [Google Scholar] [CrossRef]
  115. Visser, M.S.; Heron, N.M.; Didiuk, M.T.; Sagal, J.F.; Hoveyda, A.H. Catalytic and Enantioselective Route to Medium-Ring Heterocycles. Asymmetric Zirconium-Catalyzed Ethylmagnesation of Seven- and Eight-Membered Rings. J. Am. Chem. Soc. 1996, 118, 4291–4298. [Google Scholar] [CrossRef]
  116. Morken, J.P.; Didiuk, M.T.; Visser, M.S.; Hoveyda, A.H. Zirconium-Catalyzed Kinetic Resolution of Pyrans. J. Am. Chem. Soc. 1994, 116, 3123–3124. [Google Scholar] [CrossRef]
  117. Visser, M.S.; Hoveyda, A.H. Zirconocene-catalyzed kinetic resolution of dihydrofurans. Tetrahedron 1995, 51, 4383–4394. [Google Scholar] [CrossRef]
  118. Adams, J.A.; Heron, N.M.; Koss, A.M.; Hoveyda, A.H. Stereoselective Chelate-Controlled Addition of Grignard Reagents to Unsaturated Medium-Ring Heterocycles. J. Org. Chem. 1999, 64, 854–860. [Google Scholar] [CrossRef]
  119. Visser, M.S.; Harrity, J.P.A.; Hoveyda, A.H. Zirconium-Catalyzed Kinetic Resolution of Cyclic Allylic Ethers. An Enantioselective Route to Unsaturated Medium Ring Systems. J. Am. Chem. Soc. 1996, 118, 3779–3780. [Google Scholar] [CrossRef]
  120. Johannes, C.W.; Visser, M.S.; Weatherhead, G.S.; Hoveyda, A.H. Zr-Catalyzed Kinetic Resolution of Allylic Ethers and Mo-Catalyzed Chromene Formation in Synthesis. Enantioselective Total Synthesis of the Antihypertensive Agent (S,R,R,R)-Nebivolol. J. Am. Chem. Soc. 1998, 120, 8340–8347. [Google Scholar] [CrossRef]
  121. Adams, J.A.; Ford, J.G.; Stamatos, P.J.; Hoveyda, A.H. Zr-Catalyzed Kinetic Resolution of Aliphatic Cyclic Allylic Ethers. Carbocycles to Heterocycles by Ru- and Mo-Catalyzed Ring-Opening and Ring-Closing Metathesis. J. Org. Chem. 1999, 64, 9690–9696. [Google Scholar] [CrossRef]
  122. Bell, L.; Whitby, R.J.; Jones, R.V.H.; Standen, M.C.H. Catalytic asymmetric carbomagnesiation of unactivated alkenes. A new, effective, active, cheap and recoverable chiral zirconocene. Tetrahedron Lett. 1996, 37, 7139–7142. [Google Scholar] [CrossRef]
  123. Bell, L.; Brookings, D.C.; Dawson, G.J.; Whitby, R.J.; Jones, R.V.H.; Standen, M.C.H. Asymmetric ethylmagnesiation of alkenes using a novel zirconium catalyst. Tetrahedron 1998, 54, 14617–14634. [Google Scholar] [CrossRef]
  124. Sato, F.; Urabe, H.; Okamoto, S. Synthesis of Organotitanium Complexes from Alkenes and Alkynes and Their Synthetic Applications. Chem. Rev. 2000, 100, 2835–2886. [Google Scholar] [CrossRef]
  125. Collins, S.; Dean, W.P.; Ward, D.G. Addition of crotyltitanocenes to aldehydes: Influence of cyclopentadienyl substituents on the diastereoselective formation of homoallylic alcohols. Organometallics 1988, 7, 2289–2293. [Google Scholar] [CrossRef]
  126. Urabe, H.; Yoshikawa, K.; Sato, F. Generation of chiral (η5-indenyl)23-allyl)Ti complex and its highly enantioselective reaction with carbonyl compounds. Tetrahedron Lett. 1995, 36, 5595–5598. [Google Scholar] [CrossRef]
  127. Justicia, J.; Sancho-Sanz, I.; Álvarez-Manzaneda, E.; Oltra, J.E.; Cuerva, J.M. Efficient Propargylation of Aldehydes and Ketones Catalyzed by Titanocene(III). Adv. Synth. Catal. 2009, 351, 2295–2300. [Google Scholar] [CrossRef]
  128. Rosales, A.; Oller-López, J.L.; Justicia, J.; Gansäuer, A.; Oltra, J.E.; Cuerva, J.M. Unprecedented Barbier-type reactions catalysed by titanocene(III). Chem. Commun. 2004, 22, 2628–2629. [Google Scholar] [CrossRef]
  129. Estévez, R.E.; Justicia, J.; Bazdi, B.; Fuentes, N.; Paradas, M.; Choquesillo-Lazarte, D.; García-Ruiz, J.M.; Robles, R.; Gansäuer, A.; Cuerva, J.M.; et al. Ti-Catalyzed Barbier-Type Allylations and Related Reactions. Chem. Eur. J. 2009, 15, 2774–2791. [Google Scholar] [CrossRef]
  130. Kondakov, D.Y.; Negishi, E.I. Zirconium-catalyzed enantioselective methylalumination of monosubstituted alkenes. J. Am. Chem. Soc. 1995, 117, 10771–10772. [Google Scholar] [CrossRef]
  131. Kondakov, D.Y.; Negishi, E.I. Zirconium-Catalyzed Enantioselective Alkylalumination of Monosubstituted Alkenes Proceeding via Noncyclic Mechanism. J. Am. Chem. Soc. 1996, 118, 1577–1578. [Google Scholar] [CrossRef]
  132. Negishi, E.I. Magical Power of Transition Metals: Past, Present, and Future (Nobel Lecture). Angew. Chem. Int. Ed. 2011, 50, 6738–6764. [Google Scholar] [CrossRef] [PubMed]
  133. Petros, R.A.; Camara, J.M.; Norton, J.R. Enantioselective methylalumination of α-olefins. J. Organomet. Chem. 2007, 692, 4768–4773. [Google Scholar] [CrossRef]
  134. Millward, D.B.; Cole, A.P.; Waymouth, R.M. Catalytic Carboalumination of Olefins with Cyclopentadienylamidotitanium Dichloride Complexes. Organometallics 2000, 19, 1870–1878. [Google Scholar] [CrossRef]
  135. Parfenova, L.V.; Berestova, T.V.; Tyumkina, T.V.; Kovyazin, P.V.; Khalilov, L.M.; Whitby, R.J.; Dzhemilev, U.M. Enantioselectivity of chiral zirconocenes as catalysts in alkene hydro-, carbo- and cycloalumination reactions. Tetrahedron Asymmetry 2010, 21, 299–310. [Google Scholar] [CrossRef]
  136. Parfenova, L.V.; Kovyazin, P.V.; Tyumkina, T.V.; Makrushina, A.V.; Khalilov, L.M.; Dzhemilev, U.M. Catalytic enantioselective ethylalumination of terminal alkenes: Substrate effects and absolute configuration assignment. Tetrahedron Asymmetry 2015, 26, 124–135. [Google Scholar] [CrossRef]
  137. Dawson, G.; Durrant, C.A.; Kirk, G.G.; Whitby, R.J.; Jones, R.V.H.; Standen, M.C.H. Zirconium-catalysed enantioselective 2-aluminoethylalumination of alkenes. Tetrahedron Lett. 1997, 38, 2335–2338. [Google Scholar] [CrossRef]
  138. Pino, P.; Cioni, P.; Wei, J. Asymmetric hydrooligomerization of propylene. J. Am. Chem. Soc. 1987, 109, 6189–6191. [Google Scholar] [CrossRef]
  139. Pino, P.; Cioni, P.; Galimberti, M.; Wei, J.; Piccolrovazzi, N. Asymmetric Hydrooligomerization of Propylene. In Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer: Berlin/Heidelberg, Germany, 1988; pp. 269–279. [Google Scholar] [CrossRef]
  140. Pino, P.; Galimberti, M. Asymmetric deuteration and deuteriooligomerization of 1-pentene. J. Organomet. Chem. 1989, 370, 1–7. [Google Scholar] [CrossRef]
  141. Kaminsky, W.; Ahlers, A.; Möller-Lindenhof, N. Asymmetric Oligomerization of Propene and 1-Butene with a Zirconocene/Alumoxane Catalyst. Angew. Chem. Int. Ed. Engl. 1989, 28, 1216–1218. [Google Scholar] [CrossRef]
  142. König, W.A. Eine neue Generation chiraler Trennphasen für die Gaschromatographie. Nachrichten Chem. Tech. Lab. 1989, 37, 471–476. [Google Scholar] [CrossRef]
  143. König, W.A.; Lutz, S.; Wenz, G. Modified Cyclodextrins—Novel, Highly Enantioselective Stationary Phases for Gas Chromatography. Angew. Chem. Int. Ed. Engl. 1988, 27, 979–980. [Google Scholar] [CrossRef]
  144. Kaminsky, W.; Niedoba, S.; Moeller-Lindenhof, N.; Rabe, O. Chapter 4: Isotactic olefin polymerization with optically active catalysts. In Catalysis in Polymer Synthesis; Vandenberg, E.J., Salamone, J.C., Eds.; American Chemical Society: Washington, DC, USA, 1991; pp. 63–71. [Google Scholar] [CrossRef]
  145. Song, W.; Yu, Z.; Chien, J.C.W. Error formation in ansa-zirconocene catalyzed isotactic propylene polymerization. J. Organomet. Chem. 1996, 512, 131–140. [Google Scholar] [CrossRef]
  146. Chien, J.C.W.; Vizzini, J.C.; Kaminsky, W. Difference in stereoselective polymerization of 4-methyl-1-hexene by homogeneous and heterogeneous Ziegler-Natta catalysts. Makromol. Chem. Rapid Commun. 1992, 13, 479–484. [Google Scholar] [CrossRef]
  147. Coates, G.W.; Waymouth, R.M. Enantioselective cyclopolymerization of 1,5-hexadiene catalyzed by chiral zirconocenes: A novel strategy for the synthesis of optically active polymers with chirality in the main chain. J. Am. Chem. Soc. 1993, 115, 91–98. [Google Scholar] [CrossRef]
  148. Galdi, N.; Della Monica, C.; Spinella, A.; Oliva, L. Enantioselective C–C bond formation in styrene dimerization with chiral ansa zirconocene-based catalyst. J. Mol. Catal. A Chem. 2006, 243, 106–110. [Google Scholar] [CrossRef]
  149. Wolczanski, P.T.; Chirik, P.J. A Career in Catalysis: John, E. Bercaw. ACS Catal. 2015, 5, 1747–1757. [Google Scholar] [CrossRef]
  150. Min, E.Y.J.; Byers, J.A.; Bercaw, J.E. Catalyst Site Epimerization during the Kinetic Resolution of Chiral α-Olefins by Polymerization. Organometallics 2008, 27, 2179–2188. [Google Scholar] [CrossRef]
  151. Byers, J.A.; Bercaw, J.E. Kinetic resolution of racemic α-olefins with ansa-zirconocene polymerization catalysts: Enantiomorphic site vs. chain end control. Proc. Natl. Acad. Sci. USA 2006, 103, 15303–15308. [Google Scholar] [CrossRef]
  152. Ota, Y.; Murayama, T.; Nozaki, K. One-step catalytic asymmetric synthesis of all-syn deoxypropionate motif from propylene: Total synthesis of (2R,4R,6R,8R)-2,4,6,8-tetramethyldecanoic acid. Proc. Natl. Acad. Sci. USA 2016, 113, 2857–2861. [Google Scholar] [CrossRef] [PubMed]
  153. Murayama, T.; Nozaki, K. Single-Step Construction of the anti-Deoxypropionate Motif from Propylene: Formal Total Synthesis of the Cuticular Hydrocarbons Isolated from Antitrogus parvulus. Angew. Chem. Int. Ed. 2018, 57, 11394–11398. [Google Scholar] [CrossRef] [PubMed]
  154. Kovyazin, P.V.; Mukhamadeeva, O.V.; Parfenova, L.V. Stereoselective synthesis of functionalized hexene oligomers catalyzed by chiral ansa-zirconocene in the presence of Al- and B-containing activators. Mendeleev Commun. 2024, 34, 61–63. [Google Scholar] [CrossRef]
  155. Corradini, P.; Guerra, G.; Cavallo, L. Do New Century Catalysts Unravel the Mechanism of Stereocontrol of Old Ziegler−Natta Catalysts? Acc. Chem. Res. 2004, 37, 231–241. [Google Scholar] [CrossRef]
  156. Gately, D.A.; Norton, J.R. Origin of Stereochemistry in the α-Amino Acid Esters and Amides Generated from Optically Active Zirconaaziridine Complexes. J. Am. Chem. Soc. 1996, 118, 3479–3489. [Google Scholar] [CrossRef]
  157. Rodewald, S.; Jordan, R.F. Stereoselective Olefin Insertion Reactions of Chiral (EBI)Zr(.eta.2-pyrid-2-yl)+ and (EBTHI)Zr(.eta.2-pyrid-2-yl)+ Complexes. J. Am. Chem. Soc. 1994, 116, 4491–4492. [Google Scholar] [CrossRef]
  158. Sweeney, Z.K.; Salsman, J.L.; Andersen, R.A.; Bergman, R.G. Synthesis of Chiral, Enantiopure Zirconocene Imido Complexes: Highly Selective Kinetic Resolution and Stereoinversion of Allenes, and Evidence for a Stepwise Cycloaddition/Retrocycloaddition Reaction Mechanism. Angew. Chem. Int. Ed. 2000, 39, 2339–2343. [Google Scholar] [CrossRef]
  159. Hicks, F.A.; Kablaoui, N.M.; Buchwald, S.L. Titanocene-Catalyzed Cyclocarbonylation of Enynes to Cyclopentenones. J. Am. Chem. Soc. 1996, 118, 9450–9451. [Google Scholar] [CrossRef]
  160. Hicks, F.A.; Buchwald, S.L. Highly Enantioselective Catalytic Pauson−Khand Type Formation of Bicyclic Cyclopentenones. J. Am. Chem. Soc. 1996, 118, 11688–11689. [Google Scholar] [CrossRef]
  161. Sturla, S.J.; Buchwald, S.L. Catalytic Asymmetric Cyclocarbonylation of Nitrogen-Containing Enynes. J. Org. Chem. 1999, 64, 5547–5550. [Google Scholar] [CrossRef]
  162. Streuff, J.; Feurer, M.; Bichovski, P.; Frey, G.; Gellrich, U. Enantioselective Titanium(III)-Catalyzed Reductive Cyclization of Ketonitriles. Angew. Chem. Int. Ed. 2012, 51, 8661–8664. [Google Scholar] [CrossRef] [PubMed]
  163. Larson, G.L.; Liberatore, R.J. Organosilanes in Metal-Catalyzed, Enantioselective Reductions. Org. Process Res. Dev. 2021, 25, 1719–1787. [Google Scholar] [CrossRef]
  164. Carter, M.B.; Schiott, B.; Gutiérrez, A.; Buchwald, S.L. Enantioselective Hydrosilylation of Ketones with a Chiral Titanocene Catalyst. J. Am. Chem. Soc. 1994, 116, 11667–11670. [Google Scholar] [CrossRef]
  165. Rahimian, K.; Harrod, J.F. The influence of the catalyst preparation protocol and silane structure on the rate and enantioselectivity of ansa-titanocene catalysed hydrosilation of prochiral ketones. Inorg. Chim. Acta 1998, 270, 330–336. [Google Scholar] [CrossRef]
  166. Yun, J.; Buchwald, S.L. Titanocene-Catalyzed Asymmetric Ketone Hydrosilylation:  The Effect of Catalyst Activation Protocol and Additives on the Reaction Rate and Enantioselectivity. J. Am. Chem. Soc. 1999, 121, 5640–5644. [Google Scholar] [CrossRef]
  167. Hong, L.; Sun, W.; Yang, D.; Li, G.; Wang, R. Additive Effects on Asymmetric Catalysis. Chem. Rev. 2016, 116, 4006–4123. [Google Scholar] [CrossRef]
  168. Beagley, P.; Davies, P.J.; Blacker, A.J.; White, C. Chiral Metallocenes. 3.1 The Enantioselective Reduction of CX Bonds (X = O or CH2) by the Chiral ansa-Metallocenes (R)- or (S)-[TiCl255- C5Me4SiMe2C5H3R*)] (R* = Menthyl or Neomenthyl). Organometallics 2002, 21, 5852–5858. [Google Scholar] [CrossRef]
  169. Halterman, R.L.; Ramsey, T.M.; Chen, Z. Catalytic Asymmetric Hydrosilation of Aryl Alkyl Ketones with C2-Symmetric Chiral Metallocene Complexes. J. Org. Chem. 1994, 59, 2642–2644. [Google Scholar] [CrossRef]
  170. Willoughby, C.A.; Buchwald, S.L. Asymmetric titanocene-catalyzed hydrogenation of imines. J. Am. Chem. Soc. 1992, 114, 7562–7564. [Google Scholar] [CrossRef]
  171. Willoughby, C.A.; Buchwald, S.L. Synthesis of highly enantiomerically enriched cyclic amines by the catalytic asymmetric hydrogenation of cyclic imines. J. Org. Chem. 1993, 58, 7627–7629. [Google Scholar] [CrossRef]
  172. Willoughby, C.A.; Buchwald, S.L. Catalytic Asymmetric Hydrogenation of Imines with a Chiral Titanocene Catalyst: Scope and Limitations. J. Am. Chem. Soc. 1994, 116, 8952–8965. [Google Scholar] [CrossRef]
  173. Willoughby, C.A.; Buchwald, S.L. Catalytic Asymmetric Hydrogenation of Imines with a Chiral Titanocene Catalyst: Kinetic and Mechanistic Investigations. J. Am. Chem. Soc. 1994, 116, 11703–11714. [Google Scholar] [CrossRef]
  174. Lee, N.E.; Buchwald, S.L. Asymmetric Hydrogenation of Enamines with a Chiral Titanocene Catalyst. J. Am. Chem. Soc. 1994, 116, 5985–5986. [Google Scholar] [CrossRef]
  175. Verdaguer, X.; Lange, U.E.W.; Reding, M.T.; Buchwald, S.L. Highly Enantioselective Imine Hydrosilylation Using (S,S)-Ethylenebis(η5-tetrahydroindenyl)titanium Difluoride. J. Am. Chem. Soc. 1996, 118, 6784–6785. [Google Scholar] [CrossRef]
  176. Hansen, M.C.; Buchwald, S.L. A Method for the Asymmetric Hydrosilylation of N-Aryl Imines. Org. Lett. 2000, 2, 713–715. [Google Scholar] [CrossRef]
  177. Verdaguer, X.; Lange, U.E.W.; Buchwald, S.L. Amine Additives Greatly Expand the Scope of Asymmetric Hydrosilylation of Imines. Angew. Chem. Int. Ed. 1998, 37, 1103–1107. [Google Scholar] [CrossRef]
  178. Reding, M.T.; Buchwald, S.L. Short Enantioselective Total Syntheses of the Piperidine Alkaloids (S)-Coniine and (2R,6R)-trans-Solenopsin A via Catalytic Asymmetric Imine Hydrosilylation. J. Org. Chem. 1998, 63, 6344–6347. [Google Scholar] [CrossRef]
  179. Hansen, M.C.; Buchwald, S.L. Synthesis of NPS R-568 utilizing titanium-catalyzed asymmetric hydrosilylation. Tetrahedron Lett. 1999, 40, 2033–2034. [Google Scholar] [CrossRef]
  180. Viso, A.; Lee, N.E.; Buchwald, S.L. Kinetic resolution of racemic disubstituted 1-pyrrolines via asymmetric reduction with a chiral titanocene catalyst. J. Am. Chem. Soc. 1994, 116, 9373–9374. [Google Scholar] [CrossRef]
  181. Yun, J.; Buchwald, S.L. Kinetic resolution and isomerization of 2,5-disubstituted pyrrolines. Chirality 2000, 12, 476–478. [Google Scholar] [CrossRef]
  182. Yun, J.; Buchwald, S.L. Efficient Kinetic Resolution in the Asymmetric Hydrosilylation of Imines of 3-Substituted Indanones and 4-Substituted Tetralones. J. Org. Chem. 2000, 65, 767–774. [Google Scholar] [CrossRef] [PubMed]
  183. Fu, P.F.; Brard, L.; Li, Y.; Marks, T.J. Regioselection and Enantioselection in Organolanthanide-Catalyzed Olefin Hydrosilylation. A Kinetic and Mechanistic Study. J. Am. Chem. Soc. 1995, 117, 7157–7168. [Google Scholar] [CrossRef]
  184. Cui, X.; Burgess, K. Catalytic Homogeneous Asymmetric Hydrogenations of Largely Unfunctionalized Alkenes. Chem. Rev. 2005, 105, 3272–3296. [Google Scholar] [CrossRef] [PubMed]
  185. Verendel, J.J.; Pàmies, O.; Diéguez, M.; Andersson, P.G. Asymmetric Hydrogenation of Olefins Using Chiral Crabtree-type Catalysts: Scope and Limitations. Chem. Rev. 2014, 114, 2130–2169. [Google Scholar] [CrossRef]
  186. Pizzano, A. Chapter One—Asymmetric hydrogenation of functionalized olefins. In Advances in Catalysis. Metal-Catalyzed Asymmetric Hydrogenation: Evolution and Prospect; Elsevier: Cambridge, MA, USA, 2021; Volume 68, pp. 1–134. [Google Scholar] [CrossRef]
  187. Pàmies, O.; Zheng, J.; Faiges, J.; Andersson, P.G. Chapter Two—Asymmetric hydrogenation of unfunctionalized olefins or with poorly coordinative groups. In Advances in Catalysis. Metal-Catalyzed Asymmetric Hydrogenation: Evolution and Prospect; Elsevier: Cambridge, MA, USA, 2021; Volume 68, pp. 135–203. [Google Scholar] [CrossRef]
  188. Waymouth, R.; Pino, P. Enantioselective hydrogenation of olefins with homogeneous Ziegler-Natta catalysts. J. Am. Chem. Soc. 1990, 112, 4911–4914. [Google Scholar] [CrossRef]
  189. Grossman, R.B.; Doyle, R.A.; Buchwald, S.L. Syntheses of [ethylene-1,2-bis(.eta.5-4,5,6,7-tetrahydro-1-indenyl)]zirconium and -hafnium hydride complexes. Improved syntheses of the corresponding dichlorides. Organometallics 1991, 10, 1501–1505. [Google Scholar] [CrossRef]
  190. Broene, R.D.; Buchwald, S.L. Asymmetric Hydrogenation of Unfunctionalized Trisubstituted Olefins with a Chiral Titanocene Catalyst. J. Am. Chem. Soc. 1996, 118, 9458. [Google Scholar] [CrossRef]
  191. Troutman, M.V.; Appella, D.H.; Buchwald, S.L. Asymmetric Hydrogenation of Unfunctionalized Tetrasubstituted Olefins with a Cationic Zirconocene Catalyst. J. Am. Chem. Soc. 1999, 121, 4916–4917. [Google Scholar] [CrossRef]
  192. Yamaura, Y.; Hyakutake, M.; Mori, M. Synthesis of Heterocycles Using Zirconium-Catalyzed Asymmetric Diene Cyclization. J. Am. Chem. Soc. 1997, 119, 7615–7616. [Google Scholar] [CrossRef]
  193. Roesky, P.W.; Stern, C.L.; Marks, T.J. Ancillary Ligand Effects on Organo-f-Element Reactivity. ansa-Metallocenes with Bridge-Tethered Donors. Organometallics 1997, 16, 4705–4711. [Google Scholar] [CrossRef]
  194. Gagné, M.R.; Brard, L.; Conticello, V.P.; Giardello, M.A.; Stern, C.L.; Marks, T.J. Stereoselection effects in the catalytic hydroamination/cyclization of amino olefins at chiral organolanthanide centers. Organometallics 1992, 11, 2003–2005. [Google Scholar] [CrossRef]
  195. Gagné, M.R.; Stern, C.L.; Marks, T.J. Organolanthanide-catalyzed hydroamination. A kinetic, mechanistic, and diastereoselectivity study of the cyclization of N-unprotected amino olefins. J. Am. Chem. Soc. 1992, 114, 275–294. [Google Scholar] [CrossRef]
  196. Hong, S.; Marks, T.J. Highly Stereoselective Intramolecular Hydroamination/Cyclization of Conjugated Aminodienes Catalyzed by Organolanthanides. J. Am. Chem. Soc. 2002, 124, 7886–7887. [Google Scholar] [CrossRef]
  197. Ryu, J.S.; Marks, T.J.; McDonald, F.E. Organolanthanide-Catalyzed Intramolecular Hydroamination/Cyclization/Bicyclization of Sterically Encumbered Substrates. Scope, Selectivity, and Catalyst Thermal Stability for Amine-Tethered Unactivated 1,2-Disubstituted Alkenes. J. Org. Chem. 2004, 69, 1038–1052. [Google Scholar] [CrossRef]
  198. Hong, S.; Marks, T.J. Organolanthanide-Catalyzed Hydroamination. Acc. Chem. Res. 2004, 37, 673–686. [Google Scholar] [CrossRef]
  199. Hong, S.; Kawaoka, A.M.; Marks, T.J. Intramolecular Hydroamination/Cyclization of Conjugated Aminodienes Catalyzed by Organolanthanide Complexes. Scope, Diastereo- and Enantioselectivity, and Reaction Mechanism. J. Am. Chem. Soc. 2003, 125, 15878–15892. [Google Scholar] [CrossRef]
  200. Colletti, S.L.; Halterman, R.L. Asymmetric epoxidation of unfunctionalized alkenes using the new C2-symmetrical 1,1′-binaphthyl-2,2′-dimethylene-bridged ansa-Bis(1-indenyl)titanium dichloride catalyst. Tetrahedron Lett. 1992, 33, 1005–1008. [Google Scholar] [CrossRef]
  201. Gansäuer, A.; Bluhm, H.; Rinker, B.; Narayan, S.; Schick, M.; Lauterbach, T.; Pierobon, M. Reagent-Controlled Stereoselectivity in Titanocene-Catalyzed Epoxide Openings: Reductions and Intermolecular Additions to α,β-Unsaturated Carbonyl Compounds. Chem. Eur. J. 2003, 9, 531–542. [Google Scholar] [CrossRef]
  202. Gansäuer, A.; Bluhm, H.; Lauterbach, T. Titanocene-Catalysed Enantioselective Opening of meso-Epoxides. Adv. Synth. Catal. 2001, 343, 785–787. [Google Scholar] [CrossRef]
  203. Gansäuer, A.; Lauterbach, T.; Bluhm, H.; Noltemeyer, M. A Catalytic Enantioselective Electron Transfer Reaction: Titanocene-Catalyzed Enantioselective Formation of Radicals from meso-Epoxides. Angew. Chem. Int. Ed. 1999, 38, 2909–2910. [Google Scholar] [CrossRef]
  204. Estévez, R.E.; Oller-Lopez, J.L.; Robles, R.; Melgarejo, C.R.; Gansäuer, A.; Cuerva, J.M.; Oltra, J.E. Stereocontrolled coupling between aldehydes and conjugated alkenals mediated by TiIII/H2O. Org. Lett. 2006, 8, 5433–5436. [Google Scholar] [CrossRef] [PubMed]
  205. Chen, Z.; Halterman, R.L. Enantioselective catalytic isomerization of an unfunctionalized achiral alkene. J. Am. Chem. Soc. 1992, 114, 2276–2277. [Google Scholar] [CrossRef]
Scheme 1. Pathways for the synthesis of enantiomerically pure ansa-complexes ([aux]*—chiral auxiliary).
Scheme 1. Pathways for the synthesis of enantiomerically pure ansa-complexes ([aux]*—chiral auxiliary).
Molecules 30 02511 sch001
Scheme 2. Derivatization of Ti complexes rac-2a,b using S-binaphthol.
Scheme 2. Derivatization of Ti complexes rac-2a,b using S-binaphthol.
Molecules 30 02511 sch002
Scheme 3. Derivatization of Zr complexes 5 and 6.
Scheme 3. Derivatization of Zr complexes 5 and 6.
Molecules 30 02511 sch003
Scheme 4. Derivatization of Ti bis-tetrahydroindenyl complex with S-binaphthol.
Scheme 4. Derivatization of Ti bis-tetrahydroindenyl complex with S-binaphthol.
Molecules 30 02511 sch004
Scheme 5. Resolution of bis-tetrahydroindenyl Ti and Zr complexes with R-acetylmandelic acid.
Scheme 5. Resolution of bis-tetrahydroindenyl Ti and Zr complexes with R-acetylmandelic acid.
Molecules 30 02511 sch005
Scheme 6. Derivatization of Ti complex rac-11 using S-binaphthol.
Scheme 6. Derivatization of Ti complex rac-11 using S-binaphthol.
Molecules 30 02511 sch006
Scheme 7. Synthesis of enantiomerically pure Zr complexes p-S,p-S-17 and p-S,p-S-18.
Scheme 7. Synthesis of enantiomerically pure Zr complexes p-S,p-S-17 and p-S,p-S-18.
Molecules 30 02511 sch007
Scheme 8. Resolution of racemic Ti (20a) and Zr (21a) complexes with R-binaphtholate and acetylmandelic acid.
Scheme 8. Resolution of racemic Ti (20a) and Zr (21a) complexes with R-binaphtholate and acetylmandelic acid.
Molecules 30 02511 sch008
Scheme 9. Derivatization of complexes rac-11 (Ti) and rac-13 (Zr) with a mixture of R-binaphthol and p-aminobenzoic acid.
Scheme 9. Derivatization of complexes rac-11 (Ti) and rac-13 (Zr) with a mixture of R-binaphthol and p-aminobenzoic acid.
Molecules 30 02511 sch009
Scheme 10. Separation of metallocenes rac-13 (Zr) and rac-28 (Hf) using chiral HPLC.
Scheme 10. Separation of metallocenes rac-13 (Zr) and rac-28 (Hf) using chiral HPLC.
Molecules 30 02511 sch010
Scheme 11. Synthesis of enantiomerically pure triflates of ansa-metallocenes Zr (29) and Ti (30).
Scheme 11. Synthesis of enantiomerically pure triflates of ansa-metallocenes Zr (29) and Ti (30).
Molecules 30 02511 sch011
Scheme 12. Synthesis of Ti (p-S,p-S-34) and Zr (p-S,p-S-35) complexes starting from enantiomerically pure ligands.
Scheme 12. Synthesis of Ti (p-S,p-S-34) and Zr (p-S,p-S-35) complexes starting from enantiomerically pure ligands.
Molecules 30 02511 sch012
Scheme 13. Synthesis of optically active titanocene 38.
Scheme 13. Synthesis of optically active titanocene 38.
Molecules 30 02511 sch013
Scheme 14. Synthesis of enantiomerically enriched Ti and Zr complexes from optically active Br-substituted binaphthyl precursors.
Scheme 14. Synthesis of enantiomerically enriched Ti and Zr complexes from optically active Br-substituted binaphthyl precursors.
Molecules 30 02511 sch014
Scheme 15. Synthesis and purification of metallocenes 4749.
Scheme 15. Synthesis and purification of metallocenes 4749.
Molecules 30 02511 sch015
Scheme 16. Synthesis of enantiomerically pure titanocene 52.
Scheme 16. Synthesis of enantiomerically pure titanocene 52.
Molecules 30 02511 sch016
Scheme 17. Stereoselective synthesis of titanocene complexes p-S,p-S-58 and p-S,p-S-59.
Scheme 17. Stereoselective synthesis of titanocene complexes p-S,p-S-58 and p-S,p-S-59.
Molecules 30 02511 sch017
Scheme 18. Synthesis of optically active meso-isomers of complexes 62ac and 63ac.
Scheme 18. Synthesis of optically active meso-isomers of complexes 62ac and 63ac.
Molecules 30 02511 sch018
Scheme 19. Synthesis of optically active Zr complex meso-66.
Scheme 19. Synthesis of optically active Zr complex meso-66.
Molecules 30 02511 sch019
Scheme 20. Stereoselective synthesis of Zr complex p-R,p-R-68.
Scheme 20. Stereoselective synthesis of Zr complex p-R,p-R-68.
Molecules 30 02511 sch020
Scheme 21. Polycyclic compounds as precursors in the synthesis of chiral titanocenes 7375.
Scheme 21. Polycyclic compounds as precursors in the synthesis of chiral titanocenes 7375.
Molecules 30 02511 sch021
Scheme 22. Stabilization of ligand conformation as a significant factor in the stereoselective synthesis of chiral complexes.
Scheme 22. Stabilization of ligand conformation as a significant factor in the stereoselective synthesis of chiral complexes.
Molecules 30 02511 sch022
Scheme 23. Bis-amide R,R-82 in the synthesis of Ti complexes 47AC.
Scheme 23. Bis-amide R,R-82 in the synthesis of Ti complexes 47AC.
Molecules 30 02511 sch023
Scheme 24. Effect of menthyl and neomenthyl substituent in Cp-ring on stereoselectivity of metal complex synthesis.
Scheme 24. Effect of menthyl and neomenthyl substituent in Cp-ring on stereoselectivity of metal complex synthesis.
Molecules 30 02511 sch024
Scheme 25. Synthesis of mixed menthyl-substituted cyclopentadienyl-octahydrofluorenyl ansa-complexes 93, 94.
Scheme 25. Synthesis of mixed menthyl-substituted cyclopentadienyl-octahydrofluorenyl ansa-complexes 93, 94.
Molecules 30 02511 sch025
Scheme 26. Synthesis of diastereomeric complexes 98.
Scheme 26. Synthesis of diastereomeric complexes 98.
Molecules 30 02511 sch026
Scheme 27. Synthesis of indenyl-fluorenyl ansa-zirconocene with a naphthyl bridge p-S-103.
Scheme 27. Synthesis of indenyl-fluorenyl ansa-zirconocene with a naphthyl bridge p-S-103.
Molecules 30 02511 sch027
Scheme 28. Synthesis of complexes 109112.
Scheme 28. Synthesis of complexes 109112.
Molecules 30 02511 sch028
Scheme 29. Synthesis of enantiomerically pure C1-symmetric complexes S-116ac.
Scheme 29. Synthesis of enantiomerically pure C1-symmetric complexes S-116ac.
Molecules 30 02511 sch029
Scheme 30. Effect of σ-ligand on the diastereoselectivity in the synthesis of complexes 122 and 123.
Scheme 30. Effect of σ-ligand on the diastereoselectivity in the synthesis of complexes 122 and 123.
Molecules 30 02511 sch030
Scheme 31. Preparation of enantiomerically pure ansa-complexes of group IIIB metals 127139.
Scheme 31. Preparation of enantiomerically pure ansa-complexes of group IIIB metals 127139.
Molecules 30 02511 sch031
Scheme 32. Synthesis of chiral C1-symmetric organolanthanide complexes S-140ac and S-141ac.
Scheme 32. Synthesis of chiral C1-symmetric organolanthanide complexes S-140ac and S-141ac.
Molecules 30 02511 sch032
Scheme 33. Application of enantiomerically pure binaphthol-bridged ligands in the synthesis of ansa-yttrocenes.
Scheme 33. Application of enantiomerically pure binaphthol-bridged ligands in the synthesis of ansa-yttrocenes.
Molecules 30 02511 sch033
Scheme 34. R,R-enantiomer of Zr amide as a precursor in the synthesis of enantiomerically pure ansa-complex p-S,p-S-151.
Scheme 34. R,R-enantiomer of Zr amide as a precursor in the synthesis of enantiomerically pure ansa-complex p-S,p-S-151.
Molecules 30 02511 sch034
Scheme 35. Enantioselective synthesis of Zr complex 13 via the association of ZrCl4 with enantiomerically pure S- or R-binaphthol ethers.
Scheme 35. Enantioselective synthesis of Zr complex 13 via the association of ZrCl4 with enantiomerically pure S- or R-binaphthol ethers.
Molecules 30 02511 sch035
Scheme 36. Complex p-R,p-R-13 as a stereoselective catalyst for alkene carbomagnesiation.
Scheme 36. Complex p-R,p-R-13 as a stereoselective catalyst for alkene carbomagnesiation.
Molecules 30 02511 sch036
Scheme 37. Carbomagnesation of heterocyclic alkenes catalyzed by enantiomerically pure complexes p-R,p-R-13 and R,p-R,p-R-16.
Scheme 37. Carbomagnesation of heterocyclic alkenes catalyzed by enantiomerically pure complexes p-R,p-R-13 and R,p-R,p-R-16.
Molecules 30 02511 sch037
Scheme 38. Carbomagnesation of rac-5,6-dihydropyrans, catalyzed with p-R,p-R-13.
Scheme 38. Carbomagnesation of rac-5,6-dihydropyrans, catalyzed with p-R,p-R-13.
Molecules 30 02511 sch038
Scheme 39. Kinetic effect of p-R,p-R-13 in the ethylmagnesiation of pyrans.
Scheme 39. Kinetic effect of p-R,p-R-13 in the ethylmagnesiation of pyrans.
Molecules 30 02511 sch039
Scheme 40. Kinetic resolution of dihydrofurans in the presence of p-R,p-R-13.
Scheme 40. Kinetic resolution of dihydrofurans in the presence of p-R,p-R-13.
Molecules 30 02511 sch040
Scheme 41. Ethylmagnesiation of seven-membered cycloolefins catalyzed by p-R,p-R-16 or R,p-R,p-R-27.
Scheme 41. Ethylmagnesiation of seven-membered cycloolefins catalyzed by p-R,p-R-16 or R,p-R,p-R-27.
Molecules 30 02511 sch041
Scheme 42. Carbomagnesiation of cyclic alkoxy-substituted olefins 166av catalyzed by R,p-R,p-R-16 or S,p-S,p-S-16.
Scheme 42. Carbomagnesiation of cyclic alkoxy-substituted olefins 166av catalyzed by R,p-R,p-R-16 or S,p-S,p-S-16.
Molecules 30 02511 sch042
Scheme 43. Carbomagnesiation of inactivated alkenes by MgEt2 in the presence of R,p-R,p-R-16.
Scheme 43. Carbomagnesiation of inactivated alkenes by MgEt2 in the presence of R,p-R,p-R-16.
Molecules 30 02511 sch043
Scheme 44. Synthesis of homoallylic alcohols by the reaction of aldehydes with crotyl titanocenes.
Scheme 44. Synthesis of homoallylic alcohols by the reaction of aldehydes with crotyl titanocenes.
Molecules 30 02511 sch044
Scheme 45. Reaction of allyl- (171) and crotyltitanocenes (172) with aldehydes.
Scheme 45. Reaction of allyl- (171) and crotyltitanocenes (172) with aldehydes.
Molecules 30 02511 sch045
Scheme 46. Carbometalation of aldehydes catalyzed by complex p-S,p-S-4b.
Scheme 46. Carbometalation of aldehydes catalyzed by complex p-S,p-S-4b.
Molecules 30 02511 sch046
Scheme 47. Dia- and enantioselectivity of titanocenes 52, 58, and 59 in the reactions of 2-alkyl-1,3-butadienes with CO2 or R’CHO.
Scheme 47. Dia- and enantioselectivity of titanocenes 52, 58, and 59 in the reactions of 2-alkyl-1,3-butadienes with CO2 or R’CHO.
Molecules 30 02511 sch047
Scheme 48. Barbier-type carbonyl compound propargylation or allylation catalyzed with p-S,p-S-11.
Scheme 48. Barbier-type carbonyl compound propargylation or allylation catalyzed with p-S,p-S-11.
Molecules 30 02511 sch048
Scheme 49. Asymmetric Zr and Ti-catalyzed enantioselective carboalumination and cycloalumination of alkenes.
Scheme 49. Asymmetric Zr and Ti-catalyzed enantioselective carboalumination and cycloalumination of alkenes.
Molecules 30 02511 sch049
Scheme 50. Mechanism of alkene methylalumination in the presence of enantiomerically pure ansa-zirconocene complexes.
Scheme 50. Mechanism of alkene methylalumination in the presence of enantiomerically pure ansa-zirconocene complexes.
Molecules 30 02511 sch050
Scheme 51. Reaction of 2,5-dihydrofurans with AlEt3 catalyzed by complexes p-R,p-R-13 or R,p-R,p-R-16.
Scheme 51. Reaction of 2,5-dihydrofurans with AlEt3 catalyzed by complexes p-R,p-R-13 or R,p-R,p-R-16.
Molecules 30 02511 sch051
Scheme 52. Propene polymerization in the presence of a catalytic system based on p-R,p-R-17 and MAO.
Scheme 52. Propene polymerization in the presence of a catalytic system based on p-R,p-R-17 and MAO.
Molecules 30 02511 sch052
Scheme 53. Polymerization of 1-pentene catalyzed by R,p-R,p-R-16 or p-R,p-R-17 and MAO, followed by treatment with D2.
Scheme 53. Polymerization of 1-pentene catalyzed by R,p-R,p-R-16 or p-R,p-R-17 and MAO, followed by treatment with D2.
Molecules 30 02511 sch053
Scheme 54. Oligomerization of propene and 1-butene catalyzed by R,R,p-S,p-S-15 and MAO.
Scheme 54. Oligomerization of propene and 1-butene catalyzed by R,R,p-S,p-S-15 and MAO.
Molecules 30 02511 sch054
Scheme 55. Cyclopolymerization of 1,5-hexadiene catalyzed by achiral precatalysts Cp2MX2 (M = Ti, Zr; X = Cl, Me) and MAO.
Scheme 55. Cyclopolymerization of 1,5-hexadiene catalyzed by achiral precatalysts Cp2MX2 (M = Ti, Zr; X = Cl, Me) and MAO.
Molecules 30 02511 sch055
Scheme 56. Synthesis of enantiomerically enriched styrene hydrooligomers catalyzed by complex p-R,p-R-16.
Scheme 56. Synthesis of enantiomerically enriched styrene hydrooligomers catalyzed by complex p-R,p-R-16.
Molecules 30 02511 sch056
Scheme 57. Kinetic resolution of racemic α-olefins bearing bulky substituents during polymerization catalyzed by enantiomerically pure S-116ad.
Scheme 57. Kinetic resolution of racemic α-olefins bearing bulky substituents during polymerization catalyzed by enantiomerically pure S-116ad.
Molecules 30 02511 sch057
Scheme 58. Stereoselectivity control model of propene polymerization catalyzed by enantiomerically pure S-116ad.
Scheme 58. Stereoselectivity control model of propene polymerization catalyzed by enantiomerically pure S-116ad.
Molecules 30 02511 sch058
Scheme 59. Stereocontrolled oligomerization of propylene via reaction with ZnEt2 in the presence of p-S,p-S-13 and MAO.
Scheme 59. Stereocontrolled oligomerization of propylene via reaction with ZnEt2 in the presence of p-S,p-S-13 and MAO.
Molecules 30 02511 sch059
Scheme 60. Oligomerization of 1-hexene catalyzed by enantiomerically pure Zr ansa-complexes.
Scheme 60. Oligomerization of 1-hexene catalyzed by enantiomerically pure Zr ansa-complexes.
Molecules 30 02511 sch060
Scheme 61. Cycloaddition of enones with cyclopentadiene catalyzed by chiral Ti and Zr complexes.
Scheme 61. Cycloaddition of enones with cyclopentadiene catalyzed by chiral Ti and Zr complexes.
Molecules 30 02511 sch061
Scheme 62. Synthesis of zirconaaziridines from dimethyl complex p-S,p-S-17.
Scheme 62. Synthesis of zirconaaziridines from dimethyl complex p-S,p-S-17.
Molecules 30 02511 sch062
Scheme 63. Asymmetric induction in the reaction of amines with alkynes, alkenes, or aldehydes catalyzed by zirconaaziridines 210.
Scheme 63. Asymmetric induction in the reaction of amines with alkynes, alkenes, or aldehydes catalyzed by zirconaaziridines 210.
Molecules 30 02511 sch063
Scheme 64. Reactions of complexes 214 with various alkenes yielding metallocycles 215ad and synthesis of substituted pyridine 216 catalyzed by p-S,p-S-17.
Scheme 64. Reactions of complexes 214 with various alkenes yielding metallocycles 215ad and synthesis of substituted pyridine 216 catalyzed by p-S,p-S-17.
Molecules 30 02511 sch064
Scheme 65. Reactions of alkyl- or aryl-substituted zirconaziridine with ethylene carbonate or isocyanates yielding amino acid esters.
Scheme 65. Reactions of alkyl- or aryl-substituted zirconaziridine with ethylene carbonate or isocyanates yielding amino acid esters.
Molecules 30 02511 sch065
Scheme 66. Proposed mechanism of stereoinduction in the reaction of substituted zirconaziridine with ethylene carbonate.
Scheme 66. Proposed mechanism of stereoinduction in the reaction of substituted zirconaziridine with ethylene carbonate.
Molecules 30 02511 sch066
Scheme 67. Reaction of the enantiomerically enriched imido-complex p-S,p-S-222 with allenes.
Scheme 67. Reaction of the enantiomerically enriched imido-complex p-S,p-S-222 with allenes.
Molecules 30 02511 sch067
Scheme 68. Pauson–Khand cyclo-condensation of enynes with carbon monoxide catalyzed with p-S,p-S-226.
Scheme 68. Pauson–Khand cyclo-condensation of enynes with carbon monoxide catalyzed with p-S,p-S-226.
Molecules 30 02511 sch068
Scheme 69. Asymmetric cyclocarbonylation of nitrogen-containing enynes catalyzed with p-S,p-S-226.
Scheme 69. Asymmetric cyclocarbonylation of nitrogen-containing enynes catalyzed with p-S,p-S-226.
Molecules 30 02511 sch069
Scheme 70. Intramolecular asymmetric reductive coupling of ketones with nitriles catalyzed by p-R,p-R-11.
Scheme 70. Intramolecular asymmetric reductive coupling of ketones with nitriles catalyzed by p-R,p-R-11.
Molecules 30 02511 sch070
Scheme 71. Asymmetric hydrosilylation of ketones by polymeric methylhydrosiloxane Me3SiO[MeSi(H)O]nSiMe3 in the presence of R,p-R,p-R-12.
Scheme 71. Asymmetric hydrosilylation of ketones by polymeric methylhydrosiloxane Me3SiO[MeSi(H)O]nSiMe3 in the presence of R,p-R,p-R-12.
Molecules 30 02511 sch071
Scheme 72. Hydrosilylation of ketones by silanes catalyzed with alkylated S,p-S,p-S-12.
Scheme 72. Hydrosilylation of ketones by silanes catalyzed with alkylated S,p-S,p-S-12.
Molecules 30 02511 sch072
Scheme 73. Titanocene p-R,p-R-235 as an effective catalyst for the hydrosilylation of ketones by PhSiH3.
Scheme 73. Titanocene p-R,p-R-235 as an effective catalyst for the hydrosilylation of ketones by PhSiH3.
Molecules 30 02511 sch073
Scheme 74. Mechanism of titanocene-catalyzed hydrosilylation of ketones by PhSiH3.
Scheme 74. Mechanism of titanocene-catalyzed hydrosilylation of ketones by PhSiH3.
Molecules 30 02511 sch074
Scheme 75. Reduction of aryl-substituted ketones by silanes catalyzed by Ti complexes S-88a, R-88a, p-S,p-S-38, and 1R,2R,4R,5R-53.
Scheme 75. Reduction of aryl-substituted ketones by silanes catalyzed by Ti complexes S-88a, R-88a, p-S,p-S-38, and 1R,2R,4R,5R-53.
Molecules 30 02511 sch075
Scheme 76. Asymmetric hydrogenation of imines catalyzed by R,p-R,p-R-12 and R,R,R-22.
Scheme 76. Asymmetric hydrogenation of imines catalyzed by R,p-R,p-R-12 and R,R,R-22.
Molecules 30 02511 sch076
Scheme 77. Hydrogenation of 1,1-disubstituted enamines catalyzed with R,p-R,p-R-12.
Scheme 77. Hydrogenation of 1,1-disubstituted enamines catalyzed with R,p-R,p-R-12.
Molecules 30 02511 sch077
Scheme 78. Asymmetric hydrosilylation of N-aryl-substituted imines using PhSiH3 or PMHS in the presence of Ti complex p-S,p-S-235.
Scheme 78. Asymmetric hydrosilylation of N-aryl-substituted imines using PhSiH3 or PMHS in the presence of Ti complex p-S,p-S-235.
Molecules 30 02511 sch078
Scheme 79. Hydrosilylation of rac-2,5-disubstituted pyrroles in the presence of the binaphtholate complex S,p-S,p-S-12.
Scheme 79. Hydrosilylation of rac-2,5-disubstituted pyrroles in the presence of the binaphtholate complex S,p-S,p-S-12.
Molecules 30 02511 sch079
Scheme 80. Kinetic resolution of N-alkyl imines derived from 3-substituted indanones and 4-substituted tetralones.
Scheme 80. Kinetic resolution of N-alkyl imines derived from 3-substituted indanones and 4-substituted tetralones.
Molecules 30 02511 sch080
Scheme 81. Hydrosilylation of phenylbutene by PhSiH3 in the presence of samarium complexes R- or S-137a.
Scheme 81. Hydrosilylation of phenylbutene by PhSiH3 in the presence of samarium complexes R- or S-137a.
Molecules 30 02511 sch081
Scheme 82. Hydrogenation of 2-phenylbut-1-ene and styrene deuteration catalyzed with Zr, Ti, or Sm complexes.
Scheme 82. Hydrogenation of 2-phenylbut-1-ene and styrene deuteration catalyzed with Zr, Ti, or Sm complexes.
Molecules 30 02511 sch082
Scheme 83. Hydrogenation of trisubstituted olefins in the presence of Ti complex S,p-S,p-S-12.
Scheme 83. Hydrogenation of trisubstituted olefins in the presence of Ti complex S,p-S,p-S-12.
Molecules 30 02511 sch083
Scheme 84. Hydrogenation of tetrasubstituted olefins in the presence of p-R,p-R- or p-S,p-S-17 and [PhMe2NH]+[B(C6F5)4].
Scheme 84. Hydrogenation of tetrasubstituted olefins in the presence of p-R,p-R- or p-S,p-S-17 and [PhMe2NH]+[B(C6F5)4].
Molecules 30 02511 sch084
Scheme 85. Synthesis of enantiomerically enriched N-heterocycles.
Scheme 85. Synthesis of enantiomerically enriched N-heterocycles.
Molecules 30 02511 sch085
Scheme 86. Asymmetric hydroamination/cyclization of aminoalkenes catalyzed with lanthanide η5-complexes.
Scheme 86. Asymmetric hydroamination/cyclization of aminoalkenes catalyzed with lanthanide η5-complexes.
Molecules 30 02511 sch086
Scheme 87. Asymmetric reduction/cyclization of amino- and phosphinoalkenes in the presence of complexes with octahydrofluorenyl ligand S-141ac.
Scheme 87. Asymmetric reduction/cyclization of amino- and phosphinoalkenes in the presence of complexes with octahydrofluorenyl ligand S-141ac.
Molecules 30 02511 sch087
Scheme 88. Cyclization of aminodienes and aminoalkenes in the presence of complexes S-141a, S-133b or S-134b.
Scheme 88. Cyclization of aminodienes and aminoalkenes in the presence of complexes S-141a, S-133b or S-134b.
Molecules 30 02511 sch088
Scheme 89. Asymmetric catalytic epoxidation of alkyl- and aryl-substituted alkenes.
Scheme 89. Asymmetric catalytic epoxidation of alkyl- and aryl-substituted alkenes.
Molecules 30 02511 sch089
Scheme 90. Ring-opening of epoxides in the presence of pure ansa-titanocene p-S,p-S-11.
Scheme 90. Ring-opening of epoxides in the presence of pure ansa-titanocene p-S,p-S-11.
Molecules 30 02511 sch090
Scheme 91. Homocoupling of benzaldehyde catalyzed with titanocene complexes 52 and 53.
Scheme 91. Homocoupling of benzaldehyde catalyzed with titanocene complexes 52 and 53.
Molecules 30 02511 sch091
Scheme 92. Ti-catalyzed synthesis of γ-lactol (+)-285.
Scheme 92. Ti-catalyzed synthesis of γ-lactol (+)-285.
Molecules 30 02511 sch092
Scheme 93. Enantioselective isomerization of trans- or cis-4-tert-butyl-1-vinylcyclohexanes in the presence of chiral titanium ansa-complexes 52 or 73.
Scheme 93. Enantioselective isomerization of trans- or cis-4-tert-butyl-1-vinylcyclohexanes in the presence of chiral titanium ansa-complexes 52 or 73.
Molecules 30 02511 sch093
Table 1. The synthesis of C1-symmetric enantiomerically pure ansa-complexes of lanthanides [84].
Table 1. The synthesis of C1-symmetric enantiomerically pure ansa-complexes of lanthanides [84].
EntryComplexR*Solvent for Synthesis (Crystallization)Yield, %Ratio of Epimers R:S (Solvent)
1127a (La)(+)-neomenthylTHF (Et2O)4775:25 (THF)
2128a (Nd)(+)-neomenthylTHF (Et2O)6574:26 (THF)
3129a (Sm)(+)-neomenthylTHF (Et2O)4774:26 (THF)
4129a (Sm)(+)-neomenthylTHF (Et2O)5834:66 (Et2O)
5130a (Y)(+)-neomenthylTHF (Et2O)6374:26 (THF)
46:54 (Et2O)
6131a (Lu)(+)-neomenthylTHF (Et2O)5774:26 (THF)
51:49 (Et2O)
7129b (Sm)(−)-menthylTHF (Et2O)60>95:5 (Et2O)
20:80 (THF)
9130b (Y)(−)-menthylTHF (Et2O)6420:80 (THF)
90:10 (Et2O)
10131b (Lu)(−)-menthylTHF (Et2O)7020:80 (THF)
85:15 (Et2O)
11130c (Y)(−)-phenylmenthylTHF (Et2O)486:94 (THF)
83:17 (Et2O)
12129b, S (Sm)(−)-menthylTHF, DME (Et2O)88
13131b, S (Lu)(−)-menthylTHF, DME (Et2O)89
14129a, R (Sm)(+)-neomenthylTHF, DME (Et2O)54
Table 2. Carbo-(182) and cycloalumination (184) of alkenes, catalyzed with enantiomerically pure metallocenes.
Table 2. Carbo-(182) and cycloalumination (184) of alkenes, catalyzed with enantiomerically pure metallocenes.
CatalystAlR3ProductR′ee%Ref.
p-S,p-S-187AlMe3182n-C6H13, i-Bu, Cy, Bn, Ph, C4H8OH, C3H6NEt265–85, R[130]
182n-C4H9, n-C6H1348–73, R[134]
AlEt3182R′ = n-C6H1392, R[135]
182n-C4H9, n-C5H11, n-C6H13, n-C7H15,
n-C8H17, i-Bu, Cy, Cy8, Ph, Bn
33–68, S[136]
184n-C6H1333[135]
184n-C4H9, Bn5–11, S[136]
Cy, Cy845–57, S
Ph31–40, R
p-R,p-R-13AlMe3182n-C8H176, R[130]
R,p-R,p-R-16AlMe3182n-C6H138, R[130]
p-S,p-S-13AlEt3182n-C6H1351, R[51]
184n-C6H1311, S
S,p-S,p-S-16AlEt3182n-C6H1315, R[51]
184n-C6H1326, S
p-S,p-S-18AlMe3182n-C6H1358, R[51]
AlEt3182n-C6H1350, R
184n-C6H1312, S
S,p-S,p-S-19AlEt3182n-C6H1315, R[51]
184n-C6H1320, S
p-S,p-S-149AlMe3182Bn29[132]
p-S,p-S-188aAlMe3182Bn25[132]
p-S,p-S-188bAlMe3182Bn29[132]
p-S,p-S-188cAlMe3182Bn33[132]
p-S,p-S-189AlMe3182Ph80[132]
p-R,S-190AlMe3182Ph,19[133]
Bn28
Table 3. The polymerization of racemic α-olefins with bulky substituents 199af, catalyzed with enantiomerically pure S-116ad in the presence of MAO.
Table 3. The polymerization of racemic α-olefins with bulky substituents 199af, catalyzed with enantiomerically pure S-116ad in the presence of MAO.
EntryAlkene-1ComplexTime, hTOF, h−1Conversion, %s%ee
1199aS-116a1872242.813.3
2S-116a474738.32.420.3
3S-116a 60 2.6
4S-116b 266 2.3
5S-116c 374 2.5
6S-116d 280 3.2
7199bS-116a13.5551751.840
8S-116b 299 1.8
9S-116c 706 1.9
10199cS-116a40.545.632.517.640.6
11S-116a693442.415.958.6
12S-116a 40 16.8
13S-116b 109 12.5
14S-116c 450 15
15199eS-116a16.577.664.41.17.6
16S-116a16.573.758.71.14.6
17S-116a 75 1.1
18S-116b 111 1.2
19S-116c 633 1.2
20199fS-116a 18 7.6
21S-116d 16 3.2
22199dS-116a22.755.837.82.016.2
23S-116a433756.12.130.3
24S-116a 37 2.1
25S-116b 49 6.4
26S-116c 111 1.6
27S-116d 988 8.5
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

Kovyazin, P.V.; Khalilov, L.M.; Parfenova, L.V. Enantiomerically Pure ansa-η5-Complexes of Transition Metals as an Effective Tool for Chirality Transfer. Molecules 2025, 30, 2511. https://doi.org/10.3390/molecules30122511

AMA Style

Kovyazin PV, Khalilov LM, Parfenova LV. Enantiomerically Pure ansa-η5-Complexes of Transition Metals as an Effective Tool for Chirality Transfer. Molecules. 2025; 30(12):2511. https://doi.org/10.3390/molecules30122511

Chicago/Turabian Style

Kovyazin, Pavel V., Leonard M. Khalilov, and Lyudmila V. Parfenova. 2025. "Enantiomerically Pure ansa-η5-Complexes of Transition Metals as an Effective Tool for Chirality Transfer" Molecules 30, no. 12: 2511. https://doi.org/10.3390/molecules30122511

APA Style

Kovyazin, P. V., Khalilov, L. M., & Parfenova, L. V. (2025). Enantiomerically Pure ansa-η5-Complexes of Transition Metals as an Effective Tool for Chirality Transfer. Molecules, 30(12), 2511. https://doi.org/10.3390/molecules30122511

Article Metrics

Back to TopTop