Next Article in Journal
Pyrolysis Kinetics of Lignin-Based Flame Retardants Containing MOFs Structure for Epoxy Resins
Previous Article in Journal
Research Progress on Chemical Constituents and Pharmacological Activities of Menispermi Rhizoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Jumping in the Chiral Pool: Asymmetric Hydroaminations with Early Metals

TU Chemnitz, Research Center for Materials, Architectures and Integration of Nanomembranes (MAIN), Research Group Organometallic Chemistry, Rosenbergstraße 6, D-09126 Chemnitz, Germany
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2702; https://doi.org/10.3390/molecules28062702
Submission received: 8 February 2023 / Revised: 3 March 2023 / Accepted: 6 March 2023 / Published: 16 March 2023
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
The application of early-metal-based catalysts featuring natural chiral pool motifs, such as amino acids, terpenes and alkaloids, in hydroamination reactions is discussed and compared to those beyond the chiral pool. In particular, alkaline (Li), alkaline earth (Mg, Ca), rare earth (Y, La, Nd, Sm, Lu), group IV (Ti, Zr, Hf) metal-, and tantalum-based catalytic systems are described, which in recent years improved considerably and have become more practical in their usability. Additional emphasis is directed towards their catalytic performance including yields and regio- as well as stereoselectivity in comparison with the group IV and V transition metals and more widely used rare earth metal-based catalysts.

1. Introduction

The coupling of carbon and nitrogen bonds is of great importance to organic chemistry [1,2]. The thusly formed nitrogen-containing compounds including N-heterocycles offer diverse applications not only in material sciences, but also in natural product synthesis and pharmaceutical chemistry. One synthetic concept in the mostly applied preparation of such molecules is the hydroamination reaction [3,4,5,6,7,8,9,10,11].
Hydroamination is the addition of an N–H bond of a primary or secondary amine across a carbon–carbon double or triple bond of, for example, alkenes, alkynes, dienes or allenes, resulting in an optimal atomic economy of 100% [5,6,12,13]. However, asymmetric C,N coupling processes including the Aza–Wacker [14], Buchwald–Hartwig [14], aminoacetoxylation [14] and photoredox (aminium radicals) [15] reactions are less atomically efficient than hydroaminations. Depending on the substrates used, hydroamination reactions occur either intermolecularly, in which the relevant functional groups are part of the separated starting materials, or intramolecularly, wherein the substrates combine both the amine and unsaturated C=C or C≡C building blocks in a single molecule [5].
Commonly, the intermolecular hydroamination of alkenes and alkynes results in the formation of Markovnikov and/or anti-Markovnikov regioisomers [5]. In the case of allenes and alkynes, E/Z isomers are produced [6,7,16]. Intramolecular hydroamination favors the Markovnikov product giving α-alkyl N-heterocycles for alkene substrates [5]. In addition, substituents in the β-position to the amino unit of the nitrogen-bonded unsaturated organic carbon–hydrogen substrate affect the reaction rate, which is known as Thorpe–Ingold effect [17].
Hydroamination reactions are thermodynamically neutral [5,6,18,19]. Due to the electrostatic repulsion between the nitrogen lone pair, the C,C π-system and the orbital symmetry-forbidden [2 + 2] cycloaddition, hydroamination reactions possess a high reaction barrier despite being kinetically favored as caused by the increase in the total bonds. Therefore it is necessary to catalyze or run the respective reactions at a high temperature [5,6].
To the best of our knowledge, the first hydroamination in solution, the C,N coupling reaction of p-toluidine with cyclohexene, was reported by Hickinbottom in 1932 [20]. Shortly after this, Kozlov et al., published the catalytically controlled hydroamination of an amine with an alkyne in the presence of mercury (II), copper (II) or silver (I) halides as catalysts [21,22]. In 1971, Coulson described the reaction of amines with alkenes by using catalytic active Rh and Ir species [23]. Since then, the field of C,N coupling via hydroamination has been expanding [6,12,24,25,26,27,28,29,30]. Early transition metals of group IV and V from the periodic table of elements were introduced by Bergman and Livinghouse [31,32]. Rare earth metal-based catalysts were launched by Marks dating back to 1989 [33], and early main-group elements as catalytic systems were established at the start of the new millennium [34].
Main-group or lanthanide-element-based catalysts are generally less tolerant towards amines and alkenes featuring polar functional groups, e.g., esters, ketones and alcohols and are more sensitive towards air and moisture as compared to late transition metal complexes. However, they exhibit an overall higher reactivity, a better regioselectivity and are more ecologically friendly compared with late transition metal hydroamination catalysts. Hence, early-metal-based catalysts are the preferred catalysts over expensive and often toxic late transition metal ones, especially in intramolecular hydroamination reactions [5,6,13].
Since intermolecular alkene hydroaminations and intramolecular cyclization reactions of aminoalkenes may form stereocenters, chiral catalysts are required to obtain enantiopure isomers. For this to occur, ligands such as 1,1′-bi-2-naphthol(=BINOL), 2,2′-diamino-1,1′-binaphthaline(=DABN) or 2,2′-bis(diphenylphosphino)-1,1′-binaphthyl(=BINAP) derivatives are best suited, due to their bulk and (excellent) enantioselectivity [35,36]. In addition to these synthetic, accessible but hard to purify and difficult to up-scale biaryls, a series of enantiopure building blocks provided by nature are also of great benefit. The so called chiral pool-based ligands are readily available, ecologically friendly and hence “green” [37,38,39,40]. Due to their low cost, high abundance and general sustainability, the chiral pool has been extensively utilized by synthetic chemists in the preparation of ligand systems in enantioselective catalysis of natural products as well as pharmaceutical agents, with an extensive literature available on these topics [39,41,42,43,44,45,46]. The most relevant chiral pool motifs in hydroamination reactions are amino acids, both proteinogenic and non-proteinogenic, alkaloids and terpenes (Figure 1).
While originally only naturally occurring, enantiopure compounds were considered to be part of the chiral pool, modern definitions tend to include on a significant scale industrially produced, enantiomerically pure compounds, which can be obtained either by racemate cleavage, enantioselective synthesis or the derivatization of enantiomerically pure natural products [47].
Herein, we focus on asymmetric homogeneous metal-catalyzed hydroamination coupling reactions using early group IV and V transition, rare earth and main-group metals as catalysts featuring ligands originating from the chiral pool. The regio- and stereoselectivities, activities, conversions and yields towards the formation of the corresponding hydroamination products will be discussed in dependence of the metals, chiral pool motifs and the appropriate catalysis conditions.

2. Chiral Pool-Based Catalysts for Asymmetric Hydroamination Reactions

In the first catalytic hydroamination reaction dating back to the early nineteen thirties, group XI and XII metal halides were applied as catalysts [21,22]. Since then, a multitude of metal compounds have been researched for their suitability as catalytic active systems in inter- and intramolecular hydroamination reactions [3,5,48]. The catalysts can be differentiated into late and early transition metals, rare earth metals and early main-group elements. In the following, the application of hydroamination catalysts especially featuring ligands originating from the natural chiral pool will be discussed in detail and compared to non-chiral pool ligands.

2.1. Late Transition Metals

Late transition metal catalysts containing, for example, neutral chiral (di)phosphine-, bipyridine- or bisoxazoline-based ligands to induce regioselectivity and chirality have recently been used [14,49,50,51]. In addition, chiral pool relevant motifs such as α-hydroxy acids and sugar acids including tartaric acid were introduced as chiral centers in phosphine ligands. Examples include 2,3-O-isopropyliden2,3dihydroxy-1,4-bis(diphenyl-phosphino)butane(=DIOP) or (2S,3S)-(−)-bis(diphenylphosphino)butane(=CHIRAPHOS). However, these systems induce generally lower ee values in comparison to non-chiral pool-derived phosphines, e.g., BINAP derivatives. Detailed discussions on this topic can be found elsewhere [3,7,14,16,49,50,51,52]. During the last three decades, focus has also been directed to hydroaminations applying early metal catalysts, including those featuring chiral-pool-derived ligand peripheries, especially for their application in intramolecular hydroaminations [4,5,6].

2.2. Early Transition Metals

Early transition metals and rare earth metals have been extensively studied in intramolecular hydroamination catalysis [24]. The HSAB principle states that “hard metals” bind to “hard ligands”. Therefore, early metals have been combined with ligands, such as amines, alcohols and ethers. For transition metals of group IV of the periodic table of elements and rare earth metals, cyclopentadienyls have also been proven to be excellent ligands for the catalytic active center [33,53,54,55,56,57,58].

2.2.1. Rare Earth Metals

The intramolecular hydroamination of non-activated olefins using rare-earth metal complexes was pioneered by the group of Marks in the 1990s [33,54]. The majority of the catalysts featuring cyclopentadienyl entities allows the efficient generation of racemic or enantio-enriched N-heterocycles [33,54,59,60,61,62,63]. Using these systems, mechanistic studies were undertaken and two mechanisms were proposed [64,65,66,67]. The σ-insertion mechanism suggested by Marks et al., (Figure 2a) postulates a rapid, reversible migratory olefin insertion of the metal amide followed by a slower, irreversible rate-determining metal alkyl bond protonolysis by a further substrate molecule [29,53,54,55,61,68,69,70]. The turnover-limiting M–C σ-bond aminolysis occurs by a substrate molecule which is followed by a kinetically favored displacement of the N-heterocycle, as confirmed by deuterium-labelling experiments [29,54,61,68,69]. Exemplary NH/ND kinetic isotope effect (=KIE) and isotopic perturbation studies on Cp*2LnR (Ln = La, Nd, Sm, Y, Lu; R = H, CH(TMS)2, ƞ3-C3H5, N(TMS)2; TMS = SiMe3) complexes were carried out to define the stereochemistry of the corresponding heterocycles [71]. These studies found that the NH/ND KIE cannot be derived from protonolysis of a previously formed Ln–C bond. In order to explain this finding a non-insertive catalytic cycle was proposed (Figure 2b), involving a second coordinated amine substrate, partially transferring one of its two NH protons to the terminal alkene carbon atom to form the pyrrolidine product by insertion (Figure 2b) [29,68,71]. Finally, the coordinated pyrrolidine is released by a new substrate molecule [29,68,71]. The two discussed mechanisms for intramolecular hydroamination reactions using rare earth (or main-group) metal-based catalysts compete with each other [71].
One example is the enantioselective and regioselective hydroamination of aminoalkenes 1a,b and 2a,b to give chiral pyrrolidines 3a,b or piperidines 4a,b with C1-symmetric lanthanide ansa complexes: (S)-[Me2Si(η5-C5Me4)(η5-C5H3R*)]Ln-E(TMS)2 (5an, Ln = Y, La, Sm, Nd, Lu; E = N, CH; R* = (−)-menthyl, (+)-neomenthyl; TMS = SiMe3) [54,73] and (S)-[Me2Si(OHF)(η5-C5H3R*)]LnN(TMS)2 (6ac, Ln = Y, Sm, Lu; OHF = η5-octahydrofluorenyl) [60] serving as catalysts (Scheme 1, Table 1) [54,60,73].
From Table 1 it can be seen that catalysts 5an and 6ac achieve moderate to high ee values despite facile epimerization under the catalytic reaction conditions, due to reversible protolytic cleavage of the metal cyclopentadienyl bond [54,60,61,73,74]. A further characteristic is that the (+)-neomenthyl-containing catalysts 5cg,j,k (Table 1, entries 3–12, 21–27) form the corresponding (R)-(−) enantiomers, while the (−)-menthyl comprising derivatives 5a,b,h,i (Table 1, entries 1, 2, 13–19) and 6ac (Table entries 32–46) give the respective (S)-(+)-configured N-heterocycles 3a,b and 4a,b, with the exception of entries 18 and 44 in Table 1. The chirality of the catalysts has no effect on the optical rotation of the product. However, when lutetium complexes 5ln (Table 1, entries 28–31) are used as catalysts, then aminopentenes 1a,b are cyclized to give enantiomers of 3a,b in the exact opposite enantioselectivity to 5ak (M = Y, La, Nd, Sm) [54,73]. The best overall enantioselectivities with up to 74% ee were obtained for 1b using the (−)-menthyl-substituted samarium complexes (S)-5h,i at −30 °C (Table 1, entry 17) [54,60,73]. Generally, both (R)- (Table 1, entries 19 and 21) and (S)-enantiomers (Table 1, entries 13 and 20) of 5h,j show comparable ee values (5h: 60 vs. 62%; 5j: 52 vs. 55%) and the same optical rotation for pyrrolidine 3a at 25 °C [54,60,61,73]. For catalysts 6ac using 1a,b and 2a,b as substrates, an ee value as high as 67% ((S)-6a) was obtained (Table 1, entry 36). The activities of 6ac are in general lower than those of 5an, hence, the use of catalysts 6ac requires higher temperatures [60].
Additionally, catalysts (S)-5h and (S)-6a,b were studied in the hydroamination/cyclization of sterically hindered aminoalkenes E-7a,b and Z-8, producing pyrrolidines 9a,b and piperidine 10 with good to excellent yields and ee values as high as 68% (with (S)-6a as a catalyst (Table 2, entries 6 and 7), albeit at much harsher conditions (Table 2) [70]. Generally, the (+)-enantiomers 9a,b and 10 are formed. However, using (S)-6a as pre-catalyst for the cyclization of aminohexene Z-8 to piperidine 10, the corresponding (-)-enantiomer was obtained.
Substrate screening was later extended by the group of Marks et al., towards the conjugated 1,3-aminodienes 11 and 12a,b using (S)-5a,b, (S)-5h and (S)-6b as organolanthanide catalysts (Table 3) [61]. The reaction rate is higher for the aminodienes 11 and 12a,b than for the corresponding aminoalkenes 1a and 2a,b, despite increased steric hindrance of the cyclization transition state [25,61]. However, the enantioselectivity is generally lower, with the exception of the formation of N-heterocycle 14a with (S)-6b as a catalyst showing up to 71% ee (Table 3, entry 10) [25,61]. The authors also show the high stereoselectivity of the intramolecularly proceeding aminodiene hydroamination by concise synthesis of naturally occurring alkaloids (±)-pinidine and (+)-coniine from easily accessible diene substrates [25,61].
In 2003, Marks et al., published a series of C2-symmetric bis(oxazolinato)lanthanum complexes and discussed their use as efficient catalysts for the intramolecular hydroamination of aminoalkenes and aminodienes [75]. Two complexes out of the reported series possess L-valinol- (15a) (Table 4, entry 1) and L-tert-leucinol-derived (15b) (Table 4, entry 2) chiral pool ligands for the cyclization of 1b (Scheme 2) [75]. However, the observed enantioselectivities were with 6% (15a) and 39% (15b) at 25 °C lower than those for the non-chiral-pool-based systems with aryl functionalities in the α-position to the nitrogen atom, which result in up to 67% ee for substrate 1b. Generally, it can be stated that lanthanides possessing the largest ionic radii display the highest turnover frequencies and enantioselectivities in the hydroamination for these systems [75].
Table 4. Catalytic asymmetric hydroamination reactions of 1ae and 2c,d using chiral rare earth metal complexes 15an, 16ad, 17a, 18 and 19a,b.
Table 4. Catalytic asymmetric hydroamination reactions of 1ae and 2c,d using chiral rare earth metal complexes 15an, 16ad, 17a, 18 and 19a,b.
EntryCat.R*[cat]
[mol-%]
Substr.Prod.T
[°C]
t
[h]
Conv.
[%]
ee
[%] a
Ref.
115aiPr5 b1b3b23 bn.a.>986 (R)[75]
215btBu5 b1b3b23 bn.a.>9839 (R)[75]
315ciPr101d3d220.25>9943 c[76]
415dBn101d3d220.25>9930 c[76]
515eiPr101b3b3072>997 c[76]
6 101d3d220.25>995 c[76]
715fBn101b3b3072>996 c[76]
8 101d3d220.25>996 c[76]
915giPr101b3b2212>9914 c[76]
10 101d3d221>9916 c[76]
1115hBn101b3b2212>9910 c[76]
12 101d3d221>9942 c[76]
1315iiPr101b3b22168--[76]
14 101d3d22 d12>9936 c[76]
1515jBn101b3b22168--[76]
16 101d3d22 d12>9946 c[76]
1715kiPr101b3b2212>9914 c[76]
18 101d3d22 d12>9930 c[76]
1915lBn101b3b2212>9912 c[76]
20 101d3d22 d12>9930 c[76]
2115miPr101d3d22 e0.25>9938 c[76]
2215nBn101d3d22 e0.25>9932 c[76]
2316a-51b3b605.5952[77]
2416b-51b3b101689566[77]
2516c-51b3b25288955[77]
2616d-71b3b2589511 c[78]
27 101c3c250.510011 c[78]
28 82c4c25641005 c[78]
2917a(+)-neomenthyl41a3a65659622 (R)[79]
30 31b3b2512.79621 (R)[79]
3118(−)-menthyl31b3b256.258011 (S)[79]
3219aiPr51c3cr.t.0.1710034 (S)[56]
33 52d4dr.t.0.8310022 (S)[56]
3419btBu51c3cr.t.0.1710093 (S)[56]
35 51d3dr.t.0.1710094 (S)[56]
36 51e3er.t.39589 (S)[56]
a Enantiomeric excesses (=ee) determined either by 1H or 19F NMR spectroscopy after reaction of Mosher’s acid chloride or by HPLC after naphthoylation, or determined by chiral shift 1H NMR spectroscopy using (R)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers. b 6 mol-% ligand. c Absolute configuration not determined. d In toluene-d8. e In bromobenzene-d5. n.a.—not applicable.
Scheme 2. Catalytic asymmetric hydroamination of aminoalkenes 1ae and 2c,d using chiral rare earth complexes 15an, 16ad, 17a, 18 and 19a,b [56,75,78,79]. (For more details concerning catalysis data see Table 4).
Scheme 2. Catalytic asymmetric hydroamination of aminoalkenes 1ae and 2c,d using chiral rare earth complexes 15an, 16ad, 17a, 18 and 19a,b [56,75,78,79]. (For more details concerning catalysis data see Table 4).
Molecules 28 02702 sch002
Ward et al., discussed the application of bis(oxazolinylphenyl)amide(=BOPA) rare earth metal complexes 15cl (M = Y, La, Pr, Nd, Sm) in the hydroamination/cyclization of 1b,d (Table 4, entries 3–20) [76]. Enantioselectivities of a maximum of 46% for catalyst 15j (Table 4, entry 16) could be reached. Additionally, anionic yttrium catalysts 15m,n were studied, showing lower enantiomeric excesses for 3d than their respective neutrally charged counterparts 15c,d.
Kim et al., reported three chiral-pool-based yttrium catalysts (16ac) in which two alkylated (16a,b, R = Me; 16c, R = CH2-tert-Bu) L-proline-derived moieties are attached to a 2,2′-diaminobinaphthyl (16a) or 1,2-diaminobenzene (16a,b) backbone for the intramolecular hydroamination of 1b (Scheme 2, Table 4, entries 23–25) [77]. Complexes 16ac displayed excellent activities with conversions of 95%; only 16b showed good selectivity (66% ee) (Table 4, entry 4), while both 16a and 16c displayed only very low ee values [77].
In 2007, Carpentier et al., reported on the successful application of the yttrium catalyst 16d (Scheme 2; Table 4, entries 26–28), comprising a C2-symmetric chiral tetradentate diamine–diamide ligand with two L-proline-derived building blocks attached to the N,N’-dimethylethylenediamine backbone, in the intramolecular hydroamination of aminoalkenes 1b,c and 2c. Despite the high activities, only ee values as high as 11% could be reached for 3b,c [78]. In addition, 16d is suited for the rac-lactide ring-opening polymerization at ambient temperatures, whereby isotactic-enriched polylactides were formed [78].
The Hultzsch group published the synthesis, chemical and physical properties of (+)-neomenthyl-functionalized cyclopentadienyl and indenyl yttrocene complexes 17a and 18 (Scheme 2, Table 4 and Table 11) [79]. The synthetic methodology to prepare 17a includes a facile arene elimination starting from [Y(o-C6H4CH2NMe2)3], while 18 was accessible by salt metathesis from the lithium species and YCl3. for comparison, the (−)-phenylmenthyl derivative 17b was also prepared. Complexes 17a and 18 displayed moderate to good catalytic activity in the tested asymmetric hydroamination reactions (Table 4, entries 9–1, and Table 11, entries 1 and 2), but only low to moderate enantioselectivities of up to 22% (Table 4, entry 29) for 17a and 11% ee (Table 4, entry 31) for the sterically more hindered catalyst 18 were observed in the cyclization of 1a,b [79]. The catalytic activity and enantioselectivity of non-chiral-pool-derived 17b was comparable to 17a. Furthermore, the authors indicated that the protolytic loss of an indenyl ligand in 18 occurs at low catalyst loading (⩽0.5 mol-%), when applying the sterically undemanding substrate 1a [79].
In 2011, Manna et al., introduced a highly enantioselective bis(amido)yttrium complex based on chiral cyclopentadienylbis(oxazolinyl)borates(19a,b), in which the chirality is induced by L-valinol-(19a) and L-tert-leucinol-derived(19b) moieties (Scheme 2) [56]. The catalyst 19b in the intramolecular hydroaminations of primary aminoalkenes 1ce (Table 4, entries 12–14) and aminodialkenes 20ad (Table 11, entries 3–6) showed excellent activities and yielded the corresponding pyrrolidines with high optical purities ranging from 89% to 94% ee (Table 4, entries 34–36) in the synthesis of 3ce or from 92% to 96% (Table 11, entries 3–6) for the transformations of 20ad. The achieved values for the enantiomeric excess are comparable to those obtained for the isostructural zirconium complex (Tables 10 and 11) [56]. However, the (R)-configuration of the generated stereocenter is opposite to the pyrrolidines 3ce formed with the yttrium analog 19b. Furthermore, the authors report on mechanistic studies, indicating that 19b reacts by concerted C–N and C–H bond formations, which is maintained by the kinetic rate law for conversion, saturation of the respective substrate under initial rate conditions, isotopic enantioselectivity disruption and kinetic isotope effects [56]. By carrying out N–H/N–D kinetic studies, Manna et al., were able to show that the stereochemistry determining step for both Y and Zr catalysts involves an N–H (or N–D) bond. They demonstrated that the (S)-diastereomeric pathway is slowed down to greater extent than the (R)-pathway for both metal centers. Based on these results, they conclude that the catalysts have similar transition states but are of opposite energetic favorability, resulting in the observed difference in stereoselectivity [56].
Rare earth metal catalysts in which chiral-pool-modified ligands are present impose enantioselectivity, showing moderate to excellent activities in the intramolecular hydroamination for a variety of substrates. BOX-based yttrium complex 19b displays a 96% ee, and shows similar activities and ee values in comparison to non-chiral pool catalysts of which biaryls such as BINOL or 2,2′-bis-(diphenylphosphinoamino)-1,1′-binaphthyl(=BINAM) derivatives are the best studied examples [4,5,30,80,81,82,83,84,85,86,87]. Hultzsch et al., for example, described 3,3′-bis(trisarylsilyl)- and 3,3′-bis(arylalkylsilyl)-substituted binaphtholate rare earth metal complexes (M = Y, Lu) for the hydroamination/cyclization of 1ad and 2d with enantioselectivities of up to 95% (M = Lu) and 90% (M = Y) [88]. Overall, for the conversions, no difference is observable (>95%). On the other hand, enantioselectivities often vary significantly. For 1a,b, the difference in ee for the yttrium catalysts is with 14% (1a, (R)-5a,b) and 6% (1b, 16b) rather moderate. For 1c,d the chiral-pool-based system 19b performs with a difference in ee of 4% and 9%, which is better than the non-chiral-pool-derived ones. For 2d, the respective difference in enantioselectivity is 29% (19a), higher in favor of the non-chiral-pool-based catalysts. For luthetium, the difference between chiral pool and non-chiral pool catalysts grows even larger: 59% for substrate 1a ((R/S)-5n) and 49% for 1b ((R)-5m) [88].
Chai et al., reported on a tridentate-linked amido–indenyl yttrium complex on the basis of 1,2-diaminocyclohexane, which transforms amino-olefins 1bd and 2bd into the corresponding N-heterocycles with ee values of up to 97% [89]. Those systems show a similar enantioselectivity as the 3,3′-bis(arylalkylsilyl)-substituted binaphtholate complexes towards aminoalkenes 1bd and a higher enantioselectivity towards 2d, which increases the difference to the chiral-pool-derived catalyst even further.

2.2.2. Group IV and V Metals

Aminoalcoholates of titanium (21ac, 22ac, 23ai, 25ai, 27aj and 29ah), zirconium (32ag, 33ae) and tantalum (24al, 26al and 28aj) (Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12), as well as the cyclopentadienylbis(oxazolinyl)borate group IV metal complexes 30ac and 31 are admirable enantioselective hydroamination/cyclization catalysts for a variety of different aminoalkenes, aminodialkenes and aminoallenes, as shown by Johnson [90,91,92,93,94,95] and Sadow [56,58,96]. Complexes 2133 commonly feature natural chiral-pool-derived ligands based on either L-valine (21ac, 23ai, 24al, 29e, 30ac, 32ag), L-phenylalanine (22ac, 25ai, 26al, 27aj 28aj, 29ad), L-tert-leucine (31), L-proline (33ad) and L-pipecolic acid (33e) (Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12).
Chiral titanium aminoalcohol catalysts 21ac and 22ac with N-alkyl substituents R = 2-Ad (= 2-adamantyl), cC6H11 or iPr allowed the effective ring-closing hydroamination of substituted aminoallenes 34a,b (34a, Table 5; 34b, Table 6) [90]. The cyclization of hepta-4,5-dienylamine 34a resulted in the formation of a mixture of the six-membered 6-ethyl-2,3,4,5-tetrahydropyridine 35 (19–33% yield, Table 5) and the five-membered Z-(Z-36) as well as E-pyrrolidines (E-36) (67–86% combined yield) with ee values of up to 8% (Z-36) and 16% (E-36) at 110 °C (Table 5).
In contrast, the cyclization of the more sterically hindered 6-methylhepta-4,5-dienylamine 34b afforded exclusively five-membered 2-(2-methylpropenyl)pyrrolidine 37 with high conversions (Table 6). Nevertheless, the enantiomeric excesses of 37 are with a maximum of 15% ee (Table 6, entry 6) [90]. A significantly higher rate acceleration when using 21ac and 22ac as catalysts in comparison to the titanium complex Ti(NMe2)4 was observed. It is still an open question if either isolated or in situ-generated metal imidos, which are common for group IV catalysts, are the catalytic active species [90]. Comparative experiments with phenylglycine-derived ligands (= Phg) were carried out showing similar activities towards 34a,b as for 22ac [90].
In 2009, Johnson et al., extended the series of aminoalcohol-based titanium catalysts 21ac and 22ac towards the more bulky chiral compounds 23ai and 25ai by replacing R = H by R = methyl, nbutyl or phenyl groups [91]. The corresponding ligands were prepared by a consecutive two-step synthetic procedure, whereas catalysts 23ai and 25ai were generated in situ. Intramolecular hydroamination of aminoallene 34b exclusively results in pyrrolidine 37 with enantiomeric excesses of 16% (max.) at 135 °C (Table 6, entries 30 and 33) with quantitative conversions. No correlation between the steric bulk of the ligands and the ee values could be identified [91].
The Johnson group later used the previously discussed aminoalcohols (vide supra) for the preparation of the respective tantalum complexes (catalysts 24al and 26al) [92]. In comparison with titanium complexes 21ac, 22ac, 23ai and 25ai, which are dimeric in the solid state, tantalum compounds 24al and 26al are monomeric possessing a somewhat distorted trigonal-bipyramidal structure as confirmed by single crystal X-ray structure analysis. Next to the chiral pool motifs derived from L-valine and L-phenylalanine, non-natural D-valine and D-phenylalanine were also studied. The best results in the cyclization of aminoallene 34b to pyrrolidine 37 were obtained by catalysts containing R’ = Ph as substituents (24d,h,l and 26d,h,l). Enantioselectivities ≤ 80% ee were obtained with a 5 mol-% catalyst loading (Table 6, entries 19, 23, 27, 40, 44 and 48) [92]. Generally, the tantalum derivatives show better ee values than those of the respective titanium catalysts at the cost of higher reaction times and a greater variance in conversion rates.
Crowded sulfonamides featuring a benzyl group as a bulky chiral backbone (L-phenylalanine-derived) with different steric and electronic properties were successfully introduced as ligands for the in situ generation of titanium (27aj) and tantalum (28aj) catalysts [93,95]. The respective titanium catalysts convert 6-methylhepta-4,5-dienylamine 34b solely to 2-(2-methylpropenyl)pyrrolidine 37 with an enantiomeric excesses of max. 11% (27af) (Table 7, entries 1–12) [93] or 18–41% (27gj) (Table 7, entries 13–17) [95] with conversions of 18–100%. The corresponding tantalum catalysts generally showed an ee of 5–34% (28af) (Table 7, entries 18–29) [93] and 33–39% (28gj) (Table 7, entries 30–33) [95] more selective with generally higher conversions than 27gj.
In the hydroamination/cyclization of hepta-4,5-dienylamine 34a using 27gj and 28gj as catalysts, a mixture of tetrahydropyridine 35 (48–82% yield) and Z-36a and E-36b (15–50% combined yield) was obtained with ee values of up to 55% (E-36a) and 45% (Z-36b) (Table 8, entry 3) with tantalum showing an higher pyrrolidine yield and reduced enantioselectivities.
A further modification of the earlier discussed titanium catalysts 21ac and 22ac (Table 5 and Table 6, entries 1–6), which are suitable for aminoallene ring-closing reactions, was carried out by the introduction of chiral, tridentate, dianionic imine-diol ligands at the titanium metal center, resulting in the formation of 29ah (29ae, Table 8; 29ah, Table 9) [94,97]. Nevertheless, cyclization of hepta-4,5-dienylamine (34a) resulted in a mixture of tetrahydropyridine 35 (40–72% yield) and pyrrolidines Z-36 (8–17% yield) and E-36 (17–39% yield) (Table 8). Using 34b as a substrate, 37 was exclusively produced in the presence of 29ah (Table 9) as already described for the catalytic systems 21ac and 22ac (Table 6). The ee values of a maximum of 22% are comparable to the values observed for 21ac and 22ac with comparable conversions (Table 9, entry 10) [90,97].
For the intramolecular hydroamination of aminoalkanes using chiral-pool-derived catalysts, Sadow et al., published the highly enantioselective bis(amido)zirconium complex 30b possessing a chiral cyclopentadienylbis(oxazolinyl)borate in which chirality is induced by the incorporation of L-valinol into the ligand (Scheme 3) [56,57,96]. The addition of catalytic amounts of 30b to primary aminoalkenes 1af, 2c,d and 3840 yielded the corresponding N-heterocycles 3af, 4c,d and 4143 with enantiomeric excesses ranging from 31% for 2c (Table 10; entry 22) to 98% for 1d (Table 10; entries 14; 18). It was proposed that the observed reactivity and high enantioselectivity of 30b may relate to the ability of the relevant intermediate to stabilize the proposed six-center transition state [56,57]. Curiously, complex 30b and its yttrium derivative 18 (vide supra) gave pyrrolidines 3ce and 44d with an opposite absolute configuration, despite using the same ligand system as the (R)-derivative of 30b using D-valine as chiral building block. In addition, the L-tert-leucine derivative 31 was prepared in a multiple-step synthetic procedure [56]. The catalytic performance of 31 on the cyclization of aminoalkenes 1ce and 2c corresponds to L-valine-derived 30b, resulting in similar conversions with a maximum of 93% (Table 10; entry 36) and 29% ee (Table 10; entry 38) with generally high conversions. The existence of a kinetic rate dependence was further shown, evolving from a 1st order at a low substrate concentration to zero-order at a high concentration, which is representative of a reversible catalyst/substrate interaction preceding the N–H bond cleavage in the turnover-limiting and irreversible step of the catalytic cycle [56].
Exchanging zirconium in 30b by titanium (30a) or hafnium (30c), the latter two species catalyze the cyclization of amino-olefins 1bf, 2c,d and 3840 to result in the corresponding N-heterocycles 3bf, 4c,d and 4143 in enantiomeric excesses of 76–82% (30a) or 18–97% (30c) with moderate to high conversions (for more details see Table 10) [58]. This work was extended to aminodialkenes 20ah and aminodialkynes 45ac using 19b, 30b,c and 31 as catalysts as depicted in Scheme 4 [56,57,58,96]. Diastereomers 44ah (Scheme 4) of five- to seven-membered N-heterocycles were obtained when aminodialkenes 20ah were used as substrates, while in the case of aminodialkynes 45ac, the respective imines 46ac were produced in an enantioselective reaction in high to moderate yields. Depending on the cyclization conditions applied, diastereo- and enantioselectivities of max. 99% could be reached using zirconium catalyst 30b (Table 11, entries 13–15) [96]. In comparison, yttrium-based systems 17a and 18 reached lower ee values of up to 38% (Table 11, entries 1 and 2), while 19b showed similar enantioselectivities to 30b. It was found that catalytically generated stereocenters in cyclized 44ah can be independently controlled by the catalyst’s properties and reaction conditions (Table 11). At low concentrations Z-44b is favored, and at high concentrations combined with lower temperatures, E-44b (Table 11, entries 7–9) is favored. It could be further demonstrated that isotopic substitution of hydrogen by deuterium (H2NR/D2NR in 20b) significantly improved the diastereoselectivity from the ratio of 8:1 to a maximum of 43:1 and increased the optical purity to 99% ee [96]. As demonstrated for 30b, experimental studies on aminodialkene ring-closing reactions to examine the effects of the catalyst-to-substrate ratio, the absolute catalyst concentration and the absolute original substrate concentration show that the latter parameter greatly influences the stereoselectivity, whereas the absolute configuration of the α-amino stereocenter created by the C–N bond generation is not influenced by any parameters of the concerted proton-triggered cyclization mechanism (Figure 3) [96]. Coordination of a primary amine changes the ring conformation in the transition state to place the cis group axial to avoid unfavorable interactions between the bulkier substituent and the cyclizing substrate resulting in the formation of the trans diastereomer. With decreasing concentrations of the primary amine, pathway B becomes more unlikely, while cycle A is more favored, resulting in the increased formation of the cis diastereomer [96]. With amine deuteration, the coordination becomes more hindered, resulting in the observed increase in the respective cis product.
Furthermore, dibenzyl zirconium complexes 32ag and 33ae (Scheme 5, Table 12) have been applied in the cyclization of primary aminoalkenes 1be, 2d and 20a,b [98]. The chirality of the appropriate catalyst was introduced by the L-valine- (32ag), L-proline- (33ad) or L-pipecolic acid-derived (33e) backbone of the tridentate dianionic amino–diol ligand with variation possibilities being at the ether functionality (32ag) or the substituents of the α-position to the alcohol functionality (33ae) and the aromatic substituent R’ in the ligand system (Scheme 5). These catalysts show satisfactory catalytic activities in the C–N bond formation of aminopentenes 1be and aminohexene 2b. Conversions as high as 97% and high enantiomeric excesses (32d, max. 56% ee for 3d (Table 12, entry 4); 33b, up to 94% for 3d (Table 12, entries 15 and 16)) were observed in the catalytic synthesis of five-membered pyrrolidines 3be and E/Z-44a,b [98]. The authors also proposed a mechanism involving a highly ordered transition state and a concerted bond formation pathway. Variations in the temperature for the hydroamination of 1d using 33b as catalyst resulted only in minor changes in conversion and ee values (Table 12).
Overall, group IV metal catalysts 30ac and 33ae show high quantitative conversions with enatiomeric excesses as high as 98% for aminoalkenes 1af, 2c,d and 3840 and up to 99% for aminodialkenes 20ah and aminodialkynes 45ac. Both values are comparable for aminohexene substrates 2c,d and better for aminopentenes 1af, 38, 39 than those obtained by non-chiral-pool-derived catalysts which are mainly based on bisaryl-derived or salen-type ligands [81,84,99,100]. Comparisons of the hydroamination of aminoallenes 34a,b using Ti and Ta catalysts 21ac29ah with those applying non-chiral-pool-derived catalytic systems, which are mainly based on bisaryl-derived ligands, are more complicated due to differences in substrate screenings [86,101,102,103].

2.3. Early Main-Group Elements

In contrast to transition metals, which can appear in different oxidation states defining their reactivity by d-electrons, early main-group elements of group I and II are primarily characterized by mono-(alkaline) or dicationic (alkaline earth) ions depending on their outer shell s-electrons. Hence, main-group elements cannot easily switch between oxidation states, and therefore, catalysis with those metals is solely based on polar reaction mechanisms and Lewis-acid activations [104].

2.3.1. Alkaline Metals

Group I elements can be used as pre-catalysts both in their elementary as well as ionic form [8,105,106,107,108]. In enantioselective hydroamination reactions, solely lithium-based catalysts have been reported (intermolecular: 47, 48, Scheme 6, Table 13; 49a; 50af, Table 14; intramolecular: 49a,b, Table 15) [109,110,111,112,113,114].
Ates et al., first described the suitability of nBuLi (16 mol-%) in the catalytic high-yield synthesis of five- and six-membered N-heterocycles via the intramolecular hydroamination of non-activated aminoalkenes such as 1ac [105]. Shortly after, Hultzsch et al., reported the first Li-catalyzed enantioselective ring-closing reaction of 2,2-substituted pent-4-en-1-amines 1bd, 20a and 51 (Scheme 6, Table 13), providing the corresponding pyrrolidine derivatives 3bd, 44a and 52 [110]. As a catalyst, they used the dimeric, tetranuclear (S,S,S)-N,N’-dimethylpyrrolidinediamidobinaphthyl dilithium complex 47 (Scheme 6). The catalytic reactions succeeded with almost quantitative conversions and an enantiomeric excess of a max. 75% (Table 13, entry 5). The binaphthyl-centered chelate ligand in 47 is based on a DABN backbone to which two L-proline-derived moieties are attached. In the solid state, each of the four lithium ions possess different coordination environments, which exhibit a similar structure in solution [110]. It was found that only minor differences in enantiomeric excesses exist by various catalyst loadings and/or by the addition of coordinating solvents such as tetrahydrofuran. In contrast, these variations influenced the formation of the N-heterocyclic molecules more significantly. When, instead of the DABN backbone in 47, naphthyl was introduced in chiral 48, almost no enantiomeric excess and significant lower conversions in the formation of the respective N-heterocyclic compounds 3b,d was observed. No enantioselectivities were obtained by using the combination (−)-sparteine/LiN(SiMe3)2 (49a) as a pre-catalyst, albeit the observed conversion of 98% is comparable to 47 [110].
In 2007, Tomioka and his group discussed the intramolecular hydroamination of aminoalkenes 53a,b at −60 °C in toluene by applying in situ-produced catalytic systems 50af, containing diverse chiral bisoxazoline (= BOX) ligands (Table 14) [113]. In kinetically controlled catalytic reactions, almost quantitative yields and good ee values for the formation of N-heterocycles 54a,b was observed (Table 14). Within the catalytic active system, diisopropylamine acts as coordinating and proton-donating reagent. Out of the nine studied catalysts, 50ae contain amino acid-based chiral pool ligands, of which 50d converted substrate 53a into the corresponding six-membered tetrahydroisoquinoline 54a with 84% ee (Table 14, entry 11), while catalysts 50a,c produced five-membered isoindoline 54b with high regioselectivity and an ee of 84% (Table 14, entries 2 and 10) using 54b as substrate. In none of the cases the formation of endo-cyclized 55 was observed. On the other hand, the more rigid terpene camphor-modified catalyst 50f resulted in lower activities and ee values for the cyclization of aminopentene 53b, while for aminohexene 53a comparable results to 50d could be reached. However, both synthesized N-heterocycles 54a,b using 50f as catalyst possess the (R)-configuration instead of the (S)-products favored by 50ae. The best catalytic performance for the hydroamination of 53b was found for 50g having the non-chiral-pool D-isoleucine-derived groups attached to the BOX ligand (91% ee). Exchanging the solvent from toluene to tetrahydrofuran for catalysis resulted in the formation of both 54a,b as the kinetic and endo-cyclized 55a,b as the thermodynamic product [113].
The pre-catalyst 50a was selected for intramolecular hydroamination screening of aminoalkenes 53cf (Table 14, entries 9–16) and 56, of which the synthesis of (S)-laudanosine (57) is exemplarily shown in Scheme 7 [114]. Based on these studies, Yamamoto et al., synthesized (−)-javaberine A and (−)-epi-javaberine in an asymmetric total synthetic methodology with 76% ee using 50a as catalyst (Scheme 7) [115].
Catalyst 49a (vide supra) can be successfully used in the intermolecular hydroamination of olefins 58a,b with amines 59a,b resulting in ee values of up to 14% (Table 15, entry 14) and conversions from 38–71% [109], which contrasts the earlier discussed intramolecular hydroamination reactions showing no enantioselectivity. No enantiomeric excess was observed for 49b with (−)-α-isosparteine as ligand [109].
Outside of the chiral pool, Deschamp et al., reported a non-chiral-pool-based diamidobinaphthyl building block, allowing the variation in alkyl and methylene-aryl substituents at the amino functionalities [111,112]. Addition of LiCH2SiMe3 to the respective N,N’-disubstituted binaphthyldiamine resulted in the corresponding in situ-generated chiral lithium catalysts. Their use in the cyclization of conjugated 1,3-aminodienes 11 and 12a results in 13 and 14a with E/Z selectivities and ee values of up to 72%, while aminopentenes including 1bd, 2c and 51 provided 3bd, 4c and 52 with a maximum of 58% ee [111,112]. The enantioselectivities of the latter catalysts are for 1bee = −61%), 1cee = −63%) and 51ee = −15%), significantly lower, and for 1dee = 27%), higher, than for 47 [109]. To the best of our knowledge, no other chiral lithium catalysts were so far reported for the discussed substrates.

2.3.2. Alkaline Earth Metals

The first alkaline-earth-metal-mediated hydroamination catalysis was achieved by Hill et al., in 2005 using achiral heteroleptic β-diketiminato calcium complexes [6,34].
In 2009, Hultzsch et al., published the first chiral-pool-based alkaline earth magnesium catalysts (S,S,S)-61 and (R,S,S)-61 for the cyclization of aminoalkenes 1bd (Scheme 8, Table 16) [116]. In contrast to the lithium derivative 47 (vide supra), magnesium complexes (S,S,S)-61 and (R,S,S)-61 with their L-proline-derived axial chiral tetraamine ligands show moderate to high catalytic activities, but only limited enantiomeric excesses, with a maximum of 14% (Table 16, entry 6), due to the protolytic ligand exchange processes as typical for heteroleptic alkaline earth metal complexes. This solution-based phenomenon is known as the Schlenk equilibrium [117,118,119]. Within reference [116], the zinc derivatives of (S,S,S)-61 and (R,S,S)-61 were prepared. It was found that they are active hydroamination catalysts, yielding higher ee values (up to 29%) as their magnesium homologs [116].
In 2011, Sadow et al., described the synthesis of the magnesium complex 62a comprising a chiral, pseudo C3-symmetric, mono-anionic tris(oxazolinyl)borato ligand (Scheme 8) [120]. Its use in hydroamination reactions was also reported. The chirality of 62a results from three L-tert-leucine moieties. Due to the bulkiness of the respective ligand, the Schlenk equilibrium is hindered. Catalyst 62a produced good to excellent conversions in the intramolecular hydroamination of 1bd (Table 16, entries 7–12). The enantiomeric excesses, as compared with structurally similar complexes 31 and 32 (vide supra), were with a max. of 36% ee lower [120].
Overall, complexes (S,S,S)-61, (R,S,S)-61 and 62a, with their chiral-pool-derived motifs, show similar activities and conversions in the intramolecular hydroamination of 1bd as (R,R)-[{ONN}MgCH2Ph] (63, ONN = (R,R)-tert-butyl-2-(((-2-(dimethylamino)-cyclohexyl)(methyl)amino)methyl)-6-(triphenylsilyl)phenolato) [121], however, their enantiomeric excess is considerably lower (63, up to 90% ee for 3bd). Mechanistic studies on 63 were carried out comparing an σ-insertive mechanism against a concerted non-insertive one. DFT studies confirmed that proton-assisted concerted C–H/C–N bond formation is energetically not favored, contrary to the kinetically less demanding σ-insertive path [29,68,121]. The observed ee values for chiral-pool-derived magnesium-based catalysts are overall lower than those for non-chiral pool-derived 63.
In addition to 62a, the isostructural optically active calcium complex 62b was synthesized (Scheme 8) [120]. It was observed that within this species the Schlenk equilibrium is hindered in solution, as evidenced by NMR and IR studies. Catalyst 62b showed increased activities and quantitative conversions after minutes in comparison to 62a, but the stereoselectivity decreased to 18% ee for 3b,c (Table 16, entries 13 and 14) [120].
The first chiral-pool-derived (L-valine) calcium catalysts 64ad (Scheme 9) for enantioselective hydroamination reactions of aminoalkenes 1b,d originate from the Ward group, showing for 64a,b (Table 17, entries 1–4) similar activities and conversions (> 90%) when compared to 62b (Table 16, entries 13 and 14) [26]. Nevertheless, only an enantiomeric excess of 0–12% was observed for 3b,d. It should be mentioned that the para-fluorophenyl derivative 64d displayed no activity for substrates 1b,d, even after several weeks. Catalyst 64c (R = Ph) also revealed no activity when using 1b as substrate, while in the cyclization of 1d an 80% conversion occurred with 26% ee (Table 17, entry 6) [26]. This enantioselectivity signifies a notable increase in ee as compared with 64a,b (vide supra). It is also higher than the values reported for the calcium complex 62b and other non-chiral-pool-based BOX-containing calcium systems published by Buch and Harder [117,120].
In 2011, Wixey and Ward described the use of chiral-pool-based bisimidazoline calcium complexes 65ac in the catalytic cyclization of aminoalkenes 1b,d (Scheme 9) [122]. Like 64ad, complexes 65ac are derived from L-valine as a chirality inducing motif. It was shown that the ligand redistribution through the Schlenk equilibrium depends on the substituents R [122]. The measured ee values are within <12% low, however, they compare well to those for 64ac and the complexes containing other non-chiral-pool-derived BOX ligands [117,120].
In 2012, Nixon and Ward extended the series of bisoxazoline calcium complexes by bis(oxazolinylphenyl)amines(=BOPA), of which two of the three introduced BOPA-based catalysts (66a,b) (Scheme 9) derive from the chiral pool (L-valinol, L-phenylalaninol) [123]. In the enantioselective hydroamination of 1b, quantitative conversions and ee values of up to 26% ee could be achieved (Table 17, entry 22). The conversion for aminohexene 2d was determined to be 0–83% with enantiomeric excesses as high as 16% at 80 °C (Table 17, entry 26). A major improvement in stereoselectivity (as high as 50% ee for 1d) could be reached by employing BOPA ligands based on the non-natural, non-protogenic amino acid L-α-phenylglycine [123]. This significant improvement is due to the relatively slow ligand redistribution rate. A further increase in enantioselectivity to 56% ee for substrate 1d was reported by Harder et al., using non-chiral pool BINAM derivatives as bulky dianionic ligands [28].
While the use of free alcohols as ligands is rather common for early transition metals such as titanium or tantalum, their application in alkaline-earth-metal-based catalysts is rather limited, with phenoxyamine 63 from the Hultzsch group being the most prominent one in the case of magnesium [121]. However, no system is currently used which incorporates structural motifs derived from the chiral pool. In the case of calcium, alcoholates have, up to now, not been used at all. Therefore, we expanded on this type of binding site with the synthesis of an amino acid-derived tertiary alcohol (Scheme 10). This tridentate proto-ligand is accessible from L-isoleucine via a cascade of reductive aminations followed by a Grignard reaction. The transformation of aminoalkene 1d to the respective pyrrolidine 3d in a yield of >99% with an enantiomeric excess of 67% could be obtained by in situ formation of the catalyst 67 (Scheme 10) [124]. To the best of our knowledge, this enantioselectivity is the highest observed one for calcium-based species, including non-chiral-pool-derived catalysts, which greatly illustrates the potential of such compounds in the area of intramolecular hydroamination reactions.
Another type of catalysts for the enantioselective intramolecular hydroamination is based on the application of alkaline earth metals as pure Lewis-acidic metal centers. For example, non-basic calcium iodide as a Lewis acid and an external base for deprotonation can be used [27,125]. In general, it was found that the activities of alkaline earth metal iodides decrease in the series Ca > Sr >> Mg > Ba [27]. The proposed mechanism is shown in Figure 4. Coordination of the amino alkene to CaI2 acidifies one of the two NH2 protons. Deprotonation occurs by the tBuP4 phosphazene base, followed by the cyclization of the amino-olefin at the calcium metal center. After protonation of the formed N-heterocycle by [tBuP4H]+, pyrrolidine is released [125]. As chiral catalysts, (−)-fenchone-based 68 and non-chiral-pool-based BINOL-modified 69 were applied (Table 18). Catalyst 68 along with tBuP4 gave in the enantioselective hydroamination of aminoalkenes 1bd pyrrolidines 3bd with almost quantitative conversions and ee values reaching 15% (Table 18, entries 3 and 4). The experimentally determined ee values are generally lower than those for (S)-69, which achieves enantioselectivities of a max. 33% (Table 18, entries 7 and 8) [27].
In general, amido or benzyl strontium and barium complexes are also active in hydroamination reactions. Their overall activity is, however, lower than that of calcium, and no chiral catalyst based on the natural chiral pool have yet been reported [69,118,126,127].

3. Conclusions

The hydroamination reaction is an atom economical possibility for C–N bond formation, starting from common functional groups such as an amino functionality together with unsaturated C,C bonds. One of the main challenges arises from the high reaction barrier, which is attributed to the strong electronic repulsion of the participating groups and the symmetry forbidden nature of the [2 + 2] cyclization [5,6]. Hence, the application of catalysts is required. Over recent years, a vast amount of different catalysts were investigated, not only allowing for a maximum in yield, but also to ensure the stereoselectivity required for such transformations in case of asymmetric reaction products. Alkaline (Li), alkaline earth (Mg, Ca), rare earth (Y, La, Nd, Sm, Lu), group IV (Ti, Zr, Hf) metals, and tantalum are heavily applied in this field of research. With the rising demand for cheap and easily accessible catalysts, a promising strategy for the induction of chirality is the use of moieties obtainable from the chiral pool. In this case, the majority of ligand systems is derived from amino acids, while terpenes and alkaloids are only applied scarcely.
The chiral-pool-based building blocks can be incorporated into the ligands by different strategies, with the most prominent motifs being alcoholates, ethers and (bis)oxazolines. The best performing systems for titanium (27i, Phe-derived, Table 7), tantalum (24h, Phe-derived, Table 6) and calcium (67, Ile-derived, Scheme 10) are based on bulky amino-alcohols. For magnesium (62a, Val-derived, Table 16), yttrium (19b, Tle-derived, Table 4 and Table 11) and zirconium (30b, Val-derived, Table 10) BOX-containing ligands showed the best results, while for lanthanides, (−)-menthyl-substituted ansa-complexes performed well in the intramolecular hydroamination of aminoalkenes 1af and 2ad.
The resulting complexes are often of equal reactivity and selectivity than their non-chiral-pool-based, often bisaryl-derived counterparts. Therefore, they are a good alternative to established catalytic systems, with the exception of magnesium catalysts, which show significantly lower enantioselectivity compared to non-chiral-pool-derived ones, such as the phenoxyamine-based system 63. However, comparison between chiral-pool- and non-chiral-pool-derived titanium and tantalum catalysts is complicated due to the differences in substrate screening and the low amount of chiral catalytic systems found in the literature [101,102].
While a variety of different substrates, such as aminoalkenes, aminodialkenes and aminoallenes are investigated with great success, applications on higher functionalized substrates are only viewed scarcely and are often limited to a narrow number of model systems.
In summary, a range of different catalysts based on early metals is nowadays available for the application in hydroamination reactions, greatly enhanced by motifs originating from the chiral pool. Progress has been made towards high performant systems accompanied by a detailed understanding of their reaction behavior. Today, those catalysts are comparable to their more expensive heavy-transition-metal-based counterparts, often using cheaper and more accessible ligand systems. Their main limitation resides in the scarce substrate scope against which those catalysts were tested, greatly diminishing the possibilities arising from those catalysts. Therefore, upcoming challenges for early-metal-based hydroamination reactions need to shift from a pure catalyst development stage towards applications on more complex targets relevant for, e.g., pharmaceuticals or fine chemicals. By doing so, the extensive knowledge on early-metal-based catalysts can be harnessed and tailored to further enhance the toolkit in organic chemistry towards more (atom)economic and sustainable synthetic routes.

Author Contributions

All three authors contributed equally to the conceptualization, original draft preparation, as well as review and editing of the article. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kumar, D.; Kumar Jain, S. A Comprehensive Review of N-Heterocycles as Cytotoxic Agents. Curr. Med. Chem. 2016, 23, 4338–4394. [Google Scholar] [CrossRef] [PubMed]
  2. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Colonna, P.; Bezzenine, S.; Gil, R.; Hannedouche, J. Alkene Hydroamination via Earth-Abundant Transition Metal (Iron, Cobalt, Copper and Zinc) Catalysis: A Mechanistic Overview. Adv. Synth. Catal. 2020, 362, 1550–1563. [Google Scholar] [CrossRef]
  4. Hannedouche, J.; Collin, J.; Trifonov, A.; Schulz, E. Intramolecular Enantioselective Hydroamination Catalyzed by Rare Earth Binaphthylamides. J. Organomet. Chem. 2011, 696, 255–262. [Google Scholar] [CrossRef]
  5. Müller, T.E.; Hultzsch, K.C.; Yus, M.; Foubelo, F.; Tada, M. Hydroamination: Direct Addition of Amines to Alkenes and Alkynes. Chem. Rev. 2008, 108, 3795–3892. [Google Scholar] [CrossRef]
  6. Bestgen, S.; Roesky, P.W. Inramolecular Hydroamination of Alkenes. In Early Main Group Metal Catalysis; Harder, S., Ed.; Wiley-VCH: Weinheim, Germany, 2020; ISBN 9783527344482. [Google Scholar]
  7. Huang, L.; Arndt, M.; Gooßen, K.; Heydt, H.; Gooßen, L.J. Late Transition Metal-Catalyzed Hydroamination and Hydroamidation. Chem. Rev. 2015, 115, 2596–2697. [Google Scholar] [CrossRef]
  8. Seayad, J.; Tillack, A.; Hartung, C.G.; Beller, M. Base-Catalyzed Hydroamination of Olefins: An Environmentally Friendly Route to Amines. Adv. Synth. Catal. 2002, 344, 795–813. [Google Scholar] [CrossRef]
  9. Ye, Y.; Cao, J.; Oblinsky, D.G.; Verma, D.; Prier, C.K.; Scholes, G.D.; Hyster, T.K. Using Enzymes to Tame Nitrogen-Centered Radicals for Enantioselective Hydroamination. Nat. Chem. 2022, 15, 206–212. [Google Scholar] [CrossRef]
  10. He, Y.; Chen, J.; Jiang, X.; Zhu, S. Enantioselective NiH-Catalyzed Reductive Hydrofunctionalization of Alkenes. Chin. J. Chem. 2022, 40, 651–661. [Google Scholar] [CrossRef]
  11. Nuñez Bahena, E.; Schafer, L.L. From Stoichiometric to Catalytic E-H Functionalization by Non-Metallocene Zirconium Complexes-Recent Advances and Mechanistic Insights. ACS Catal. 2022, 12, 14934–14953. [Google Scholar] [CrossRef]
  12. Brunet, J.J.; Neibecker, D. Hydroamination of unsaturated Carbon bonds. In Catalytic Heterofunctionalization: From Hydroanimation to Hydrozirconation, 1st ed.; Togni, A., Grützmacher, H., Eds.; Wiley-VCH: Weinheim, Germany, 2001; ISBN 3527302344. [Google Scholar]
  13. Müller, T.E.; Beller, M. Metal-Initiated Amination of Alkenes and Alkynes. Chem. Rev. 1998, 98, 675–703. [Google Scholar] [CrossRef] [PubMed]
  14. Giofrè, S.; Molteni, L.; Beccalli, E.M. Asymmetric Pd(II)-Catalyzed C−O, C−N, C−C Bond Formation Using Alkenes as Substrates: Insight into Recent Enantioselective Developments. Eur. J. Org. Chem. 2022, 26, e202200976. [Google Scholar] [CrossRef]
  15. Jia, S.M.; Huang, Y.H.; Wang, F. Aminium-Radical-Mediated Intermolecular Hydroamination of Nonactivated Olefins. Synlett 2022, 34, 93–100. [Google Scholar] [CrossRef]
  16. Bernoud, E.; Lepori, C.; Mellah, M.; Schulz, E.; Hannedouche, J. Recent Advances in Metal Free- and Late Transition Metal-Catalysed Hydroamination of Unactivated Alkenes. Catal. Sci. Technol. 2015, 5, 2017–2037. [Google Scholar] [CrossRef]
  17. Beesley, R.M.; Ingold, C.K.; Thorpe, J.F. The Formation and Stability of Spiro-Compounds. Part I. Spiro-Compounds from Cyclo-Hexane. J. Chem. Soc. Trans. 1915, 107, 1080–1106. [Google Scholar] [CrossRef] [Green Version]
  18. Brunet, J.; Neibecker, D.; Niedercorn, F. Functionalisation of Alkenes: Catalytic Amination of Monoolefins. J. Mol. Catal. 1989, 49, 235–259. [Google Scholar] [CrossRef]
  19. Johns, A.M.; Sakai, N.; Ridder, A.; Hartwig, J.F. Direct Measurement of the Thermodynamics of Vinylarene Hydroamination. J. Am. Chem. Soc. 2006, 128, 9306–9307. [Google Scholar] [CrossRef]
  20. Hickinbottom, W.J. Reactions of Unsaturated Compounds. Part I. Addition of Arylamines to CycloHexene and 1:4- Dihydronaphthalene. J. Chem. Soc. 1932, 2646–2654. [Google Scholar] [CrossRef]
  21. Kozlov, N.S.; Gimpelevich, E. Catalytic Condensation of Acetylene with Aromatic Amines. IV. Condensation of Acetylene with Aniline and p-Toluidine in the Presence of Silver Nitrate. J. Gen. Chem. USSR 1936, 6, 1341–1345. [Google Scholar]
  22. Kozlov, N.S.; Bogdanovskaya, R. Catalytic Condensation of Acetylene with Aromatic Amines. V. Condensation of Acetylene with o- and p-Anisidine in the Presence of Cu2Cl2 and HgCl2. J. Gen. Chem. USSR 1936, 6, 1346–1348. [Google Scholar]
  23. Coulson, D.R. Catalytic Addition of Secondary Amines to Ethylene. Tetrahedron Lett. 1971, 12, 429–430. [Google Scholar] [CrossRef]
  24. Nobis, M.; Drießen-Hölscher, B. Recent developments in transition metal catalyzed intermolecular hydroamination reactions—A breakthrough? Angew. Chemie Int. Ed. 2001, 40, 3983–3985. [Google Scholar] [CrossRef]
  25. 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] [PubMed]
  26. Wixey, J.S.; Ward, B.D. Chiral Calcium Catalysts for Asymmetric Hydroamination/Cyclisation. Chem. Commun. 2011, 47, 5449–5451. [Google Scholar] [CrossRef]
  27. Stegner, P.C.; Fischer, C.A.; Nguyen, D.T.; Rösch, A.; Penafiel, J.; Langer, J.; Wiesinger, M.; Harder, S. Intramolecular Alkene Hydroamination with Hybrid Catalysts Consisting of a Metal Salt and a Neutral Organic Base. Eur. J. Inorg. Chem. 2020, 2020, 3387–3394. [Google Scholar] [CrossRef]
  28. Stegner, P.C.; Eyselein, J.; Ballmann, G.M.; Langer, J.; Schmidt, J.; Harder, S. Calcium Catalyzed Enantioselective Intramolecular Alkene Hydroamination with Chiral C2-Symmetric Bis-Amide Ligands. Dalt. Trans. 2021, 50, 3178–3185. [Google Scholar] [CrossRef]
  29. Zhang, X.; Tobisch, S.; Hultzsch, K.C. σ-Insertive Mechanism versus Concerted Non-Insertive Mechanism in the Intramolecular Hydroamination of Aminoalkenes Catalyzed by Phenoxyamine Magnesium Complexes: A Synthetic and Computational Study. Chem. Eur. J. 2015, 21, 7841–7857. [Google Scholar] [CrossRef]
  30. Michon, C.; Abadie, M.A.; Medina, F.; Agbossou-Niedercorn, F. Recent Metal-Catalysed Asymmetric Hydroaminations of Alkenes. J. Organomet. Chem. 2017, 847, 13–27. [Google Scholar] [CrossRef]
  31. McGrane, P.L.; Jensen, M.; Livinghouse, T. Intramolecular [2 + 2] Cycloadditions of Group IV Metal-Imido Complexes. Applications to the Synthesis of Dihydropyrrole and Tetrahydropyridine Derivatives. J. Am. Chem. Soc. 1992, 114, 5459–5460. [Google Scholar] [CrossRef]
  32. Walsh, P.J.; Baranger, A.M.; Bergman, R.G. Stoichiometric and Catalytic Hydroamination of Alkynes and Allene by Zirconium Bisamides Cp2Zr(NHR)2. J. Am. Chem. Soc. 1992, 114, 1708–1719. [Google Scholar] [CrossRef]
  33. Gagné, M.R.; Marks, T.J. Organolanthanide-Catalyzed Hydroamination. Facile, Regiospecific Cyclization of Unprotected Amino Olefins. J. Am. Chem. Soc. 1989, 111, 4108–4109. [Google Scholar] [CrossRef]
  34. Crimmin, M.R.; Casely, I.J.; Hill, M.S. Calcium-Mediated Intramolecular Hydroamination Catalysis. J. Am. Chem. Soc. 2005, 127, 2042–2043. [Google Scholar] [CrossRef] [PubMed]
  35. Pandey, S.K. BINOL: A Versatile Chiral Reagent. Synlett 2006, 2006, 3366–3367. [Google Scholar] [CrossRef] [Green Version]
  36. Berthod, M.; Mignani, G.; Woodward, G.; Lemaire, M. Modified BINAP: The How and the Why. Chem. Rev. 2005, 105, 1801–1836. [Google Scholar] [CrossRef]
  37. Blaser, H.U. The Chiral Pool as a Source of Enantioselective Catalysts and Auxiliaries. Chem. Rev. 1992, 92, 935–952. [Google Scholar] [CrossRef]
  38. Casiraghi, G.; Zanardi, F.; Rassu, G.; Spanu, P. Stereoselective Approaches to Bioactive Carbohydrates and Alkaloids-with a Focus on Recent Syntheses Drawing from the Chiral Pool. Chem. Rev. 1995, 95, 1677–1716. [Google Scholar] [CrossRef]
  39. Brill, Z.G.; Condakes, M.L.; Ting, C.P.; Maimone, T.J. Navigating the Chiral Pool in the Total Synthesis of Complex Terpene Natural Products. Chem. Rev. 2017, 117, 11753–11795. [Google Scholar] [CrossRef]
  40. Stout, C.N.; Renata, H. Reinvigorating the Chiral Pool: Chemoenzymatic Approaches to Complex Peptides and Terpenoids. Acc. Chem. Res. 2021, 54, 1143–1156. [Google Scholar] [CrossRef]
  41. Paek, S.M.; Jeong, M.; Jo, J.; Heo, Y.M.; Han, Y.T.; Yun, H. Recent Advances in Substrate-Controlled Asymmetric Induction Derived from Chiral Pool α-Amino Acids for Natural Product Synthesis. Molecules 2016, 21, 951. [Google Scholar] [CrossRef] [Green Version]
  42. Money, T.; Wong, M.K.C. The Use of Cyclic Monoterpenoids as Enantiopure Starting Materials in Natural Product Synthesis. Stud. Nat. Prod. Chem. 1995, 16, 123–288. [Google Scholar] [CrossRef]
  43. Koskinen, A.M.P. Chirospecific Synthesis: Catalysis and Chiral Pool Hand in Hand. Pure Appl. Chem. 2011, 83, 435–443. [Google Scholar] [CrossRef]
  44. Li, F.; Renata, H. A Chiral-Pool-Based Strategy to Access Trans-Syn-Fused Drimane Meroterpenoids: Chemoenzymatic Total Syntheses of Polysin, N-Acetyl-Polyveoline and the Chrodrimanins. J. Am. Chem. Soc. 2021, 143, 18280–18286. [Google Scholar] [CrossRef] [PubMed]
  45. Upadhyay, R.; Rana, R.; Maurya, S.K. Organocatalyzed C−N Bond-Forming Reactions for the Synthesis of Amines and Amides. ChemCatChem 2021, 13, 1867–1897. [Google Scholar] [CrossRef]
  46. Amadji, M.; Vadecard, J.; Plaquevent, J.; Duhamel, L.; Duhamel, P. First Catalytic Enantioselective Proton Abstraction. J. Am. Chem. Soc. 1996, 118, 12483–12484. [Google Scholar] [CrossRef]
  47. Krix, G.; Bommarius, A.S.; Drauz, K.; Kottenhahn, M.; Schwarm, M.; Kula, M.R. Enzymatic Reduction of α-Keto Acids Leading to L-Amino Acids, D- or L-Hydroxy Acids. J. Biotechnol. 1997, 53, 29–39. [Google Scholar] [CrossRef]
  48. Hannedouche, J.; Schulz, E. Hydroamination and Hydroaminoalkylation of Alkenes by Group 3-5 Elements: Recent Developments and Comparison with Late Transition Metals. Organometallics 2018, 37, 4313–4326. [Google Scholar] [CrossRef]
  49. Chen, Q.A.; Chen, Z.; Dong, V.M. Rhodium-Catalyzed Enantioselective Hydroamination of Alkynes with Indolines. J. Am. Chem. Soc. 2015, 137, 8392–8395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Otsuka, M.; Yokoyama, H.; Endo, K.; Shibata, T. Ru-Catalyzed β-Selective and Enantioselective Addition of Amines to Styrenes Initiated by Direct Arene-Exchange. Org. Biomol. Chem. 2012, 10, 3815–3818. [Google Scholar] [CrossRef]
  51. Flaget, A.; Zhang, C.; Mazet, C. Ni-Catalyzed Enantioselective Hydrofunctionalizations of 1,3-Dienes. ACS Catal. 2022, 12, 15638–15647. [Google Scholar] [CrossRef]
  52. Rocard, L.; Chen, D.; Stadler, A.; Zhang, H.; Gil, R.; Bezzenine, S.; Hannedouche, J. Earth-Abundant 3d TransitionMetal Catalysts for Hydroalkoxylation and Hydroamination of Unactivated Alkenes. Catalysts 2021, 11, 674. [Google Scholar] [CrossRef]
  53. Li, Y.; Marks, T.J. Organolanthanide-Catalyzed Intramolecular Hydroamination/Cyclization of Aminoalkynes. J. Am. Chem. Soc. 1996, 7863, 9295–9306. [Google Scholar] [CrossRef]
  54. Gagné, M.R.; Brard, L.; Conticello, V.P.; Giardello, M.A.; Stern, C.L.; Marks, T.J. Stereoselection Effects In the Catalytic Hydroaminatlon/Cyclization of Aminoolefins at Chiral Organolanthanide Centers. Organometallics 1992, 11, 2003–2005. [Google Scholar] [CrossRef]
  55. Li, Y.; Fu, P.; Marks, T.J. Organolanthanide-Catalyzed Carbon-Heteroatom Bond Formation. Observations on the Facile, Regiospecific Cyclization of Aminoalkynes. Organometallics 1994, 13, 4349–4440. [Google Scholar] [CrossRef]
  56. Manna, K.; Kruse, M.L.; Sadow, A.D. Concerted C-N/C-H Bond Formation in Highly Enantioselective Yttrium(III)-Catalyzed Hydroamination. ACS Catal. 2011, 1, 1637–1642. [Google Scholar] [CrossRef] [Green Version]
  57. Manna, K.; Xu, S.; Sadow, A.D. A Highly Enantioselective Zirconium Catalyst for Intramolecular Alkene Hydroamination: Significant Isotope Effects on Rate and Stereoselectivity. Angew. Chem. 2011, 123, 1905–1908. [Google Scholar] [CrossRef]
  58. Manna, K.; Everett, W.C.; Schoendorff, G.; Ellern, A.; Windus, T.L.; Sadow, A.D. Highly Enantioselective Zirconium-Catalyzed Cyclization of Aminoalkenes. J. Am. Chem. Soc. 2013, 135, 7235–7250. [Google Scholar] [CrossRef] [Green Version]
  59. Gagné, M.R.; Nolan, S.P.; Marks, T.J. Organolanthanide-Centered Hydroamination/Cyclizatlon of Aminoolefins. Expedient Oxidative Access to Catalytic Cycles. Organometallics 1990, 9, 1716–1718. [Google Scholar] [CrossRef]
  60. 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]
  61. 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]
  62. Molander, G.A.; Dowdy, E.D. Catalytic Intramolecular Hydroamination of Hindered Alkenes Using Organolanthanide Complexes. J. Org. Chem. 1998, 63, 8983–8988. [Google Scholar] [CrossRef]
  63. Molander, G.A.; Dowdy, E.D. Lanthanide-Catalyzed Hydroamination of Hindered Alkenes in Synthesis: Rapid Access to 10,11-Dihydro-5H-Dibenzo[a,d]Cyclohepten-5,10-Imines. J. Org. Chem. 1999, 64, 6515–6517. [Google Scholar] [CrossRef]
  64. Tobisch, S. Organolanthanide-Mediated Intermolecular Hydroamination of 1,3-Dienes: Mechanistic Insights from a Computational Exploration of Diverse Mechanistic Pathways for the Stereoselective Hydroamination of 1,3-Butadiene with a Primary Amine Supported by an Ansa-Neodymocene-Based Catalyst. Chem. Eur. J. 2005, 11, 6372–6385. [Google Scholar] [CrossRef]
  65. Tobisch, S. Mechanism and Exo-Regioselectivity of Organolanthanide-Mediated Intramolecular Hydroamination/Cyclization of 1,3-Disubstituted Aminoallenes: A Computational Study. Chem. Eur. J. 2006, 12, 2520–2531. [Google Scholar] [CrossRef] [PubMed]
  66. Motta, A.; Fragalà, I.L.; Marks, T.J. Energetics and Mechanism of Organolanthanide-Mediated Phosphinoalkene Hydrophosphination/Cyclization. A Density Functional Theory Analysis. Organometallics 2005, 24, 4995–5003. [Google Scholar] [CrossRef]
  67. Motta, A.; Fragalà, I.L.; Marks, T.J. Organolanthanide-Catalyzed Hydroamination/Cyclization Reactions of Aminoalkynes. Computational Investigation of Mechanism, Lanthanide Identity, and Substituent Effects for a Very Exothermic C-N Bond-Forming Process. Organometallics 2006, 25, 5533–5539. [Google Scholar] [CrossRef]
  68. Tobisch, S. Computational Mechanistic Elucidation of the Intramolecular Aminoalkene Hydroamination Catalysed by Iminoanilide Alkaline-Earth Compounds. Chem. Eur. J. 2015, 21, 6765–6779. [Google Scholar] [CrossRef]
  69. Arrowsmith, M.; Crimmin, M.R.; Barrett, A.G.M.; Hill, M.S.; Kociok-Köhn, G.; Procopiou, P.A. Cation Charge Density and Precatalyst Selection in Group 2-Catalyzed Aminoalkene Hydroamination. Organometallics 2011, 30, 1493–1506. [Google Scholar] [CrossRef]
  70. 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]
  71. 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]
  72. Tobisch, S. Intermolecular Hydroamination of Vinylarenes by Iminoanilide Alkaline-Earth Catalysts: A Computational Scrutiny of Mechanistic Pathways. Chem. Eur. J. 2014, 20, 8988–9001. [Google Scholar] [CrossRef]
  73. Giardello, M.A.; Conticello, V.P.; Brard, L.; Gagné, M.R.; Marks, T.J. Chiral Organolanthanides Designed for Asymmetric Catalysis. J. Am. Chem. Soc. 1994, 116, 10241–10254. [Google Scholar] [CrossRef]
  74. 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]
  75. Hong, S.; Tian, S.; Metz, M.V.; Marks, T.J. C2-Symmetric Bis (Oxazolinato) Lanthanide Catalysts for Enantioselective Intramolecular Hydroamination/Cyclization. J. Am. Chem. Soc. 2003, 125, 14768–14783. [Google Scholar] [CrossRef] [PubMed]
  76. Bennett, S.D.; Core, B.A.; Blake, M.P.; Pope, S.J.A.; Mountford, P.; Ward, B.D. Chiral Lanthanide Complexes: Coordination Chemistry, Spectroscopy, and Catalysis. Dalt. Trans. 2014, 43, 5871–5885. [Google Scholar] [CrossRef] [PubMed]
  77. Kim, H.; Kim, Y.K.; Shim, J.H.; Kim, M.; Han, M.; Livinghouse, T.; Lee, P.H. Internal Alkene Hydroaminations Catalyzed by Zirconium(IV) Complexes and Asymmetric Alkene Hydroaminations Catalyzed by Yttrium(III) Complexes. Adv. Synth. Catal. 2006, 348, 2609–2618. [Google Scholar] [CrossRef]
  78. Heck, R.; Schulz, E.; Collin, J.; Carpentier, J.F. Group 3 Metal Complexes Based on a Chiral Tetradentate Diamine-Diamide Ligand: Synthesis and Use in Polymerization of (d,l)-Lactide and Intramolecular Alkene Hydroamination Catalysis. J. Mol. Catal. A Chem. 2007, 268, 163–168. [Google Scholar] [CrossRef]
  79. Vitanova, D.V.; Hampel, F.; Hultzsch, K.C. (+)-Neomenthyl- and (-)-Phenylmenthyl-Substituted Cyclopentadienyl and Indenyl Yttrocenes as Catalysts in Asymmetric Hydroamination/Cyclization of Aminoalkenes (AHA). J. Organomet. Chem. 2007, 692, 4690–4701. [Google Scholar] [CrossRef]
  80. Huynh, K.; Anderson, B.K.; Livinghouse, T. Enantioselective Hydroamination/Cyclization of Aminoalkenes by (Bis)-C2 Symmetric and (Mono)-C2 Symmetric Anionic Tetraamide Complexes of La(III). Tetrahedron Lett. 2015, 56, 3658–3661. [Google Scholar] [CrossRef] [Green Version]
  81. Reznichenko, A.L.; Hultzsch, K.C. C2-Symmetric Zirconium Bis(Amidate) Complexes with Enhanced Reactivity in Aminoalkene Hydroamination. Organometallics 2010, 29, 24–27. [Google Scholar] [CrossRef]
  82. Chapurina, Y.; Guillot, R.; Lyubov, D.; Trifonov, A.; Hannedouche, J.; Schulz, E. LiCl-Effect on Asymmetric Intramolecular Hydroamination Catalyzed by Binaphthylamido Yttrium Complexes. J. Chem. Soc. Dalt. Trans. 2013, 42, 507–520. [Google Scholar] [CrossRef]
  83. Chapurina, Y.; Ibrahim, H.; Guillot, R.; Kolodziej, E.; Collin, J.; Trifonov, A.; Schulz, E.; Hannedouche, J. Catalytic, Enantioselective Intramolecular Hydroamination of Primary Amines Tethered to Di- and Trisubstituted Alkenes. J. Org. Chem. 2011, 76, 10163–10172. [Google Scholar] [CrossRef] [PubMed]
  84. Yonson, N.; Yim, J.C.H.; Schafer, L.L. Alkene Hydroamination with a Chiral Zirconium Catalyst. Connecting Ligand Design, Precatalyst Structure and Reactivity Trends. Inorganica Chim. Acta 2014, 422, 14–20. [Google Scholar] [CrossRef]
  85. Reznichenko, A.L.; Hultzsch, K.C. C1-Symmetric Rare-Earth-Metal Aminodiolate Complexes for Intra- and Intermolecular Asymmetric Hydroamination of Alkenes. Organometallics 2013, 32, 1394–1408. [Google Scholar] [CrossRef]
  86. Reznichenko, A.L.; Emge, T.J.; Audörsch, S.; Klauber, E.G.; Hultzsch, K.C.; Schmidt, B. Group 5 Metal Binaphtholate Complexes for Catalytic Asymmetric Hydroaminoalkylation and Hydroamination/Cyclization. Organometallics 2011, 30, 921–924. [Google Scholar] [CrossRef]
  87. Teng, H.L.; Luo, Y.; Wang, B.; Zhang, L.; Nishiura, M.; Hou, Z. Synthesis of Chiral Aminocyclopropanes by Rare-Earth-Metal-Catalyzed Cyclopropene Hydroamination. Angew. Chem. 2016, 128, 15632–15636. [Google Scholar] [CrossRef]
  88. Nguyen, H.N.; Lee, H.; Audörsch, S.; Reznichenko, A.L.; Nawara-Hultzsch, A.J.; Schmidt, B.; Hultzsch, K.C. Asymmetric Intra- and Intermolecular Hydroamination Catalyzed by 3,3′-Bis(Trisarylsilyl)- and 3,3′-Bis(Arylalkylsilyl)-Substituted Binaphtholate Rare-Earth-Metal Complexes. Organometallics 2018, 37, 4358–4379. [Google Scholar] [CrossRef]
  89. Chai, Z.; Hua, D.; Li, K.; Chu, J.; Yang, G. A Novel Chiral Yttrium Complex with a Tridentate Linked Amido-Indenyl Ligand for Intramolecular Hydroamination. Chem. Commun. 2014, 50, 177–179. [Google Scholar] [CrossRef]
  90. Hoover, J.M.; Petersen, J.R.; Pikul, J.H.; Johnson, A.R. Catalytic Intramolecular Hydroamination of Substituted Aminoallenes by Chiral Titanium Amino-Alcohol Complexes. Organometallics 2004, 23, 4614–4620. [Google Scholar] [CrossRef]
  91. Hickman, A.J.; Hughs, L.D.; Jones, C.M.; Li, H.; Redford, J.E.; Sobelman, S.J.; Kouzelos, J.A.; Johnson, A.R. Sterically Encumbered Chiral Amino Alcohols for Titanium-Catalyzed Asymmetric Intramolecular Hydroamination of Aminoallenes. Tetrahedron Asymmetry 2009, 20, 1279–1285. [Google Scholar] [CrossRef]
  92. Hansen, M.C.; Heusser, C.A.; Narayan, T.C.; Fong, K.E.; Hara, N.; Kohn, A.W.; Venning, A.R.; Rheingold, A.L.; Johnson, A.R. Asymmetric Catalytic Intramolecular Hydroamination of Aminoallenes by Tantalum Amidoalkoxide Complexes. Organometallics 2011, 30, 4616–4623. [Google Scholar] [CrossRef]
  93. Near, K.E.; Chapin, B.M.; McAnnally-Linz, D.C.; Johnson, A.R. Asymmetric Hydroamination of Aminoallenes Catalyzed by Titanium and Tantalum Complexes of Chiral Sulfonamide Alcohol Ligands. J. Organomet. Chem. 2011, 696, 81–86. [Google Scholar] [CrossRef]
  94. Sha, F.; Mitchell, B.S.; Ye, C.Z.; Abelson, C.S.; Reinheimer, E.W.; Lemagueres, P.; Ferrara, J.D.; Takase, M.K.; Johnson, A.R. Catalytic Intramolecular Hydroamination of Aminoallenes Using Titanium Complexes of Chiral, Tridentate, Dianionic Imine-Diol Ligands. Dalt. Trans. 2019, 48, 9603–9616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Sha, F.; Shimizu, E.A.; Slocumb, H.S.; Towell, S.E.; Zhen, Y.; Porter, H.Z.; Takase, M.K.; Johnson, A.R. Catalytic Intramolecular Hydroamination of Aminoallenes Using Titanium and Tantalum Complexes of Sterically Encumbered Chiral Sulfonamides. Dalt. Trans. 2020, 49, 12418–12431. [Google Scholar] [CrossRef] [PubMed]
  96. Manna, K.; Eedugurala, N.; Sadow, A.D. Zirconium-Catalyzed Desymmetrization of Aminodialkenes and Aminodialkynes through Enantioselective Hydroamination. J. Am. Chem. Soc. 2015, 137, 425–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Fok, E.Y.; Show, V.L.; Johnson, A.R. Intramolecular Hydroamination of Trisubstituted Aminoallenes Catalyzed by Titanium Complexes of Diaryl Substituted Tridentate Imine-Diols. Polyhedron 2021, 198, 115070. [Google Scholar] [CrossRef]
  98. Zhou, X.; Wei, B.; Sun, X.L.; Tang, Y.; Xie, Z. Asymmetric Hydroamination Catalyzed by a New Chiral Zirconium System: Reaction Scope and Mechanism. Chem. Commun. 2015, 51, 5751–5753. [Google Scholar] [CrossRef] [Green Version]
  99. Hussein, L.; Purkait, N.; Biyikal, M.; Tausch, E.; Roesky, P.W.; Blechert, S. Highly Enantioselective Hydroamination to Six-Membered Rings by Heterobimetallic Catalysts. Chem. Commun. 2014, 50, 3862–3864. [Google Scholar] [CrossRef] [Green Version]
  100. Wood, M.C.; Leitch, D.C.; Yeung, C.S.; Kozak, J.A.; Schafer, L.L. Chiral Neutral Zirconium Amidate Complexes for the Asymmetric Hydroamination of Alkenes. Angew. Chem. 2007, 119, 358–362. [Google Scholar] [CrossRef]
  101. Eisenberger, P.; Schafer, L.L. Catalytic Synthesis of Amines and N-Containing Heterocycles: Amidate Complexes for Selective C-N and C-C Bond-Forming Reactions. Pure Appl. Chem. 2010, 82, 1503–1515. [Google Scholar] [CrossRef]
  102. Wang, Q.; Song, H.; Zi, G. Synthesis, Structure, and Catalytic Activity of Group 4 Complexes with New Chiral Biaryl-Based NO2 Ligands. J. Organomet. Chem. 2010, 695, 1583–1591. [Google Scholar] [CrossRef]
  103. Zhang, F.; Song, H.; Zi, G. Synthesis and Catalytic Activity of Group 5 Metal Amides with Chiral Biaryldiamine-Based Ligands. Dalt. Trans. 2011, 40, 1547–1566. [Google Scholar] [CrossRef] [PubMed]
  104. Begouin, J.M.; Niggemann, M. Calcium-Based Lewis Acid Catalysts. Chem. Eur. J. 2013, 19, 8030–8041. [Google Scholar] [CrossRef] [PubMed]
  105. Ates, A.; Quinet, C. Efficient Intramolecular Hydroamination of Unactivated Alkenes Catalysed by Butyllithium. Eur. J. Org. Chem. 2003, 2003, 1623–1626. [Google Scholar] [CrossRef]
  106. Peng, X.; Kaga, A.; Hirao, H.; Chiba, S. Hydroamination of Alkenyl: N-Arylhydrazones Mediated by t -BuOK for the Synthesis of Nitrogen Heterocycles. Org. Chem. Front. 2016, 3, 609–613. [Google Scholar] [CrossRef]
  107. Chen, Z.Y.; Wu, L.Y.; Fang, H.S.; Zhang, T.; Mao, Z.F.; Zou, Y.; Zhang, X.J.; Yan, M. Intramolecular Hydroamidation of Ortho-Vinyl Benzamides Promoted by Potassium Tert-Butoxide/N,N-Dimethylformamide. Adv. Synth. Catal. 2017, 359, 3894–3899. [Google Scholar] [CrossRef]
  108. Zhang, Z.; Wang, J.; Guo, S.; Fan, J.; Fan, X. T - BuOK-Catalyzed Regio- and Stereoselective Intramolecular Hydroamination Reaction Leading to Phthalazinoquinazolinone Derivatives. J. Org. Chem. 2023, 8, 1282–1291. [Google Scholar] [CrossRef]
  109. Horrillo-Martínez, P.; Hultzsch, K.C.; Gil, A.; Branchadell, V. Base-Catalyzed Anti-Markovnikov Hydroamination of Vinylarenes—Scope, Limitations and Computational Studies. Eur. J. Org. Chem. 2007, 2007, 3311–3325. [Google Scholar] [CrossRef]
  110. Martínez, P.H.; Hultzsch, K.C.; Hampel, F. Base-Catalysed Asymmetric Hydroamination/Cyclisation of Aminoalkenes Utilising a Dimeric Chiral Diamidobinaphthyl Dilithium Salt. Chem. Commun. 2006, 37, 2221–2223. [Google Scholar] [CrossRef]
  111. Deschamp, J.; Olier, C.; Schulz, E.; Guillot, R.; Hannedouche, J.; Collin, J. Simple Chiral Diaminobinaphthyl Dilithium Salts for Intramolecular Catalytic Asymmetric Hydroamination of Amino-1,3-Dienes. Adv. Synth. Catal. 2010, 352, 2171–2176. [Google Scholar] [CrossRef]
  112. Deschamp, J.; Collin, J.; Hannedouche, J.; Schulz, E. Easy Routes towards Chiral Lithium Binaphthylamido Catalysts for the Asymmetric Hydroamination of Amino-1,3-Dienes and Aminoalkenes. Eur. J. Org. Chem. 2011, 2011, 3329–3338. [Google Scholar] [CrossRef]
  113. Ogata, T.; Ujihara, A.; Tsuchida, S.; Shimizu, T.; Kaneshige, A.; Tomioka, K. Catalytic Asymmetric Intramolecular Hydroamination of Aminoalkenes. Tetrahedron Lett. 2007, 48, 6648–6650. [Google Scholar] [CrossRef]
  114. Ogata, T.; Kimachi, T.; Yamada, K.I.; Yamamoto, Y.; Tomioka, K. Catalytic Asymmetric Synthesis of (S)-Laudanosine by Hydroamination. Heterocycles 2012, 86, 469–485. [Google Scholar] [CrossRef]
  115. Uenishi, S.; Kakigi, R.; Hideshima, K.; Miyawaki, A.; Matsuoka, J.; Ogata, T.; Tomioka, K.; Yamamoto, Y. Asymmetric Total Synthesis of (−)-Javaberine A and (−)-Epi-Javaberine A Based on Catalytic Intramolecular Hydroamination of N-Methyl-2-(2-Styrylaryl)Ethylamine. Tetrahedron 2021, 90, 132165. [Google Scholar] [CrossRef]
  116. Horrillo-Martínez, P.; Hultzsch, K.C. Intramolecular Hydroamination/Cyclization of Aminoalkenes Catalyzed by Diamidobinaphthyl Magnesium- and Zinc-Complexes. Tetrahedron Lett. 2009, 50, 2054–2056. [Google Scholar] [CrossRef]
  117. Buch, F.; Harder, S. A Study on Chiral Organocalcium Complexes: Attempts in Enantioselective Catalytic Hydrosilylation and Intramolecular Hydroamination of Alkenes. Z. Naturforsch. B 2008, 63, 169–177. [Google Scholar] [CrossRef]
  118. Liu, B.; Roisnel, T.; Carpentier, J.F.; Sarazin, Y. Cyclohydroamination of Aminoalkenes Catalyzed by Disilazide Alkaline-Earth Metal Complexes: Reactivity Patterns and Deactivation Pathways. Chem. Eur. J. 2013, 19, 2784–2802. [Google Scholar] [CrossRef]
  119. Schlenk, W.; Schlenk, W. Über Die Konstitution Der Grignardschen Magnesiumverbindungen. Berichte Dtsch. Chem. Gesellschaft 1929, 62, 920–924. [Google Scholar] [CrossRef]
  120. Neal, S.R.; Ellern, A.; Sadow, A.D. Optically Active, Bulky Tris(Oxazolinyl)Borato Magnesium and Calcium Compounds for Asymmetric Hydroamination/Cyclization. J. Organomet. Chem. 2011, 696, 228–234. [Google Scholar] [CrossRef]
  121. Zhang, X.; Emge, T.J.; Hultzsch, K.C. A Chiral Phenoxyamine Magnesium Catalyst for the Enantioselective Hydroamination/Cyclization of Aminoalkenes and Intermolecular Hydroamination of Vinyl Arenes. Angew. Chemie Int. Ed. 2012, 51, 394–398. [Google Scholar] [CrossRef]
  122. Wixey, J.S.; Ward, B.D. Modular Ligand Variation in Calcium Bisimidazoline Complexes: Effects on Ligand Redistribution and Hydroamination Catalysis. Dalt. Trans. 2011, 40, 7693–7696. [Google Scholar] [CrossRef]
  123. Nixon, T.D.; Ward, B.D. Calcium Amido-Bisoxazoline Complexes in Asymmetric Hydroamination/Cyclisation Catalysis. Chem. Commun. 2012, 48, 11790–11792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Notz, S. Ph.D. Thesis, TU Chemnitz: Chemnitz, Germany, ongoing.
  125. Penafiel, J.; Maron, L.; Harder, S. Early Main Group Metal Catalysis: How Important Is the Metal? Angew. Chemie 2015, 54, 201–206. [Google Scholar] [CrossRef] [PubMed]
  126. Jenter, J.; Köppe, R.; Roesky, P.W. 2,5-Bis{N-(2,6-Diisopropylphenyl)Iminomethyl}pyrrolyl Complexes of the Heavy Alkaline Earth Metals: Synthesis, Structures, and Hydroamination Catalysis. Organometallics 2011, 30, 1404–1413. [Google Scholar] [CrossRef]
  127. Romero, N.; Roşca, S.C.; Sarazin, Y.; Carpentier, J.F.; Vendier, L.; Mallet-Ladeira, S.; Dinoi, C.; Etienne, M. Highly Fluorinated Tris(Indazolyl)Borate Silylamido Complexes of the Heavier Alkaline Earth Metals: Synthesis, Characterization, and Efficient Catalytic Intramolecular Hydroamination. Chem. Eur. J. 2014, 21, 4115–4125. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Examples of chiral pool motifs within the three important natural product classes for non-phosphine-based ligand systems.
Figure 1. Examples of chiral pool motifs within the three important natural product classes for non-phosphine-based ligand systems.
Molecules 28 02702 g001
Figure 2. The σ-insertive (a) and concerted proton-triggered cyclization mechanism (b) for intramolecular hydroaminations [29,33,54,59,61,64,65,68,71,72].
Figure 2. The σ-insertive (a) and concerted proton-triggered cyclization mechanism (b) for intramolecular hydroaminations [29,33,54,59,61,64,65,68,71,72].
Molecules 28 02702 g002
Scheme 1. Catalytic asymmetric hydroamination of aminoalkenes 1a,b and 2a,b using chiral lanthanocene complexes (S)/(R)-5an and (S)-6ac [54,60,73]. (For more details concerning catalysis data see Table 1).
Scheme 1. Catalytic asymmetric hydroamination of aminoalkenes 1a,b and 2a,b using chiral lanthanocene complexes (S)/(R)-5an and (S)-6ac [54,60,73]. (For more details concerning catalysis data see Table 1).
Molecules 28 02702 sch001
Scheme 3. Catalytic asymmetric hydroamination of aminoalkenes 1af, 2c,d and 3840 using chiral group IV complexes 30ac and 31 [56,57,96]. (For more details concerning catalysis data see Table 10).
Scheme 3. Catalytic asymmetric hydroamination of aminoalkenes 1af, 2c,d and 3840 using chiral group IV complexes 30ac and 31 [56,57,96]. (For more details concerning catalysis data see Table 10).
Molecules 28 02702 sch003
Scheme 4. Catalytic asymmetric hydroamination of aminoalkenes 20ah and 45ac using chiral group IV complexes 30b,c and 31 (Scheme 3) and the yttrium complexes 17a, 18 and 19b (Scheme 2) [56,57,58,96]. (For more details concerning catalysis data and assignments of R and R’ see Table 11).
Scheme 4. Catalytic asymmetric hydroamination of aminoalkenes 20ah and 45ac using chiral group IV complexes 30b,c and 31 (Scheme 3) and the yttrium complexes 17a, 18 and 19b (Scheme 2) [56,57,58,96]. (For more details concerning catalysis data and assignments of R and R’ see Table 11).
Molecules 28 02702 sch004
Figure 3. Proposed catalytic cycle explaining the concentration-dependent Z/E selectivity with an exchanging two-site catalytic model [96].
Figure 3. Proposed catalytic cycle explaining the concentration-dependent Z/E selectivity with an exchanging two-site catalytic model [96].
Molecules 28 02702 g003
Scheme 5. Catalytic asymmetric hydroamination of aminoalkenes 1be, 2d and 20a,b using chiral zirconium complexes 32ag and 33ae [98]. (For more details concerning catalysis data see Table 12).
Scheme 5. Catalytic asymmetric hydroamination of aminoalkenes 1be, 2d and 20a,b using chiral zirconium complexes 32ag and 33ae [98]. (For more details concerning catalysis data see Table 12).
Molecules 28 02702 sch005
Scheme 6. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1bd, 20a and 51 using lithium-based catalysts 47 and 48 [110]. (For more details concerning catalysis data see Table 13).
Scheme 6. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1bd, 20a and 51 using lithium-based catalysts 47 and 48 [110]. (For more details concerning catalysis data see Table 13).
Molecules 28 02702 sch006
Scheme 7. Catalytic asymmetric hydroamination of 56 [114,115].
Scheme 7. Catalytic asymmetric hydroamination of 56 [114,115].
Molecules 28 02702 sch007
Scheme 8. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1bd using magnesium-based catalysts (S,S,S)-61, (R,S,S)-61 and 62a and calcium complex 62b [116,120]. (For more details concerning catalysis data see Table 16).
Scheme 8. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1bd using magnesium-based catalysts (S,S,S)-61, (R,S,S)-61 and 62a and calcium complex 62b [116,120]. (For more details concerning catalysis data see Table 16).
Molecules 28 02702 sch008
Scheme 9. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1b,d and 2d using calcium-based catalysts 64ad, 65ac and 66a,b [26,122,123]. (For more details concerning catalysis data see Table 17).
Scheme 9. Catalytic asymmetric intramolecular hydroamination of aminoalkenes 1b,d and 2d using calcium-based catalysts 64ad, 65ac and 66a,b [26,122,123]. (For more details concerning catalysis data see Table 17).
Molecules 28 02702 sch009
Scheme 10. Catalytic asymmetric intramolecular hydroamination of aminopentene 1d using the calcium-based catalytic system 67 [124].
Scheme 10. Catalytic asymmetric intramolecular hydroamination of aminopentene 1d using the calcium-based catalytic system 67 [124].
Molecules 28 02702 sch010
Figure 4. Catalytic cycle of the intramolecular hydroamination of aminoalkenes using CaI2 as catalyst and tBuP4 as external base [27].
Figure 4. Catalytic cycle of the intramolecular hydroamination of aminoalkenes using CaI2 as catalyst and tBuP4 as external base [27].
Molecules 28 02702 g004
Table 1. Catalytic asymmetric hydroamination reactions of 1a,b and 2a,b using chiral rare earth metal complexes (S)/(R)-5an and (S)-6ac.
Table 1. Catalytic asymmetric hydroamination reactions of 1a,b and 2a,b using chiral rare earth metal complexes (S)/(R)-5an and (S)-6ac.
EntryCat.MER*Substr.Prod.T a
[°C]
ee
[%] b,c
Ref.
1(R)-5a,b dYN,CH b(−)-menthyl1a3a2569 (+)[73]
21b3b2543 (+)[73]
3(R)-5cYN(+)-neomenthyl1a3a2550 (−)[73]
41b3b2540 (−)[73]
5(R/S)-5d eYCH(+)-neomenthyl1a3a2547 (−)[73]
61b3b2536 (−)[73]
7(R)-5eLaN(+)-neomenthyl1a3a2531 (−)[54,73]
81b3b2514 (−)[54,73]
9(R,S)-5fLaCH(+)-neomenthyl1a3a2536 (−)[73]
10(R/S)-5g eNdCH(+)-neomenthyl1a3a2555 (−)[73]
11 064 (−)[73]
121b3b−2061 (−)[73]
13(S)-5h,i dSmN,CH b(−)-menthyl1a3a2562 (+)[54,73]
14 072 (+)[54,73]
151b3b2553 (+)[54,73]
16 061 (+)[54,73]
17 −3074 (+)[54,73]
182b4b2515 (−)[54,73]
19(R)-5hSmN(−)-menthyl1a3a2560 (+)[73]
20(S)-5jSmN(+)-neomenthyl1a3a2555 (−)[73]
21(R)-5j,k dSmN,CH b(+)-neomenthyl1a3a2552 (−)[54,73]
22 058 (−)[54,73]
231b3b2551 (−)[54,73]
24 054 (−)[54,73]
25 −3064 (−)[54,73]
262b4b2517 (−)[54,73]
27(R/S)-5k eSmCH(+)-neomenthyl1a3a2561 (−)[73]
28(R)-5lLuCH(−)-menthyl1b3b2529 (−)[73]
29(R)-5mLuN(+)-neomenthyl1b3b2540(+)[73]
30(R/S)-5n eLuCH(+)-neomenthyl1a3a2529 (+)[73]
311b3b2536 (+)[73]
32(S)-6aYN(−)-menthyl1a3a605 (+)[60]
331b3b2517 (+)[60]
342a4a603 (+)[60]
352b4b6054 (+)[60]
36 2567 (+)[60]
37(S)-6bSmN(−)-menthyl1a3a6037 (+)[60]
38 2546 (+)[60]
391b3b2532 (+)[60]
402a4a6010 (+)[60]
412b4b6043 (+)[60]
42 4041 (+)[60]
43 2541 (+)[60]
44(S)-6cLuN(−)-menthyl1a3a6016 (−)[60]
451b3b252 (+)[60]
462b4b6015 (+)[60]
a M = La, Nd, Sm, t = 1–12 h; M = Y, Lu, t = 1–3 d. b Enantiomeric excesses (=ee), 100% Conversion and >95% regiospecifity as determined by GLC and 1H NMR measurements. c (−) = (R)-(−)-2-Methylpyrrolidine. d Both derivatives show the same behavior. e Mixture of (R)- and (S)-diastereomers.
Table 2. Catalytic asymmetric hydroamination of E-7a,b and Z-8 using chiral rare earth metal complexes (S)-5h and (S)-6a,b a.
Table 2. Catalytic asymmetric hydroamination of E-7a,b and Z-8 using chiral rare earth metal complexes (S)-5h and (S)-6a,b a.
Molecules 28 02702 i001
EntryCat.MR*Substr.Prod.T
[°C]
TOF
[h−1]
ee
[%] b
Ref.
1(S)-5hSm(−)-menthylE-7a9a800.2628 (+)[70]
2E-7b9b800.1532 (+)[70]
3Z-810800.1616 (+)[70]
3(S)-6aY(+)-neomenthylE-7a9a1000.0726 (+)[70]
4E-7b9b1000.0628 (+)[70]
5Z-810100 c0.3058 (−)[70]
6 80 d0.1664 (−)[70]
7 60 c0.0368 (−)[70]
8(S)-6bSm(+)-neomenthylE-7a9a800.1824 (+)[70]
9 E-7b9b800.0622 (+)[70]
10Z-810800.1116 (+)[70]
a 5 mol-% catalyst (unless otherwise noted). b Enantiomeric excesses (=ee) at >95% conversion determined by chiral HPLC analysis. c In o-xylene-d10. d 20 mol-% catalyst in benzene-d6.
Table 3. Catalytic asymmetric hydroamination of 11 and 12a,b using chiral rare earth metal complexes (S)-5a,b, (S)-5h and (S)-6b a.
Table 3. Catalytic asymmetric hydroamination of 11 and 12a,b using chiral rare earth metal complexes (S)-5a,b, (S)-5h and (S)-6b a.
Molecules 28 02702 i002
EntryCat.Substr.Prod.T
[°C]
Ratio
E/Z b
ee
[%] c
Ref.
1(S)-5a,b11E/Z-132598:241[61]
2(S)-5h11E/Z-132398:225 d[61]
312aE/Z-14a2598:237 (R)[61]
4(S)-6b11E/Z-132593:723[25,61]
512aE/Z-14a2597:363 (R)[25,61]
6 0 e96:464 (R)[61]
7 25 f96:464 (R)[61]
8 0 f95:569 (R)[25,61]
9 25 g97:364 (R)[61]
10 0 g97:371 (R)[61]
11 25 h97:365 (R)[61]
1212bE/Z-14b2596:419[61]
13 0 f93:724[61]
a Conditions: 4–7 mol-% or 20 mol-% catalyst, ∼0.6 mL of solvent. b Determined by GC-MS. c Enantiomeric excesses (=ee) determined by optical rotation of the HCl salt of the hydrogenated product. Absolute configuration of the major isomer. d Determined by chiral HPLC analysis of the hydrogenated product. e In toluene-d8. f In cyclohexane-d12. g In methylcyclohexane-d14. h In pentane.
Table 5. Catalytic asymmetric hydroamination of 34a using chiral rare earth metal complexes 21ac and 22ac.
Table 5. Catalytic asymmetric hydroamination of 34a using chiral rare earth metal complexes 21ac and 22ac.
Molecules 28 02702 i003
EntryCat.R’ at
[h] b
35
[%] c
Z-36
[%] c
ee
[%] d
E-36
[%] c
ee
[%] d
Ref.
121aiPr4820411394[90]
221bcC6H119119410415[90]
321c2-Ad9424400365[90]
422aiPr2233346334[90]
522bcC6H112232338355[90]
622c2-Ad43224273616[90]
a 2-Ad = 2-Adamantyl. b Conditions: 5 mol-% catalyst, T = 110 °C. c Yields determined at 95% conversion. d Enantiomeric excesses (=ee) determined of the benzamide derivative, determined by chiral GC ± 2%.
Table 6. Catalytic asymmetric hydroamination of 34b using chiral titanium 21ac, 22ac, 23ai and 25ai and tantalum catalysts 24al and 26al a.
Table 6. Catalytic asymmetric hydroamination of 34b using chiral titanium 21ac, 22ac, 23ai and 25ai and tantalum catalysts 24al and 26al a.
Molecules 28 02702 i004
EntryCat.MnR*R bR’t
[h]
Conv.
[%]
ee
[%] c
Ref.
121aTi4iPriPrH18>954 (+)[90,91]
221bTi4iPrcC6H11H18>954 (+)[90,91]
321cTi4iPr2-AdH17>955 (+)[90,91]
422aTi4BniPrH16>952 (+)[90,91]
522bTi4BncC6H11H16>956 (+)[90,91]
622cTi4Bn2-AdH20>9515 (+)[90,91]
723aTi4iPriPrMe181002 (−)[91]
823bTi4iPriPrnBu181005 (+)[91]
923c dTi4iPriPrPh18100n.a.[91]
1023dTi4iPrcC6H11Me181001 (−)[91]
1123eTi4iPrcC6H11nBu181001 (+)[91]
1223fTi4iPrcC6H11Ph181005 (+)[91]
1323gTi4iPr2-AdMe181002 (+)[91]
1423hTi4iPr2-AdnBu1810010 (+)[91]
1523iTi4iPr2-AdPh181000[91]
1624aTa5iPriPrH1627810 (−)[92]
1724bTa5iPriPrMe231002 (−)[92]
1824cTa5iPriPrnBu983924 (−)[92]
1924d dTa5iPriPrPh1510029 (−)[92]
2024eTa5iPrcC6H11H286647 (−)[92]
2124fTa5iPrcC6H11Me46903 (−)[92]
2224gTa5iPrcC6H11nBu431008 (−)[92]
2324hTa5iPrcC6H11Ph1810074 (−)[92]
2424iTa5iPr2-AdH336652 (−)[92]
2524jTa5iPr2-AdMe334443 (−)[92]
2624kTa5iPr2-AdnBu1341006 (−)[92]
2724lTa5iPr2-AdPh429137 (−)[92]
2825aTi4BniPrMe181004 (−)[91]
2925bTi4BniPrnBu181003 (+)[91]
3025cTi4BniPrPh1810016 (+)[91]
3125dTi4BncC6H11Me181001 (−)[91]
3225eTi4BncC6H11nBu181005 (−)[91]
3325fTi4BncC6H11Ph1810016 (+)[91]
3425gTi4Bn2-AdMe181002 (−)[91]
3525hTi4Bn2-AdnBu181001 (+)[91]
3625iTi4Bn2-AdPh181007 (+)[91]
3726aTa5BniPrH692413 (−)[92]
3826bTa5BniPrMe1510046 (−)[92]
3926cTa5BniPrnBu231001 (−)[92]
4026d dTa5BniPrPh1810065 (−)[92]
4126eTa5BncC6H11H503513 (−)[92]
4226fTa5BncC6H11Me6510035 (−)[92]
4326gTa5BncC6H11nBu2310032 (−)[92]
4426hTa5BncC6H11Ph1610080 (−)[92]
4526iTa5Bn2-AdH116336 (−)[92]
4626jTa5Bn2-AdMe241002 (−)[92]
4726kTa5Bn2-AdnBu11510024 (−)[92]
4826lTa5Bn2-AdPh2302825 (−)[92]
a Conditions: 5 mol-% catalyst, 135 °C. b 2-Ad = 2-Adamantyl. c Enantiomeric excesses (=ee) determined by chiral shift 1H NMR spectroscopy using (R)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers, or determined by GC using Chiraldex B-DM (±5%). d Ligand for 26d obtained in 50% ee. n.a.—not applicable.
Table 7. Catalytic asymmetric hydroamination of 34b using chiral titanium 27aj and tantalum catalysts 28aj a.
Table 7. Catalytic asymmetric hydroamination of 34b using chiral titanium 27aj and tantalum catalysts 28aj a.
Molecules 28 02702 i005
EntryCat.MnRArRT a
[°C]
t
[h]
Conv.
[%]
ee b
[%]
Ref.
127aTi44-CH3H125 c513811 (+)[93]
2 13540653 (+)[93]
327bTi44-CF3H125 c51619 (+)[93]
4 135 c18734 (+)[93]
527cTi43,5-di-CF3H12549290[93]
6 13515857 (+)[93]
727dTi44-CH3Me12540334 (+)[93]
8 13518956 (+)[93]
927eTi44-CF3Me125 c49558 (+)[93]
10 135 c40617 (+)[93]
1127fTi43,5-di-CF3Me12527455 (+)[93]
12 13537802 (+)[93]
1327gTi44-CH3Ph135188218 (−)[95]
1427hTi44-CF3Ph135329524 (−)[95]
1527iTi43,5-di-CF3Ph1351061841 (−)[95]
16 Ph135 c2010021 (−)[95]
1727jTi42,4,6-tri-CH3Ph135757127 (−)[95]
1828aTa54-CH3H1251322024 (−)[93]
19 1351246021 (−)[93]
2028bTa54-CF3H1251158528 (−)[93]
21 13569325 (−)[93]
2228cTa53,5-(CF3)2H1257110034 (−)[93]
23 135693417 (−)[93]
2428dTa54-CH3Me125178824 (−)[93]
25 1351810015 (−)[93]
2628eTa54-CF3Me1251510020 (−)[93]
27 1351810026 (−)[93]
2828fTa53,5-(CF3)2Me1251510023 (−)[93]
29 135191007 (−)[93]
3028gTa54-CH3Ph1355710037 (−)[95]
3128hTa54-CF3Ph1354910033 (−)[95]
3228iTa53,5-di-CF3Ph1352310035 (−)[95]
3328jTa52,4,6-tri-CH3Ph1354710039 (−)[95]
a 5 mol-% catalyst. b Enantiomeric excesses (=ee) determined by chiral shift 1H NMR spectroscopy using (R)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers. c 10 mol-% catalyst.
Table 8. Catalytic asymmetric hydroamination of 34a using chiral titanium catalysts 27gj, 28gj and 29ae a.
Table 8. Catalytic asymmetric hydroamination of 34a using chiral titanium catalysts 27gj, 28gj and 29ae a.
Molecules 28 02702 i006
EntryCat.MnR*RArRt
[h]
Conv.
[%]
35
[%]
E-36
[%]
ee
[%] b
Z-36
[%]
ee
[%] b
Ref.
127gTi4Bn4-CH3Ph648460617185[95]
227hTi4Bn4-CF3Ph4095734461820[95]
327iTi4Bn3,5-di-CF3Ph2398825551045[95]
427jTi4Bn2,4,6-tri-CH3Ph308359419208[95]
528gTa5Bn4-CH3Ph30964910403717[95]
628hTa5Bn4-CF3Ph30975010253616[95]
728iTa5Bn3,5-di-CF3Ph2398481883221[95]
828jTa5Bn2,4,6-tri-CH3Ph23954910403615[95]
929aTi-BnHMe437646179228[94]
1029bTi-BnHPh318740141362[94]
1129cTi-Bn3,5-di-tBuPh54967283177[94]
1229dTi-Bn5-FPh549144132381[94]
1329eTi-iPr5-FPh548842132392[94]
a 5 mol-% catalyst. b Enantiomeric excesses (=ee) determined by chiral shift 1H NMR spectroscopy using (R)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers.
Table 9. Catalytic asymmetric hydroamination of 34b using chiral catalysts 29ah a.
Table 9. Catalytic asymmetric hydroamination of 34b using chiral catalysts 29ah a.
Molecules 28 02702 i007
EntryCat.R*RArR[cat.]
[%]
t
[h]
Conv.
[%]
ee
[%] b
Ref.
129aBnHMe567711 (−)[94]
229bBnHPh543866 (−)[94]
329cBn3,5-di-tBuPh5348717 (−)[94]
429dBn5-FPh581731 (−)[94]
529eiPr5-FPh565743 (−)[94]
629fBn3,5-di-C6H5Ph20181008 (−)[97]
7 10662515 (−)[97]
8 2018 c10021 (−)[97]
9 2066 d10019 (−)[97]
10 2018 e10022 (−)[97]
1129gBn3,5-di-(4-C6H4(CF3))Ph20131008 (−)[97]
12 10131006 (−)[97]
13 2021 c1006 (−)[97]
14 2019 d1007 (−)[97]
15 2019 e1008 (−)[97]
1629hBn3,5-di-(3,5-C6H3(CF3)2)Ph20181006 (−)[97]
17 10184012 (−)[97]
18 2018 c1005 (−)[97]
19 1018 c2518 (−)[97]
20 2018 d1007 (−)[97]
21 1066 d6321 (−)[97]
22 2018 e1006 (−)[97]
a Conditions: 5 mol-% catalyst (29ae) or 20 mol-% (29fh), T = 135 °C. b Enantiomeric excesses (=ee) determined by chiral shift 1H NMR spectroscopy using (R)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers. c T = 125 °C. d T = 115 °C. e T = 105 °C.
Table 10. Catalytic asymmetric hydroamination of 1af, 2c,d and 3840 using chiral group IV catalysts 30ac and 31 a.
Table 10. Catalytic asymmetric hydroamination of 1af, 2c,d and 3840 using chiral group IV catalysts 30ac and 31 a.
EntryCat.MR*Substr.Prod.Solvent bT
[°C]
t
[h]
Conv.
[%]
ee
[%] c
Ref.
130aTiiPr1c3cbenzene-d6251207583 (R)[58]
2 1d3dbenzene-d6251209376 (R)[58]
330bZriPr1a3abenzene-d61101524n.d.[57,58]
4 1b3bbenzene-d62578989 (R)[57]
5 benzene2578989 (R)[58]
6 toluene-d8−301929593 (R)[58]
7 1c3cbenzene-d6251.25>9590 (R)[57]
8 benzene251.259690 (R)[58]
9 toluene-d825 d6.5>9590 (R)[57]
10 thf-d80119394 (R)[57,58]
11 1d3dbenzene-d6251.25>9593 (R)[57]
12 benzene251.259593 (R)[58]
13 benzene-d625 d69693 (R)[57]
14 toluene−301209898 (R) e[58]
15 dcm-d2255>9594 (R)[57,58]
16 thf-d8255>9595 (R)[57,58]
17 thf-d8012>9596 (R) e[57]
18 thf-d8−30120>9598 (R) e[57]
19 1e3ebenzene-d62548892 (R)[57]
20 benzene2548892 (R)[58]
21 1f3fbenzene-d625309097 (R)[58]
22 2c4cbenzene-d625404831 (R) e[57,58]
23 2d4dbenzene-d625966546 (R)[57,58]
24 3841benzene-d625158966/57 f[58]
25 3942benzene-d625968589 (R)[58]
26 4043benzene251207391 (R)[58]
2730cHfiPr1b3bbenzene-d625209087 (R)[58]
28 1c3cbenzene2559593 (R)[58]
29 1d3dtoluene0159897 (R) e[58]
30 1e3etoluene-d8088595 (R) e[58]
31 1f3fbenzene251209096 (R)[58]
32 2c4cbenzene85308526 (R) e[58]
33 2d4dbenzene85208918 (R)[58]
34 3841benzene25249065/58 g[58]
3531ZrtBu1c3cbenzene25309287 (R)[56]
36 1d3dbenzene25189593 (R)[56]
37 1e3ebenzene25487788 (R)[56]
38 2c4cbenzene25488029 (R) e[58]
a 10 mol-% precatalyst. b dcm = Dichloromethane, thf = tetrahydrofuran. c Enantiomeric excesses (=ee) determined by 1H and/or 19F NMR spectroscopy after reaction with Mosher’s acid chloride. d 2 mol-% precatalyst. e The ee values were determined by HPLC on a chiral stationary phase. f dr = 3:1; (R)-enantiomers. g dr = 2.5:1; (R)-enantiomers. n.d.—not displayed.
Table 11. Catalytic asymmetric hydroamination of 20ah and 45ac using chiral catalysts 17a, 18, 19b, 30b,c and 31 a.
Table 11. Catalytic asymmetric hydroamination of 20ah and 45ac using chiral catalysts 17a, 18, 19b, 30b,c and 31 a.
EntryCat.RR’nSubstr.Prod.t
[h]
Conv.
[%]
dr bee
[%] c
Ref.
117aMeH120a44a2.6 d1001.36:1 e5/21 (R)[79]
218MeH120a44a0.5 f961.55:1 e38/25 (S)[79]
319bMeH120a44a0.251001.2:1 e95/95[56]
4 PhH120b44b0.251001.2:1 e95/96[56]
5 4-C6H4BrH120c44c0.251001.9:1 e95/92[56]
6 ethenylH120d44d0.17100-96 (S)[56]
730bMeH120a44a0.51001.1:193/92[96]
8 48 g1004.2:193/92[96]
9 144 h,i1001:6.596/97[96]
10 PhH120b44b0.51003.3:196/96[96]
11 6 g1008.9:196/95[96]
12 55 j1002:196/96[96]
13 96 g,k951:1.199/99[96]
14 96 h951:4.599/99[96]
15 144 h,i,k1001:699/99[96]
16 4-C6H4BrH120c44c0.51004:197/95[96]
17 3 g1008:195 (cis)[96]
18 ethenylH120d44d0.7598-92 (R)[58]
19 OMeH120e44e4890>20:197 (cis)[96]
20 48 h9010:197 (cis)[96]
21 PhMe120f44f96872:193/95[96]
22 192 l918:192 (cis)[96]
23 PhH220g44g72902.4:133/12[96]
24 96906.6:132/12[96]
25 PhH320h44h96902.8:189/92[96]
26 144 m867:189/91[96]
27 PhMe145a46a0.6100-87[96]
28 2 g100-91[96]
29 PhMe245b46b20100-71[96]
30 48 g100-77[96]
31 PhMe345c46c72 g100-89[96]
3230cMeH120a44a20811.4:1 e87/63[58]
33 4-C6H4BrH120c44c3842:1 e93/96[58]
34 ethenylH120d44d20 n90-96 (R)[58]
3531MeH120a44a30851.2:1 e92/91[56,58]
36 PhH120b44b48932.5:1 e88/92[56,58]
37 4-C6H4BrH120c44c48901.2:1 e96/98[56,58]
38 ethenylH120d44d3090-88 (R)[56,58]
a Conditions: 10 mol-% precatalyst in benzene, ambient temperature; substrate, c = 65.4 mM. b Where not indicated otherwise, dr is given as the ratio of cis:trans. c Enantiomeric excesses (=ee) determined by 1H and/or 19F NMR spectroscopy after reaction with Mosher’s acid chloride. Where the absolute configuration on the 2-position of the formed product was determined, it is given in parenthesis. For products for which the cis/trans configuration was determined, they are ordered as cis/trans. d Conditions: 3 mol-% precatalyst in benzene, 25 °C. e cis/trans Configuration not emphasized for the reaction products. f Conditions: 1.5 mol-% precatalyst in benzene, ambient temperature. g Substrate, c = 5.45 mM. h Substrate, c = 327 mM. i Propyl amine, c = 100 mM. j Amyl amine, c = 29 mM. k T = −30 °C. l Substrate, c = 10.9 mM. m Substrate, c = 16.4 mM. n T = 0 °C.
Table 12. Catalytic asymmetric hydroamination of 1be, 2d and 20a,b using chiral zirconium catalysts 32ag and 33ae a.
Table 12. Catalytic asymmetric hydroamination of 1be, 2d and 20a,b using chiral zirconium catalysts 32ag and 33ae a.
EntryCat.RR’nSubstr.Prod.t
[h]
T
[°C]
Conversion
[%]
ee
[%] b
Ref.
132aEttBu-1d3d4100>9527[98]
232bEt2-Ad-1d3d8.5100>9528[98]
332cEtPh3Si-1d3d4100>9538[98]
432dEtPh3C-1d3d2100>9556[98]
532eMePh3C-1d3d3100>9549[98]
632fBnPh3C-1d3d1.5100>9540[98]
732gtBuPh3C-1d3d2100>951[98]
833aHtBu11d3d2.51009374[98]
933bMetBu11b3b120859689[98]
10 1c3c21859793[98]
11 1d3d4.51159584[98]
12 4.51009489[98]
13 11859292[98]
14 14809093[98]
15 19709594[98]
16 29559394[98]
17 1e3e72859187[98]
18 2d4d2859166[94]
19 20aE/Z-44a898589 c90/93[98]
20 20bE/Z-44b248591 d88/92[98]
2133cEttBu11d3d111009668[98]
2233dPhtBu11d3d59.510085−13[98]
2333eHMe21d3d41009477[98]
a Conditions: 10 mol-% precatalyst in benzene. Conversions measured by 1H NMR spectroscopy. b Products were converted to N-Ts (Ts = Tosyl) compounds and the enantiomeric excess (=ee) was determined by chiral HPLC analysis; a positive value refers to the (R)-enantiomer; the dr value was determined by the analysis of the 1H NMR spectrum of the crude product. c dr = 1:1.2. d dr = 1:1.7.
Table 13. Catalytic asymmetric hydroamination of 1b-d, 20a and 51 using chiral lithium catalysts 47 and 48 a.
Table 13. Catalytic asymmetric hydroamination of 1b-d, 20a and 51 using chiral lithium catalysts 47 and 48 a.
EntryCat.Substr.Prod.[cat.]
[mol-%] b
t a
[h]
T
[°C]
Conv.
[%]
ee
[%] c
Ref.
1471b3b7.59229664 (S)[110]
2 542229668 (S)[110]
3 2.545229367 (S)[110]
4 5 d407806653 (S)[110]
5 1c3c2.51.1229175 (S)[110]
6 5 e2209874 (S)[110]
7 5 d91606469 (S)[110]
8 1d3d50.8229731 (S)[110]
9 5 d27807024 (S)[110]
10 515250.08229817[110]
11 20a44a52229864/72 f[110]
12481b3b10333120562[110]
13 1d3d101401200-[110]
a Conditions: benzene-d6, Ar atm.; conversions measured by 1H NMR spectroscopy. b Calculated for dimeric species (S,S,S)-47 and (S,S,S)-47·4thf containing four Li atoms each. c Enantiomeric excess (=ee) was determined by 19F NMR spectroscopy after derivatization with Mosher’s acid chloride. d (S,S,S)-47·4thf used as catalyst. e Reaction in toluene. f dr(E-44a:Z-44a) = 1.2:1.
Table 14. Catalytic asymmetric hydroamination of 53af using chiral lithium catalysts 50a–f a.
Table 14. Catalytic asymmetric hydroamination of 53af using chiral lithium catalysts 50a–f a.
Molecules 28 02702 i008
EntryCat.R*RR’R’’nSubstr.Prod.t
[h]
Yield
[%] b
ee
[%]
Ref.
150aiPrMePhMe253a54a599(0)71 (S)[113]
2 PhMe153b54b599(0)84 (S)[113]
3 Ph4-C6H4OMe253c54c31 c94(0)11 d[114]
4 Ph2-propenyl153d54d2 e98(0)83 (S)[114]
5 Ph2-propenyl153e f54e f21 e90(0)18 d[114]
6 HMe153f54f1 g33(7)43 d[114]
750btBuMePhMe253a54a5 h99(0)31 (S)[113]
8 PhMe153b54b589(0)19 (S)[113]
950ciPrEtPhMe253a54a2799(0)62 (S)[113]
10 PhMe153b54b599(0)84 (S)[113]
1150diPrCH2MePhMe253a54a597(0)84 (S)[113]
12 PhMe153b54b599(0)79 (S)[113]
1350esecBuMePhMe253a54a525(0)81 (S)[113]
14 PhMe153b54b599(0)66 (S)[113]
1550f-MePhMe253a54a2254(0)62 (R)[113]
16 PhMe153b54b598(0)86 (R)[113]
a The reaction was conducted with 0.4 equiv of ligand, 0.2–0.4 equiv of butyllithium at −60 °C. b Yield of 55 in parentheses. c Carried out at 0 °C. d Absolute configuration not determined. e Carried out at −40 °C. f -(CH2)2- instead of aromatic backbone C6H4 in substrate 53e. g Carried out at −20 °C. h Carried out at ambient temperature.
Table 15. Catalytic asymmetric intermolecular hydroamination of vinylarenes 58a,b with amines 59a,b using the chiral lithium catalytic system 49a,b a.
Table 15. Catalytic asymmetric intermolecular hydroamination of vinylarenes 58a,b with amines 59a,b using the chiral lithium catalytic system 49a,b a.
Molecules 28 02702 i009
EntryCat.AdditiveVinylareneAmineProd.t a
[h]
Yield
[%] b
ee
[%]
Ref.
149a(−)-sparteine58a59a60a3757-[109]
2 58a59b60b65407[109]
3 58b59a60c13.571-[109]
4 58b59b60d186014[109]
549bα-isosparteine58a58b60b7038-[109]
a Reaction conditions: 2 mmol of vinylarene, 2 mmol of amine, 2 mol-% of LiN(TMS)2, 2 mol-% of additive, and 0.1 mL of toluene at 120 °C. b Isolated yield by column chromatography.
Table 16. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1bd using the chiral alkaline-earth-metal-based catalysts (S,S,S)-61, (R,S,S)-61 and 62a,b a.
Table 16. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1bd using the chiral alkaline-earth-metal-based catalysts (S,S,S)-61, (R,S,S)-61 and 62a,b a.
EntryCat.Substr.Prod.[cat]
[mol-%]
T
[°C]
t a
[h]
Conv.
[%]
ee
[%] b
Ref.
1(S,S,S)-611b3b510022≥994 (S)[116]
2 1c3c52222940[116]
3 1d3d4220.33≥996 (R)[116]
4(R,S,S)-611b3b1010021≥990[116]
5 1c3c10223.5806 (R[116]
6 1d3d10220.17≥9914 (R)[116]
762a1b3b10r.t.1680-[120]
8 10801208027 (R)[120]
9 1c3c10r.t.240-[120]
10 1060269336 (R)[120]
11 1d3d10r.t.24890[120]
12 106012≥990[120]
1362b1b3b10r.t.0.0810018 (S)[120]
14 1c3c10r.t.0.08≥9918 (S)[120]
15 180168≤10n.d.[120]
16 1d3d10r.t.0.08≥990[120]
a Reaction conditions: benzene-d6, Ar atm. b Enantiomeric exess (=ee) determined by 1H and/or 19F NMR spectroscopy of amide after derivatization with Mosher’s acid chloride. n.d.—not displayed.
Table 17. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1b,d and 2d using the chiral calcium catalysts 64ad, 65ac and 66a,b a.
Table 17. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1b,d and 2d using the chiral calcium catalysts 64ad, 65ac and 66a,b a.
EntryCat.R’Substr.Prod.T
[°C]
t
[h]
Conv.
[%] b
ee
[%] c
Ref.
164atBu1b3br.t.168900[26]
2 1d3dr.t.24≥996[26]
364biPr1b3br.t.120≥9912[26]
4 1d3dr.t.1≥995[26]
564cPh1b3br.t.5040-[26]
6 1d3dr.t.728026[26]
764d4-C6H4F1b3br.t.3360-[26]
8 1d3dr.t.3360-[26]
965a4-C6H4Me1b3br.t.n/a d85[122]
10 1d3dr.t.n/a d950[122]
1165b4-C6H4F1b3br.t.n/a d39[122]
12 1d3dr.t.n/a d>999[122]
1365ctBu1b3br.t.n/a d1812[122]
14 1d3dr.t.n/a d≥9912[122]
1566a-1d3d30245114[123]
16 4024≥9922[123]
17 50728224[123]
18 2d4d80120140[123]
19 50 e24268[123]
20 80 e120836[123]
2166b-1d3d2124≥9925[123]
22 3024≥9926[123]
23 40248820[123]
24 2d4d80120trace-[123]
25 50240-[123]
26 801201416[123]
a Reaction conditions: 10 mol-% catalyst, benzene-d6, Ar atm. b Determined from 1H NMR spectroscopy. c Enantiomeric excess (=ee) determined by 1H NMR spectroscopy using (R)-(-)-(O)-acetylmandelic acid to allow for the distinction between the two enantiomers. No absolute configuration was determined. d Initial rates were determined instead. e 20 mol-% catalyst. n/a—not applicable.
Table 18. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1bd using the chiral calcium catalysts 68 and (S)-69 a.
Table 18. Catalytic asymmetric intermolecular hydroamination of aminoalkenes 1bd using the chiral calcium catalysts 68 and (S)-69 a.
Molecules 28 02702 i010
EntryCat.Substr.Prod.T
[°C]
t
[h]
Conv.
[%] b
ee
[%] c
Ref.
1681b3b9018>988[27]
2 1c3c605>988[27]
3 1d3d205>9915[27]
4 601>9915[27]
5(S)-691b3b9024>9726[27]
6 1c3c605>9823[27]
7 1d3d205>9933[27]
8 601>9933[27]
a Reaction conditions: 10 mol-% catalyst, benzene-d6, Ar atm. b Determined from 1H NMR spectroscopy. c Enantiomeric excess (=ee) determined by 1H and/or 19F NMR spectroscopy after derivatization with Mosher’s acid. The absolute configuration was not determined for the reaction products.
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

Notz, S.; Scharf, S.; Lang, H. Jumping in the Chiral Pool: Asymmetric Hydroaminations with Early Metals. Molecules 2023, 28, 2702. https://doi.org/10.3390/molecules28062702

AMA Style

Notz S, Scharf S, Lang H. Jumping in the Chiral Pool: Asymmetric Hydroaminations with Early Metals. Molecules. 2023; 28(6):2702. https://doi.org/10.3390/molecules28062702

Chicago/Turabian Style

Notz, Sebastian, Sebastian Scharf, and Heinrich Lang. 2023. "Jumping in the Chiral Pool: Asymmetric Hydroaminations with Early Metals" Molecules 28, no. 6: 2702. https://doi.org/10.3390/molecules28062702

Article Metrics

Back to TopTop