Next Article in Journal
Nylons with Applications in Energy Generators, 3D Printing and Biomedicine
Previous Article in Journal
Carbon Dots Derived from Non-Biomass Waste: Methods, Applications, and Future Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies for Accessing cis-1-Amino-2-Indanol

Aix Marseille Univ, CNRS, ICR, Marseille, France
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(11), 2442; https://doi.org/10.3390/molecules29112442
Submission received: 26 April 2024 / Revised: 14 May 2024 / Accepted: 15 May 2024 / Published: 22 May 2024
(This article belongs to the Section Organic Chemistry)

Abstract

:
cis-1-amino-2-indanol is an important building block in many areas of chemistry. Indeed, this molecule is currently used as skeleton in many ligands (BOX, PyBOX…), catalysts and chiral auxiliaries. Moreover, it has been incorporated in numerous bioactive structures. The major issues during its synthesis are the control of cis-selectivity, for which various strategies have been devised, and the enantioselectivity of the reaction. This review highlights the various methodologies implemented over the last few decades to access cis-1-amino-2-indanol in racemic and enantioselective manners. In addition, the various substitution patterns on the aromatic ring and their preparations are listed.

1. Introduction

1-Amino-2-alcohols are crucial building blocks in many fields, including medicinal and organic chemistry. Consequently, both their synthesis and the development of new methodologies to obtain them constitute an important area of organic chemistry [1,2,3,4]. More particularly, cis-1-amino-2-indanol (1) plays a central role in organic synthesis as a ligand or chiral auxiliary due to its rigid cyclic skeleton (Scheme 1). This structure is the key moiety of BOX and PyBOX ligands widely used in asymmetric catalysis [5,6,7]. Notably, oxazaborilidine catalysts derived from cis-1-amino-2-indanol are often more efficient than other chiral 1-amino-2-alcohol structures in the enantioselective reduction in carbonyl compounds [8]. In addition, cis-1-amino-2-indanol is an important derivative when used as a chiral auxiliary in several asymmetric transformations, such as diastereoselective enolate alkylation or diastereoselective reduction [9]. This compound, like other chiral amines, also has applications in the resolution of racemic carboxylic acid bearing a chiral carbon at position α [9]. It is also an interesting substructure for drug design. It is present in Indinavir sulfate (Crixivan®), an HIV protease inhibitor for the treatment of acquired immunodeficiency syndrome (AIDS) developed by Merck (Rahway, NJ, USA) [10,11], or in KNI-10006 for anti-malarial treatment [12].
Due to its wide range of applications, the synthesis of cis-1-amino-2-indanol has attracted the interest of numerous academic and industrial researchers, leading to the development of various synthetic pathways (Scheme 2) [9,13]. Most of these strategies involve indane skeleton as the starting point, but total syntheses from other building blocks have also been reported. Whatever the starting material, the same synthetic difficulties have to be overcome. One of the first points to control is the selective introduction of nitrogen and oxygen atoms in the -C1 and C2 positions, respectively. Afterward, the main challenge is to control the cis relationship between the oxygen and nitrogen atoms and to carry out the process enantioselectively.
From the indane skeleton, the main strategy for obtaining the cis relationship between nitrogen and oxygen atoms involves the key intramolecular formation of a cis-5-membered ring (Scheme 2, (A)). This approach requires a final hydrolysis step to deliver the desired cis-1-amino-2-indanol. Alternative methodologies have been developed relying on the epimerization of C1 or C2 centers, or diastereoselective reduction (Scheme 2, (B) and (C)). Interestingly, other syntheses have been developed from non-indanic precursors, allowing the use of the chiral pool to introduce the right configurations (Scheme 2, (D)).

2. Syntheses from Indane Skeleton

2.1. cis Stereochemistry Controlled by Intramolecular Formation of 5-Membered Ring

2.1.1. Nitrilium Trapping—Ritter Type Reaction

Ritter type reaction on indene oxide is a practical route for the synthesis of enantiopure cis-1-amino-2-indanol. This strategy was reported by Senanayake in 1995, first starting from indene oxide (2) prepared via the reaction of indene with hydrogen peroxide. This allowed the isolation of cis-1-amino-2-indanol (−/+)-1 under acid conditions after hydrolysis of oxazoline intermediate 3 (Scheme 3) [14]. The first attempts using 2.0 equivalent of 97% sulfuric acid in acetonitrile from −40 °C to RT gave the desired compound with 55–60% yield. Interestingly, when the reaction was performed with fuming sulfuric acid (21% SO3), the formation of indanone 4 was suppressed, and the yield improved to 78–80%.
This strategy was then optimized, and a larger scale synthesis was proposed, enabling the attainment of 17.1 g of enantiopure (1S,2R)-1 starting from enantioenriched indene oxide (1S,2R)-2 obtained via Jacobsen enantioselective epoxidation of indene (5) (Scheme 4) [15]. After applying typical sequence Ritter reaction and hydrolysis, the resulting sulfuric acid salt (1S,2R)-1 was engaged in an enantioenrichment process by using L-tartaric acid, giving the desired compound (1S,2R)-1 with a high enantiomeric excess (>99%).
From a mechanistic point of view, the authors proposed that in the presence of H2SO4 alone, the epoxide 2 is in equilibrium with the opened form 6 (Scheme 5). This carbocation could be reversibly trapped by the counter anion (HSO4), giving 7, which was observed via NMR analysis, or evolve toward the formation of the undesired ketone 4 via 1,2 hydride shift. However, the major pathway consists of its trapping by acetonitrile, leading to the corresponding nitrilium ions cis-8 and trans-8 as a mixture of diastereomers. The success of this reaction is due to the equilibrium between the cis- and trans-nitrilium 8, which is displaced by the fast cyclization of the cis-nitrilium, leading to 3. In addition, this mechanism explains the transfer of chiral information from the C2 carbon, since no epimerization via the formation of a carbocation occurs at this position. When the reaction is performed in the presence of SO3, epoxide 2 is proposed to evolve toward the formation of cyclic sulfate 9 observed at −40 °C via NMR. This intermediate is in equilibrium with the corresponding carbocation 10 through regioselective ring-opening. Suppression of the indanone by-product is likely to result from this fast equilibrium, which probably enables the 1,2 hydride shift process. The trapping of 10 by acetonitrile renders a mixture of cis and trans-nitrilium 11, which cyclizes while releasing SO3 to finally deliver the desired cis-1-amino-2-indanol with 78–80% yield after hydrolysis.
Interestingly, and in accordance with the mechanism, stereochemistry is controlled by the configuration of the C2 center; consequently, the reaction works with enantiopure diol 12 as the starting material (Scheme 6) [14,16]. Indeed, when the reaction was performed with cis-1,2 diol or trans-1,2 diol, the product was obtained with high yield and without loss of chiral information. It is worthwhile noting that a better yield was obtained when cis-diol was used as the substrate instead of epoxide in the presence of H2SO4 (81% (Scheme 6) vs. 55–60% (Scheme 3)).
To date, numerous methodologies have been developed to obtain enantioenriched epoxide, such as Jacobsen type reactions [17,18], biocatalytic/biomimetic epoxidations [19,20,21] or direct racemic epoxide resolution [22,23]. In addition, epoxide could be obtained from common enantioenriched precursors (i.e., β-hydroxy selenides [24], indene bromohydrin [25]). The enantioselective synthesis of diols 12 is also described in the literature [26,27,28,29,30,31]. It should be underlined that these methodologies—some of which were developed subsequently to the strategies presented here—would most likely be very useful in the preparation of enantiopure cis-1-amino-2-indanol.
More recently, Lambert’s group described an electrophotocatalytic amino-oxygenation reaction of aryl olefins (Scheme 7) [32]. The reaction used a trisaminocyclopropenium (TAC) ion catalyst under electrochemical conditions and compact fluorescent lamp (CFL) irradiation in the presence of acetonitrile and 5.0 equivalent of water. This enabled the formation of oxazolines 3 and 14 with 55% and 43% yield, respectively, from the corresponding indene derivatives.
Regarding the mechanism (Scheme 8), the authors first proposed an electrochemical oxidation of the photocatalyst 15 to generate the corresponding radical dication 16. Its photoexcitation renders 17, which shows high oxidizing properties (E*red = 3.3 V vs. SCE) and can thus oxidize indene to furnish the corresponding radical cation 18. The regioselective trapping of this intermediate by water at C2 position leads to benzylic radical intermediate 19, which in turn could be oxidized by 16 or directly by the anode. This reaction furnishes the benzylic carbocation 6, which is subsequently reversibly trapped by an equivalent of acetonitrile. The highly favored cis-cyclization delivers the final oxazoline 3. The authors stressed the importance of controlling the amount of water used in the reaction (5.00 equiv.), as a large excess could lead to the trapping of highly electrophilic intermediates generated during the process. Indeed, during the optimization course, when water was used in excess (50.00 equiv.), the authors isolated an aldehyde derivative resulting from probable oxidative cleavage of 1,2 diol intermediates.

2.1.2. Intramolecular Amide Cyclization

Intramolecular amide cyclization is an important class of reaction for the synthesis of cis-1-amino-2-indanol. The key point of this strategy is to form an amide/urethane derivative at C1 position and a leaving group at C2 position, which is engaged in a O-amide cyclization step to give the corresponding cis-5-membered ring. Two main approaches have been developed to access the key intermediate via either the N-opening of epoxide or direct electrophilic activation of indene.
This strategy was first reported in 1951 by Lutz and Wayland based on the key formation of a cis-oxazoline obtained from the corresponding trans-1-amino-2-indanol (−/+)-1′ (Scheme 9) [33]. The latter was obtained in two steps via addition of ammonia to 2-bromo-1-indanol (20), which probably proceeded through an epoxide intermediate to form 1, followed by quantitative amidation reaction rendering 21. The addition of thionyl chloride allowed an intramolecular cyclization with inversion of the configuration at C2, leading to the cis-oxazoline 22. Finally, acidic hydrolysis at reflux delivered the cis-1-amino-2-indanol (−/+)-1 with 68% yield.
Based on this strategy, researchers from Sepracor developed an enantioselective synthesis of (1S,2R)-1 using an enantioselective epoxidation reaction as the key step (Scheme 10) [34]. Indeed, (1R,2S)-indene oxide (2) was prepared in 80–85% ee in the presence of NaOCl and (R,R)-Mn-Salen. Then, the epoxide was opened with ammonia, giving trans-1-amino-2-indanol (1′), which subsequently engaged in the amidation reaction under Schotten–Baumann conditions. Recrystallization from a mixture of DMF/MeOH delivered benzamide 23 with 47% yield in three steps from indene 5 and 99% enantiomeric excess. The addition of 80% H2SO4 at 80 °C to 23 led to the formation of oxazoline 24, which was directly hydrolyzed, delivering the enantiopure (1S,2R)-1 with 84% yield and an enantiomeric excess exceeding 99%.
In 1997, Ghosh et al. reported the synthesis of both enantiomers of cis-1-amino-2-indanol (Scheme 11) [35]. In this approach, racemic trans-1-azido-2-indanol (25) was obtained in two steps from indene via epoxidation, followed by opening with sodium azide. Enzymatic acylation proceeded in the presence of lipase PS 30 in a mixture of dimethoxyethane (DME) and isopropenyl acetate to give the unreactive alcohol (1S,2S)-25 (46% yield, ee > 96%) and acylated enantiomer (1R,2R)-26 (44% yield, ee > 96%). Once isolated, both enantiomers were then engaged in the same reaction sequence. The azido group was reduced via hydrogenation and in situ converted into carbamates 27 in the presence of diethylpyrocarbonate. Key cyclization with epimerization of the C2 center occurred in SOCl2 giving 28 with excellent yields, and a final basic hydrolysis delivered the enantiopure products.
Didier proposed an alternative route for both enantiomers of 1 from β–ketoester 29 (Scheme 12) [36]. Baker’s yeast reduction of 29 yielded the enantiopure 2-hydroxy ester 30. The hydrolysis of 30 in the presence of NaOH led to a partial racemization at the C1 position, providing trans-isomer 31 after a selective crystallization. C1 epimerization was overcome by enzymatic hydrolysis, leading only to the cis isomer 31. The formation of the C–N bond at the C1 position occurred via a Curtius rearrangement triggered by diphenyl phosphorazidate (DPPA), rendering oxazolidinone 28 and carbamate 27 from cis and trans carboxylic acids, respectively. Basic hydrolysis of 28 led to (1R,2S)-1, whereas its enantiomer (1S,2R)-1 was obtained after inversion of the configuration at the C2 position thanks to an intramolecular cyclization of 27 followed by hydrolysis.
In 1967, Heatchcok et al. proposed a more direct approach through iodoamidation of indene, leading to a trans iodocarbamate 32, which, under reflux in diglyme, yielded the cis-oxazolidinone 28 with 88% yield. A final hydrolysis under basic conditions in methanol at reflux afforded the desired racemic compound (−/+)-1 with 79% yield (Scheme 13) [37].
In 1989, Ogura and coworkers developed a strategy for accessing oxazolidinone from alkene involving organotelluric compounds under Lewis acid catalysis (Scheme 14) [38,39]. Under the best conditions, indene (5) was converted into oxazolidinone 28 with 79% yield in the presence of benzenetellurinyl trifluoroacetate, ethyl carbamate and boron trifluoride etherate under reflux of dichloroethane. In the mechanism proposed by the authors, the activation of the double bond enables its amidotellurinylation via intermediate 33. Under reflux of dichloroethane and assisted by BF3, cyclization of 34 with the inversion at C2 center occurred, leading to the corresponding oxazolidinone 28.
Recently, there has been renewed interest in the amino-oxygenation of alkenes. In 2015, Tepe’s group developed an approach with Br-N-(CO2Me)2 as the reagent (Scheme 15) [40]. More particularly, indene 5 was transformed into the oxazolidinone 35 precursor of cis-1-amino-2-indanol with 48% yield. The authors proposed the conditions for hydrolysis for oxazolidinones (2M LiOH in THF); however, such conditions were not directly applied to 35. Li and coworkers reported a similar strategy using electrophilic iodine and urea as a partner, rendering N-substituted oxazoline 36 [41]. Later, the use of benzamide as a partner was developed to isolate oxazoline 24 with 62% yield [42]. Finally, Hashimoto’s group identified a new carbamate as a bifunctional N- and O-nucleophile, giving—in the presence of chiral hypervalent iodine—the corresponding oxazolidinone 37 with 75% yield and 50% ee [43].
From a mechanistic point of view, this is an electrophilic activation of olefin by bromine—the iodine of hypervalent iodine—leading to intermediates 38 or 38′ (Scheme 16). Then, the N-nucleophile reacts with the electrophilic benzylic position of 38 or 38′, leading to the formation of trans amido product 39 or 39′, followed by an intramolecular cyclization with the nucleophilic oxygen, giving the desired compound 40.

2.1.3. Benzylic Csp3-H Amination

  • Benzylic Csp3-H amination via radical pathway
The use of trichloroacetamidates for radical β-amination reactions for the synthesis of cis-1-amino-2-indanol has been proposed by Nagib in 2017 (Scheme 17) [44]. Indeed, when compound 41 was engaged in the reaction in the presence of PhI(OAc)2 and NaI under light irradiation in acetonitrile, the desired compound was obtained with a high diastereomeric ratio (>20:1) and 81% yield after hydrolysis of the corresponding oxazoline 45. The reaction consisted of the formation of an imidate radical 42, which underwent 1,5-hydrogen atom transfer (HAT) to generate the corresponding benzylic radical 43. The trapping of the carbon-centered radical by iodine led to 44, followed by intramolecular cyclization. The resulting oxazoline 45 ring was hydrolyzed under acidic conditions, delivering cis-1-amino-2-indanol. A similar approach was developed in the meantime by He [45] using three equivalents of NIS in DCE under thermal conditions, and an approach using a catalytic amount of iodine was devised in 2019 by Nagib [46]. Finally, Shi has proposed a new way of obtaining the imidate radical via direct abstraction of the hydrogen of the N–H bond hydrogen atom bond by the malonyl peroxide (MPO) radical [47].
In a very well-designed study on the enantioselective synthesis of β-amino alcohols via radical amination, Nagib’s group used the achiral 2-indanol imidate derivative 46 as a substrate for mechanistic studies (Scheme 18) [48]. When placed under their best conditions, the corresponding oxazoline 49 was obtained with 67% ee and a diastereomeric ratio exceeding 20:1. However, the compound was not isolated, and low yield was obtained (12% NMR yield). The key point in the process is the enantioselective HAT via CuL*-bound imidate 47, giving the corresponding chiral benzylic radical 48. It is important to note that the behavior of this substrate seemed to be different to others described in the publication, since better results were obtained. However, this result paved the way for further developments.
Recently, He and coworkers have shown that binap-containing copper photocatalysts on pillar-layered MOF supports were interesting and stable photocatalysts for the conversion of N-acyloxy imidates 50 into oxazoline 24 (Scheme 19) [49]. The oxazoline 24 precursor of cis-aminoindanol was obtained with 42% yield in the presence of a catalytic amount of the Cu photocatalyst and DABCO in N,N-dimethylacetamide (DMA) under blue LED irradiation. The reaction starts with the reduction of the N-acyloxy imidate 50 by the excited state of the photocatalyst, leading to the formation of a N-centered radical 51 and the oxidized photocatalyst (Scheme 19). In the following step, a 1,5 HAT occurs to furnish the benzylic radical 52, which reacts via single electron transfer (SET) with the oxidized photocatalyst to regenerate it at its ground state. On the other hand, this reaction enables the formation of the corresponding benzylic carbenium ion 53, which is trapped via intramolecular N-cyclization assisted by the base (DABCO).
Carbamates 54 derived from 2-indanol are interesting substrates for the synthesis of cis-aminoindanol via a radical pathway. Nicholas’ group developed a copper-diimine catalyzed oxidative C-H insertion of carbamates (Scheme 20) [50]. The reaction is proposed to work through the generation of a L-Cu-imido derivative behaving as a triplet-state diradical specie, in which the C-H amination is operated via a stepwise radical process [51]. The authors developed an enantioselective version of this reaction, but moderate enantiomeric excesses were obtained (13–18% ee). The resulting carbamate can be hydrolyzed under several developed conditions [52].
  • Benzylic Csp3-H amination via nitrene chemistry
2-Indanol 55 is also an attractive substrate for creating the C–N benzylic bond. The amination of the C–H bond occurs from the corresponding primary carbamate 54 [53,54], sulfamate ester 56a [54,55,56,57,58,59,60], azidoformate 56b [61], alkylsulfonyloxycarbamate 56c [62,63,64,65] or carbamimidate 56d [66] in the presence of various catalysts, such as the complexes of rhodium [53,55,62,63,64,65,66], ruthenium [56,67], silver [54], cobalt [61], manganese [59,60] or iron (Scheme 21) [57,58,59]. The cis insertion of the nitrogen atom into the C–H benzylic bond enables the control of stereoselectivity of the reaction thanks to the involvement of a metal-nitrene intermediate.
The first attempts to desymmetrize carbamate 54 or sulfamate 56a are reported in the literature. The use of chiral ligands with Rh(II) [68,69], Ru(II) [67] or Mn(III) [60] led to low ee, ranging from 43 to 75%. Recently, Meggers’ group has developed an efficient general approach to chiral oxazolidinone via C(sp3)-H nitrene insertion catalyzed with the chiral complexes of Ru(II) associated with N-(2-pyridyl)-substituted N-heterocyclic carbine (NHC) ligands [70]. In 1,2-dichlorobenzene at 30 °C, the precursor 58 delivered oxazolidinone (3S,8R)-28 with 99% yield and 95% ee (Scheme 22).
In all these strategies, the protected amino alcohol must be released from the tricyclic compounds. Hydrolysis of oxazolidinone 28 under the basic condition, typically KOH, led to 1. For instance, this approach has been widely used in the synthesis of BOX ligands [71,72,73] or biologically active molecules [35]. Recently, new conditions for the hydrolysis of cyclic carbamate in the presence of diethylenetriamine were developed and resulted in 1 with 86% yield [52]. Surprisingly, the conversion of 57a into (−/+)-1 was not described in the literature. However, it is well known that the reduction of cyclic sulfatame with LiAlH4 yields the corresponding amino alcohol with retention of the configuration [74,75], unlike hydrolysis, which would lead to an inversion of the configuration of the C–O bond and thus to the trans amino-indanol [76]. No example of hydrolysis of a carbamimidate 57b leading to 1 has yet been reported. However, analogous structures were cleaved under acidic conditions (H2SO4), releasing the amino-alcohol moiety with good yield [77,78].

2.2. cis Stereochemistry Controlled by Epimerization

2.2.1. Epimerization via SN2 at C1 Position

Resnick’s group proposed a straightforward approach to the synthesis of 1 [79] from (1S,2R)-indanediol (12, ee 92%) prepared via dioxygenase hydroxylation of 2-indanol 55 (Scheme 23) [80]. The carbon–nitrogen bond in 26 was formed through a two-step sequence, with an overall yield of 69%: (i) concomitant introduction of chlorine in the benzylic position and the acetylation of alcohol in position 2, delivering 59; and (ii) chlorine substitution by an azide. The aminolysis of 26 followed by the hydrogenation of azide 25 led to (1S,2R)-1.

2.2.2. Epimerization via Mitsunobu Reaction at C2 Position

In 1995, Ogasawara and Takahashi reported the synthesis of both enantiomers of cis-1-amino-2-indanol via the resolution of trans-1-azido-2-indanol (25) (Scheme 24) [81]. The racemic epoxide 2 was obtained in a two-step sequence via treatment of indene 5 with NBS to render trans-bromohydrin with 82% yield, followed by an epoxidation step mediated by NaOH. The opening of the epoxide by sodium azide rendered the racemic trans-1-azido-2-indanol (25) with 93% yield. A screening of the conditions allowed the authors to identify Lipase PS (Pseudomonas sp. Amano) and vinyl acetate in tert-butyl methyl ether as the best conditions for alcohol resolution. Unreacted trans-(1S,2S)-azido-2-indanol (25) was recovered with 48% yield and 99% ee, whereas azido acetate (1R,2R)-26 was obtained with 49% yield and 98% ee. The latter was then converted into the corresponding alcohol via methanolysis with 92% yield without the erosion of chirality. On the other hand, Mitsunobu inversion of 25 was carried out in the presence of p-nitrobenzoic acid, diethyl azodicarboxylate (DEAD) and triphenylphosphine, rendering the corresponding cis-azido ester 60 with 75% yield. At this stage, two pathways were proposed to reach the enantiopure cis-1-amino-2-indanol (1S,2R)-1, either via concomitant reduction of the azido group and the ester with LiAlH4 in THF (65% yield) or through a two-step sequence consisting of methanolysis of the ester followed by hydrogenation over palladium on carbon (98% yield). The same strategy was successfully applied to transform the other trans-azido-alcohol enantiomer into (1R,2S)-1.

2.3. cis Stereochemistry Controlled by Diastereoselective Imino Alcohol Reduction

In the 1-amino-2-indanol 1 route, 2-hydroxy-1-indanone (61) turned out to be an interesting intermediate (Scheme 25). Different strategies were proposed for its preparation in its enantiomerically pure form. Using the chiral pool, (R)-phenylalanine (64) was converted in five steps to (R)-61 [82]. Alternatively, (R)-61 was also synthesized in two steps from 1-indanone (63) thanks to an α acetoxylation of the ketone, followed by enzymatic resolution [82,83]. Desymmetrization of protected 2-indanol 65 through the oxidation of the benzylic position with Mn(salen) complexes also yielded (R)-61 with a modest yield (13%, 70% ee) [84]. The conversion of (R)-2 into (1S, 2R)-1 was performed in two steps: (i) formation of the oxime; and (ii) diastereoselective reduction with H2/Pd [82,84] or in the presence of oxazaborolidine complexes prepared from different chiral amino alcohols and BH3 [83].
Direct reduction of 1,2-indanedion-1-oxime (66)—prepared from 2-indanone (4) via oximination—with H2 in the presence of Pd/BaSO4 led to the corresponding racemic cis-N-hydroxyamine 67 (Scheme 26A) [85]. Subsequent hydrogenation of the N-hydroxyamine moiety enabled (−/+)-1.
An analogous enantioselective approach was developed with sequential reductions (Scheme 26B). The first reduction with a biocatalyst—Daucus carota—led to alcohol 62, and then, the oxime function was hydrogenated to an amino group with H2 in the presence of Pd/C with 99% ee and 95% conversion in two steps [86].

3. Syntheses from Non-Indane Skeleton

3.1. Synthesis from (E)-Cinnamate Ethyl Ester

In 2006, Ko’s group reported an eight-step enantioselective synthesis of cis-1-amino-2-indanol from (E)-cinnamate ethyl ester (68) (Scheme 27) [87]. The synthesis started with a Sharpless asymmetric dihydroxylation, leading to the corresponding syn-diol 69 with 97% yield and 99% ee. The benzylic alcohol was then selectively substituted with inversion of the configuration under Mitsunobu conditions using HN3 as a nucleophile delivering 70 [88]. After reduction of the azide 70 under a hydrogen atmosphere, the corresponding amino alcohol 71 was judiciously protected with 79% yield under oxazolidinone form 72 in the presence of triphosgene, allowing the protection and further deprotection of both alcohol and amine in a single step. The ethyl ester 72 is then engaged in a saponification reaction, in which the co-solvent (Et2O) plays an important role. Indeed, the choice of a more polar solvent, such as THF, leads to detectable epimerization via 1H NMR. The indane ring skeleton was introduced with 83% yield via a Friedel–Crafts acylation, first converting the carboxylic acid 73 into its acyl chloride derivative and then via addition of an excess of AlCl3. The corresponding indanone 74 was then reduced by an excess of silane catalyzed by BF3Et2O under microwave irradiation, since low conversions were observed under thermal conditions (oil bath). The cis-1-amino-2-indanol was obtained after final deprotection of the oxazolidinone ring 28 under basic conditions.

3.2. Synthesis from 7,3-Xylofuranose Derivative

In 2021, a successful enantioselective synthesis of cis-1-amino-2-indanol using the Diels–Alder reaction as a key step was reported (Scheme 28) [89]. The chiral pool employed for the synthesis was the versatile chiron 7,3-xylofuranose derivative (7,3-LXF) prepared in two steps from diacetone-D-glucose [90]. 7,3-LXF was first transformed into diacetylated derivative 76 under acidic conditions. Then, diastereomeric allylation was achieved with 85% yield in the presence of allylsilane and BF3Et2O, followed by Pd-catalyzed β-elimination, rendering 78 with 60% yield. Acidic deprotection allowed the formation of diol 79, followed by a selective tosylation reaction, enabling the introduction—in the second sequence—of the amine group with an appropriate configuration. In order to avoid the formation of by-products during the Diels–Alder step, the amine 81 was first protected to render 82 and then refluxed in toluene at 150 °C, rendering—after CO2 extrusion from 83—compound 84. Rearomatization via DDQ led to 85 with 81% yield, and the final removal of the Boc group under acidic conditions afforded the desired cis-1-amino-2-indanol with 80% yield.

3.3. Synthesis from D-Phenylalanine

α-amino acids such as phenylalanine provide a readily available chiral pool for synthesis. Hiyama and coworkers reported a synthesis of enantiomerically pure (1S,2R)-1-amino-2-indanol (1), starting from phenylalanine (Scheme 29) [91]. The reaction began by the conversion of D-phenylalanine (64) into the corresponding optically pure hydroxylated compound 86 with 82% yield using a mixture of NaNO2–H2SO4 [92]. The latter was then protected and the carboxylic acid transformed in acyl chloride with thionyl chloride at 50 °C. The addition of AlCl3 to 87 led to the cyclic structure 88 via Friedel–Crafts acylation without the loss of chiral information. A screening of several conditions for the hydrolysis of 88 without epimerization enabled the identification of the best conditions, i.e., Sc(OTf)3 (20 mol%) in a mixture of H2O–MeOH (1:4) at room temperature, rendering 61 with 82% yield and >99% ee. The α-hydroxy ketone 61 was transformed into its oxime equivalent 62 obtained as a mixture of isomers. After optimization, the authors found out that diastereoselective hydrogenation using Pd black in MeOH/HBr afforded the final product (1S,2R)-1 with 66% yield.

4. Resolution

As previously mentioned, enantiopure cis-1-amino-2-indanol can be prepared from an enantiopure starting material, such as an epoxide, a diol or an azido-alcohol. However, a significant number of syntheses provide a racemic form of cis-1-amino-2-indanol. Various approaches have been proposed for resolving this racemic mixture.

4.1. Chemical Resolution

Interestingly, the resolution of racemic cis-1-amino-2-indanol was performed after its derivatization and chromatographic separation (Scheme 30) [93]. After the preparation of racemic cis-1-amino-2-indanol, following the Lutz and Wayland strategy, the authors performed peptide coupling with Boc-Phe-OH followed by Boc deprotection, leading to the formation of a pair of diastereomers 89 and 90 separable via chromatography on silica gel. Both diastereomers were isolated, each with 40% yield, and enantioenriched cis-1-amino-2-indanol was isolated with 93% yield after cleavage of the amide bond.
Another general strategy for β-amino-alcohol resolution is to perform a kinetic resolution via enantioselective acylation of the alcohol in the presence of a chiral nucleophilic catalyst. Kawabata’s group implemented this approach in cis-1-amino-2-indanol with a chiral aminopyridine as a catalyst (Scheme 31) [94]. After a screening, 4-(dimethylamino)benzamide was determined to be the best protecting group for the amine function. In the presence of (i-PrCO)2O as the acylating agent and the chiral amino pyridine, the S alcohol from mixture 91 was preferentially acylated to 92, and the unconverted protected β-amido-alcohol (1S,2R)-91 could be recovered with an 99% ee after 64% conversion. A treatment with 6M HCl provided the targeted (1S,2R)-1-amino-2-indanol 1 with 68% yield. An analogous approach was proposed by Campbell with a N-4′-pyridinyl-α-methyl proline derivative as the catalyst and trifluoroacetyl as the nitrogen protecting group in order to facilitate its removal [95].
Resolution of the racemic cis-1-amino-2-indanol can also be achieved via diastereomeric salt formation (Scheme 32). (S)-2-Phenylpropionic acid proved to be an efficient resolving agent, inducing a selective crystallization of the ammonium salt formed with (1R,2S)-1-amino-2-indanol (1). After filtration, the amino alcohol could be released under a basic work-up, leading to an enantiopure product with 35% yield, and the resolving agent recovered with 93% yield [96]. Similar results were obtained using tartaric acid as a resolving agent, either to perform enantioenrichment [15] or complete resolution [97].

4.2. Enzymatic Resolution

Enzymatic resolution is an interesting alternative. Various approaches have been devised involving the reactivity of alcohol or amine functions. Gotor’s group proposed the acylation of N-Cbz protected racemic cis-1-amino-2-indanol 93 with vinyl acetate catalyzed by Pseudomonas cepacia lipase (PSL) (Scheme 33). These conditions enabled a R selective acylation of the alcohol 93 (43%, >99% ee) with 44% conversion [98].
The enzyme’s ability to hydrolyze an ester was tested on N,O-diacetyl-cis-1-amino-2-indanol 96 (Scheme 34) [99]. The alcoholysis of 96 was carried out in the presence of Candida antartica lipase B (CAL-B) and n-butanol. The very high enantiomeric ratio (E > 500) provided both the hydrolyzed ester 97 as well as the unreactive ester 98 with excellent ee and yields. The targeted amino-indanol was released under basic conditions without eroding the ee.
Interestingly, more recently, a continuous-flow resolution was directly performed on cis-1-amino-2-indanol with CAL-B immobilized on an acrylic resin (Novozym 435® (N435)) with EtOAc as the acyl donor (Scheme 35) [100]. The flow system enabled selective acylation of the amino group of the (1S,2R) substrate, whereas analogous conditions in vials were much less effective, certainly because the controlled flow rate increased the local concentration of immobilized CAL-B, and consequently, the amino group acylation rate. The result was a conversion of 50% and an ee of 99% for the two alcohols recovered with E > 200.

5. Substituted cis-1-Amino-2-Indanol

Due to the high interest in them for catalysis and for the development of drug candidates, numerous cis-1-amino-2-indanol derivatives with a substituent on the aromatic ring have been prepared. Two different strategies are implemented: (i) a post-functionalization approach, in which the cis-1-amino-2-indanol moiety is already formed, and substituents are introduced via coupling reactions; or (ii) a pre-functionalization approach, in which the introduction of the substituent to the aromatic ring is performed before the formation of the cis-1-amino-2-indanol skeleton with the strategies described above. This section presents an overview of the structure of substituted cis-1-amino-2-indanols and their preparation strategy.

5.1. Post-Functionalization

The post-functionalization strategy has been used in particular to prepare new catalysts based on cis-1-amino-2-indanol architecture (Scheme 36). This strategy offers the advantage of utilizing the commercially available enantiopure cis-1-amino-2-indanol, which is first protected under the carbamate form. At this stage, functionalization at position 6 can be performed through an electrophilic reaction, such as bromination [71,72,101], nitration [102] or Friedel–Crafts reaction [71]. In addition to the Friedel–Crafts reaction, the C–C bond can be formed in two steps via a metal-catalyzed coupling reaction on the corresponding bromide [71,72,73,101]; an alkyl or aryl substituent can be introduced in this manner (Table 1). It should be pointed out that the regioselectivity of the electrophilic reaction mainly leads to substitution at position 6 (R6), which limits the scope of this approach.

5.2. Pre-Functionalization

The pre-functionalization approach implements the various strategies discussed earlier in this review, namely the Ritter reaction (Strategy A, Scheme 2), the Mitsunobu reaction (Strategy B, Scheme 2) or a reductive approach (Strategy C, Scheme 2). The difference lies in the starting substrate, indanone, which already carries substituents on the aromatic ring (Scheme 37). It offers the possibility of choosing different regioisomers, as well as poly-substituted products (Table 2). However, depending on the regioisomer desired, the scope of the reaction is limited. For example, at position 7 (R7), only alkyl groups were described (entries 1 and 2) [86,104,105,106,107]. Regarding position 6 (R6), a larger variety of electron-donating or electron-withdrawing substituents (fluorine [108], bromine [109], methoxy [110,111] and nitro [112]) were introduced onto indanone prior to conversion, mainly via a Ritter-type reaction (entries 3–6). After the formation of cis-1-amino-2-indanol, the nitro group can be easily reduced to an amino group [112]. The introduction of fluorine [108], chlorine [113] or bromine [109] using this strategy is also possible at position 5 (R5) (entries 7–9). It may be underlined that the presence of bromine in position 5, which is not possible using the post-functionalization approach, opens the way to subsequent functionalization. Substitution at position 4 (R4) seems to be more difficult (or less interesting) to achieve, since only methyl (entry 10) [86,106] and fluoro (entry 11) [114] derivatives were prepared. Di-bromo-substituted [86] and polyaromatic indanones [115,116,117,118] were also used for the preparation of cis-1-amino-2-indanol moieties (entries 12–14).

6. Conclusions

cis-1-amino-2-indanol is a ubiquitous substructure in many areas of chemistry, and its synthesis has attracted the attention of many chemists in both academic and industrial fields for decades. A wide range of strategies have been implemented to prepare this molecule in order to overcome its main structural feature, namely the cis-introduction of an amino group and an alcohol group in positions 1 and 2, respectively. Most of the time, stereochemistry is controlled by forming a 5-membered intermediate ring. In this context, indene has proven to be a versatile starting material, either as a precursor of key functional groups, such as epoxide, diol, halohydrin in their racemic or enantiopure form, or directly functionalized via electrophilic activation of its double bond. Among the implemented strategies, direct C-H functionalization has recently attracted a great deal of interest. These promising approaches involve nitrene or radical intermediates. However, such strategies are mainly developed in their racemic versions, and the few attempts of enantioselective approaches have yielded low enantiomeric excesses. Moreover, the scope of all reactions is particularly limited in terms of the diversity of substitution on the aromatic ring. Position 6 offers the largest possibility of substitution by an electron-withdrawing or electron-donating group. The examples for positions 4, 5 and 7 are scarce and only concern alkyl or halogen groups. No electron-rich or electron-poor groups have been introduced on these positions, and consequently, their impact on the methods implemented for the introduction of the amino-alcohol moiety—in the pre-functionalization strategy—remains unknown. In this context, the development of functionalizations through Csp3-H activation provides interesting alternatives, as the reaction intermediates involved are different from those involved in older synthetic approaches.

Author Contributions

Conceptualization, S.G. and J.-S.S.; methodology, S.G. and J.-S.S.; writing—original draft preparation, I.M., S.G. and J.-S.S.; writing—review and editing, I.M., S.G. and J.-S.S.; supervision, S.G. and J.-S.S.; project administration, S.G. and J.-S.S.; funding acquisition, J.-S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Agence Nationale de la Recherche (ANR-21-CE07-0012-01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Saddique, F.A.; Zahoor, A.F.; Faiz, S.; Naqvi, A.A.R.; Usman, M.; Ahmad, M. Recent trends in ring opening of epoxides by amines as nucleophiles. Synth. Commun. 2016, 46, 831–868. [Google Scholar] [CrossRef]
  2. Weng, C.; Zhang, H.; Xiong, X.; Lu, X.; Zhou, Y. Evolution of Epoxides to Synthesize β-amino Alcohols. Asian J. Org. Chem. 2014, 26, 3761–3768. [Google Scholar] [CrossRef]
  3. Karjalainen, O.K.; Koskinen, A.M.P. Diastereoselective synthesis of vicinal amino alcohols. Org. Biomol. Chem. 2012, 10, 4311–4326. [Google Scholar] [CrossRef]
  4. Bergmeier, S.C. The Synthesis of Vicinal Amino Alcohols. Tetrahedron 2000, 56, 2561–2576. [Google Scholar] [CrossRef]
  5. Desimoni, G.; Faita, G.; Jørgensen, K.A. C2-Symmetric Chiral Bis(Oxazoline) Ligands in Asymmetric Catalysis. Chem. Rev. 2006, 106, 3561–3651. [Google Scholar] [CrossRef]
  6. Desimoni, G.; Faita, G.; Quadrelli, P. Pyridine-2,6-bis(oxazolines), Helpful Ligands for Asymmetric Catalysts. Chem. Rev. 2003, 103, 3119–3154. [Google Scholar] [CrossRef]
  7. Babu, S.A.; Krishnan, K.K.; Ujwaldev, S.M.; Anilkumar, G. Applications of Pybox Complexes in Asymmetric Catalysis. Asian J. Org. Chem. 2018, 7, 1033–1053. [Google Scholar] [CrossRef]
  8. Higashijima, S.; Itoh, H.; Senda, Y.; Nakano, S. cis-1-Amino-1,2,3,4-tetrahydro-2-naphthalenol: Resolution and application to the catalytic enantioselective reduction of ketones. Tetrahedron Asymmetry 1997, 8, 3107–3110. [Google Scholar] [CrossRef]
  9. Gallou, I.; Senanayake, C.H. cis-1-Amino-2-indanol in Drug Design and Applications to Asymmetric Processes. Chem. Rev. 2006, 106, 2843–2874. [Google Scholar] [CrossRef]
  10. Dorsey, B.D.; Levin, R.B.; McDaniel, S.L.; Vacca, J.P.; Guare, J.P.; Darke, P.L.; Zugay, J.A.; Emini, E.A.; Schleif, W.A.; Quintero, J.C.; et al. RL-735,524: The Design of a Potent and Orally Bioavailable HIV Protease Inhibitor. J. Med. Chem. 1994, 37, 3443–3451. [Google Scholar] [CrossRef]
  11. Reider, P.J. Advances in AIDS Chemotherapy: The Asymmetric Synthesis of CRIXIVAN®. Chimia 1997, 51, 306–308. [Google Scholar] [CrossRef]
  12. Abdel-Rahman, H.; Kimura, T.; Hidaka, K.; Kiso, A.; Nezami, A.; Freire, E.; Hayashi, Y.; Kiso, Y. Design of inhibitors against HIV, HTLV-I, and Plasmodium falciparum aspartic proteases. Biol. Chem. 2004, 385, 1035–1039. [Google Scholar] [CrossRef]
  13. Ghosh, A.K.; Fidanze, S.; Senanayakeb, C.H. cis-1-Aminoindan-2-ol in Asymmetric Syntheses. Synthesis 1998, 1998, 937–961. [Google Scholar] [CrossRef]
  14. Senanayake, C.; Roberts, F.E.; DiMichele, L.M.; Ryan, K.M.; Liu, J.; Fredenburgh, L.E.; Foster, B.S.; Douglas, A.W.; Larsen, R.D.; Verhoeven, T.R.; et al. The Behavior of Indene Oxide in the Ritter Reaction: A Simple Route to cis-Aminoindanol. Tetrahedron Lett. 1995, 36, 3993–3996. [Google Scholar] [CrossRef]
  15. Larrow, J.F.; Roberts, E.; Verhoeven, T.R.; Ryan, K.M.; Senanayake, C.H.; Reider, P.J.; Jacobsen, E.N. (1S,2R)-1-AMINOINDAN-2-OL. Org. Synth. 1998, 76, 46. [Google Scholar]
  16. Senanayake, C.H.; DiMichele, L.M.; Liu, J.; Fredenburgh, L.E.; Ryan, K.M.; Roberts, F.E.; Larsen, R.D.; Verhoeven, R.R.; Reider, P.J. Regio- and Stereocontrolled Syntheses of Cyclic Chiral cis-Amino Alcohols from 1,2-Diols or Epoxides. Tetrahedron Lett. 1995, 36, 7615–7618. [Google Scholar] [CrossRef]
  17. Alexander, G.; Medvedev, A.G.; Grishanov, D.A.; Mikhaylov, A.A.; Churakov, A.V.; Tripol’skaya, T.A.; Ottenbacher, R.V.; Bryliakov, K.P.; Shames, A.I.; Lev, O.; et al. V Triphenyllead Hydroperoxide: A 1D Coordination Peroxo Polymer, Single-Crystal-to-Single-Crystal Disproportionation to a Superoxo/Hydroxo Complex, and Application in Catalysis. Inorg. Chem. 2022, 61, 8193–8205. [Google Scholar]
  18. Huang, J.; Liu, S.; Ma, Y.; Cai, J. Chiral salen Mn (III) immobilized on ZnPS-PVPA through alkoxyl-triazole for superior performance catalyst in asymmetric epoxidation of unfunctionalized olefins. J. Organomet. Chem. 2019, 886, 27–33. [Google Scholar] [CrossRef]
  19. Kratky, J.; Eggerichs, D.; Heine, T.; Hofmann, S.; Sowa, P.; Weiße, R.H.; Tischler, D.; Sträter, N. Structural and Mechanistic Studies on Substrate and Stereoselectivity of the Indole Monooxygenase VpIndA1: New Avenues for Biocatalytic Epoxidations and Sulfoxidations. Angew. Chem. Int. Ed. 2023, 62, e202300657. [Google Scholar] [CrossRef] [PubMed]
  20. Toda, H.; Imae, R.; Itoh, N. Efficient biocatalysis for the production of enantiopure (S)-epoxides using a styrene monooxygenase (SMO) and Leifsonia alcohol dehydrogenase (LSADH) system. Tetrahedron Asymmetry 2012, 23, 1542–1549. [Google Scholar] [CrossRef]
  21. Martínez-Montero, L.; Tischler, D.; Süss, P.; Schallmey, A.; Franssen, M.C.R.; Hollmann, F.; Paul, C.E. Asymmetric azidohydroxylation of styrene derivatives mediated by a biomimetic styrene monooxygenase enzymatic cascade. Catal. Sci. Technol. 2021, 11, 5077–5085. [Google Scholar] [CrossRef]
  22. He, G.-H.; Liu, Y.-L.; Liu, H.; Lu, X.-B. Enantioselective Resolution Copolymerization of Racemic cis-Epoxides and Cyclic Anhydrides Mediated by Multichiral Bimetallic Chromium Complexes. Macromolecules 2022, 55, 3869–3876. [Google Scholar] [CrossRef]
  23. Zhang, J.; Reddy, J.; Roberge, C.; Senanayake, C.; Greashman, R.; Chartrain, M. Chiral Bio-Resolution of Racemic Indene Oxide by Fungal Epoxide Hydrolases. J. Ferment. Bioeng. 1995, 80, 244–246. [Google Scholar] [CrossRef]
  24. Gruttadauria, M.; Lo Meo, P.; Riela, S.; D’Anna, F.; Noto, R. Lipase-catalyzed resolution of β-hydroxy selenides. Tetrahedron Asymmetry 2006, 17, 2713–2721. [Google Scholar] [CrossRef]
  25. Igarashi, Y.; Otsutomo, S.; Harada, M.; Nakano, S.; Watanabe, S. Lipase-Mediated Resolution of Indene Bromohydrin. Synthesis 1997, 5, 549–552. [Google Scholar] [CrossRef]
  26. Hao, B.; Gunaratna, M.J.; Zhang, M.; Weerasekara, S.; Seiwald, S.N.; Nguyen, V.T.; Meier, A.; Hua, D.H. Chiral-Substituted Poly-N-vinylpyrrolidinones and Bimetallic Nanoclusters in Catalytic Asymmetric Oxidation Reactions. J. Am. Chem. Soc. 2016, 138, 16839–16848. [Google Scholar] [CrossRef]
  27. Köhler, V.; Mao, J.; Heinisch, T.; Pordea, A.; Sardo, A.; Wilson, Y.M.; Knörr, L.; Creus, M.; Prost, J.-C.; Schirmer, T.; et al. OsO4·Streptavidin: A Tunable Hybrid Catalyst for the Enantioselective cis-Dihydroxylation of Olefins. Angew. Chem. Int. Ed. 2011, 50, 10863–10866. [Google Scholar] [CrossRef] [PubMed]
  28. Yan, L.; Meng, Y.; Haeffner, F.; Leon, R.M.; Crockett, M.P.; Morken, J.P. Carbohydrate/DBU Cocatalyzed Alkene Diboration: Mechanistic Insight Provides Enhanced Catalytic Efficiency and Substrate Scope. J. Am. Chem. Soc. 2018, 140, 3663–3673. [Google Scholar] [CrossRef]
  29. Zhu, J.; Sun, X.-T.; Wang, X.-D.; Wu, L. Enantioselective Dihydroxylation of Alkenes Catalyzed by 1,4-Bis(9-O-dihydroquinidinyl) phthalazine-Modified Binaphthyl-Osmium Nanoparticles. ChemCatChem 2018, 10, 1788–1792. [Google Scholar] [CrossRef]
  30. Tetour, D.; Paška, T.; Máková, V.; Holubová, B.N.; Karpíšková, J.; Řezanka, M.; Brus, J.; Hodačová, J. Cinchonine-based organosilica materials as heterogeneous catalysts of enantioselective alkene dihydroxylation. J. Catal. 2021, 404, 493–500. [Google Scholar] [CrossRef]
  31. Nakano, S.; Igarashi, Y.; Nohira, H. Lipase-mediated kinetic resolution of cis-1,2-indandiol and the Ritter reaction of its mono-acetate. Tetrahedron Asymmetry 2001, 12, 59–62. [Google Scholar] [CrossRef]
  32. Huang, H.; Lambert, T.H. Regiodivergent Electrophotocatalytic Aminooxygenation of Aryl Olefins. J. Am. Chem. Soc. 2022, 144, 18803–18809. [Google Scholar] [CrossRef] [PubMed]
  33. Lutz, R.E.; Wayland, R.L. Further Studies on the Stability of β-Hydroxyethylamines toward the Oppenauer Oxidation. cis- and trans-1-Amino-2-indanols. J. Am. Chem. Soc. 1951, 73, 1639–1641. [Google Scholar]
  34. Gao, Y.; Hong, Y.; Nie, X.; Bakale, R.P.; Feinberg, R.R.; Zepp, C.M. Optically Pure 1-Amino 2-Indanols. U.S. Patent 5,616,808, 1997. [Google Scholar]
  35. Ghosh, A.K.; Kincaid, J.F.; Haske, M.G. A Convenient Enzymatic Route to Optically Active l-Aminoindan-2-ol: Versatile Ligands for HIV-1 Protease Inhibitors and Asymmetric Syntheses. Synthesis 1997, 5, 541–544. [Google Scholar] [CrossRef] [PubMed]
  36. Didier, E.; Loubinoux, B.; Ramos Tombo, G.M.; Rihs, G. Chemo-Enzymatic Synthesis of 1,2- and 1,3- Amino-Alcohols and their Use in the Enantioselective Reduction of Acetophenone and Anti-Acetophenone Oxine Methyl Ether with Borane. Tetrahedron 1991, 47, 4941–4958. [Google Scholar] [CrossRef]
  37. Hassner, A.; Lorber, M.E.; Heathcock, C. Addition of Iodine Isocyanate to Olefins. Scope and Synthetic Utility. J. Org. Chem. 1967, 32, 540–549. [Google Scholar] [CrossRef] [PubMed]
  38. Hu, N.X.; Aso, Y.; Otsubo, T.; Ogura, F. Aminotellurinylation of Olefins and Its Utilization for Synthesis of 2-Oxazolidinones. J. Org. Chem. 1989, 54, 4398–4404. [Google Scholar] [CrossRef]
  39. Hu, N.X.; Aso, Y.; Ogura, O.J. Amidotellurinylation of Olefins and a Novel One-Pot Synthesis of 4,5-Dihydro-oxazoles from Olefins. J. Chem. Soc. Perkin Trans. 1 1989, 1775–1780. [Google Scholar] [CrossRef]
  40. Kuszpit, M.R.; Giletto, M.B.; Jones, C.L.; Bethel, T.K.; Tepe, J.J. Hydroxyamination of Olefins Using Br-N-(CO2Me)2. J. Org. Chem. 2015, 80, 1440–1445. [Google Scholar] [CrossRef]
  41. Wu, F.; Alom, N.-E.; Ariyarathna, J.P.; Naß, J.; Li, W. Regioselective Formal [3+2] Cycloadditions of Urea Substrates with Activated and Unactivated Olefins for Intermolecular Olefin Aminooxygenation. Angew. Chem. Int. Ed. 2019, 58, 11676–11680. [Google Scholar] [CrossRef]
  42. Wu, F.; Kaur, N.; Alom, N.-E.; Li, W. Chiral Hypervalent Iodine Catalysis Enables an Unusual Regiodivergent Intermolecular Olefin Aminooxygenation. JACS Au 2021, 1, 734–741. [Google Scholar] [CrossRef]
  43. Wata, C.; Hashimoto, T. Organoiodine-Catalyzed Enantioselective Intermolecular Oxyamination of Alkenes. J. Am. Chem. Soc. 2021, 143, 1745–1751. [Google Scholar] [CrossRef] [PubMed]
  44. Wappes, E.A.; Nakafuku, K.M.; Nagib, D.A. Directed β C−H Amination of Alcohols via Radical Relay Chaperones. J. Am. Chem. Soc. 2017, 139, 10204–10207. [Google Scholar] [CrossRef] [PubMed]
  45. Mou, X.-Q.; Chen, X.-Y.; Chen, G.; He, G. Radical-mediated intramolecular β–C(sp3)–H amidation of alkylimidates: Facile synthesis of 1,2-amino alcohols. Chem. Commun. 2018, 54, 515–518. [Google Scholar] [CrossRef] [PubMed]
  46. Stateman, L.M.; Wappes, E.A.; Nakafuku, K.M.; Edwards, K.M.; Nagib, D.A. Catalytic β C–H amination via an imidate radical Relay. Chem. Sci. 2019, 10, 2693–2699. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, R.; Fu, K.; Fang, Y.; Zjou, J.; Shi, L. Site-Specific C(sp3)–H Aminations of Imidates and Amidines Enabled by Covalently Tethered Distonic Radical Anions. Angew. Chem. Int. Ed. 2020, 59, 20682–20690. [Google Scholar] [CrossRef]
  48. Nakafuku, K.M.; Zhang, Z.; Wappes, E.A.; Stateman, L.M.; Chen, A.D.; Nagib, D.A. Enantioselective radical C–H amination for the synthesis of β-amino alcohols. Nat. Chem. 2020, 12, 697–704. [Google Scholar] [CrossRef] [PubMed]
  49. Ma, B.; Xia, Q.; Wand, D.; Jin, J.-K.; Li, Z.; Liand, Q.-J.; Sun, M.-Y.; Liu, D.; Liu, L.-J.; Shu, H.-X.; et al. Metal-Organic Framework Supported Copper Photoredox Catalysts for Iminyl Radical-Mediated Reactions. Angew. Chem. Int. Ed. 2023, 62, e202300233. [Google Scholar] [CrossRef] [PubMed]
  50. Barman, D.N.; Nicholas, K.M. Copper-Catalyzed Intramolecular C–H Amination. Eur. J. Org. Chem. 2011, 2011, 908–911. [Google Scholar] [CrossRef]
  51. Barman, D.N.; Liu, P.; Houk, K.N.; Nicholas, K.M. On the Mechanism of Ligand-Assisted, Copper-Catalyzed Benzylic Amination by Chloramine-T. Organometallics 2010, 29, 3404–3412. [Google Scholar] [CrossRef]
  52. Noshita, M.; Shimizu, Y.; Morimoto, H.; Oshsima, T. Diethylenetriamine-Mediated Direct Cleavage of Unactivated Carbamates and Ureas. Org. Lett. 2016, 18, 6062–6065. [Google Scholar] [CrossRef] [PubMed]
  53. Espino, C.G.; Du Bois, J. A Rh-Catalyzed C-H Insertion Reaction for the Oxidative Conversion of Carbamates to Oxazolidinones. Angew. Chem. Int. Ed. 2001, 40, 598–600. [Google Scholar] [CrossRef]
  54. Cui, Y.; He, C. A Silver-Catalyzed Intramolecular Amidation of Saturated C-H Bonds. Angew. Chem. Int. Ed. 2004, 43, 4210–4212. [Google Scholar] [CrossRef] [PubMed]
  55. Espino, C.G.; Wehn, P.M.; Chow, J.; Du Bois, J. Synthesis of 1,3-Difunctionalized Amine Derivatives through Selective C-H Bond Oxidation. J. Am. Chem. Soc. 2001, 123, 6935–6936. [Google Scholar] [CrossRef]
  56. Liang, J.-L.; Yuan, S.-X.; Huang, J.-S.; Yu, W.-Y.; Che, C.-M. Highly Diastereo- and Enantioselective Intramolecular Amidation of Saturated C-H Bonds Catalyzed by Ruthenium Porphyrins. Angew. Chem. Int. Ed. 2002, 41, 3465–3468. [Google Scholar] [CrossRef]
  57. Liu, P.; Wong, E.L.-M.; Yuen, A.W.-H.; Che, C.-M. Highly Efficient Alkene Epoxidation and Aziridination Catalyzed by Iron(II) Salt + 4,4′,4″-Trichloro-2,2′:6′,2″-terpyridine/ 4,4″-Dichloro-4′-O-PEG-OCH3-2,2′:6′,2″-terpyridine. Org. Lett. 2008, 10, 3275–3278. [Google Scholar] [CrossRef]
  58. Liu, Y.; Guan, X.; Wong, E.L.-M.; Liu, P.; Huang, J.-S.; Che, C.-M. Nonheme Iron-Mediated Amination of C(sp3)−H Bonds. Quinquepyridine-Supported Iron-Imide/Nitrene Intermediates by Experimental Studies and DFT Calculations. J. Am. Chem. Soc. 2013, 135, 7194–7204. [Google Scholar] [CrossRef]
  59. Liang, J.-L.; Yuan, S.-X.; Huang, J.-S.; Che, C.-M. Intramolecular C-N Bond Formation Reactions Catalyzed by Ruthenium Porphyrins: Amidation of Sulfamate Esters and Aziridination of Unsaturated Sulfonamides. J. Org. Chem. 2004, 69, 3610–3619. [Google Scholar] [CrossRef]
  60. Zhang, J.; Chan, P.W.H.; Che, C.-M. Enantioselective intramolecular amidation of sulfamate esters catalyzed by chiral manganese(III) Schiff-base complexes. Tetrahedron Lett. 2005, 46, 5403–5408. [Google Scholar] [CrossRef]
  61. Lee, J.; Lee, J.; Jung, H.; Kim, D.; Park, J.; Chang, S. Versatile Cp*Co(III)(LX) Catalyst System for Selective Intramolecular C−H Amidation Reactions. J. Am. Chem. Soc. 2020, 142, 12324–12332. [Google Scholar] [CrossRef]
  62. Lebel, H.; Huard, K.; Lectard, S. N-Tosyloxycarbamates as a Source of Metal Nitrenes: Rhodium-Catalyzed C-H Insertion and Aziridination Reactions. J. Am. Chem. Soc. 2005, 127, 14198–14199. [Google Scholar] [CrossRef] [PubMed]
  63. Huard, K.; Lebel, H. N-Tosyloxycarbamates as Reagents in Rhodium-Catalyzed C-H Amination Reactions. Chem. Eur. J. 2008, 14, 6222–6230. [Google Scholar] [CrossRef]
  64. Lebel, H.; Mamani Laparra, L.; Khalifa, M.; Trudel, C.; Audubert, C.; Szponarski, M.; Dicaire Leduc, C.; Azek, E.; Ernzerhof, M. Synthesis of oxazolidinones: Rhodium-catalyzed C–H amination of N-mesyloxycarbamates. Org. Biomol. Chem. 2017, 15, 4144–4158. [Google Scholar] [CrossRef] [PubMed]
  65. Jeong, J.; Jung, H.; Kim, D.; Chang, S. Multidimensional Screening Accelerates the Discovery of Rhodium Catalyst Systems for Selective Intra- and Intermolecular C−H Amidations. ACS Catal. 2022, 12, 8127–8138. [Google Scholar] [CrossRef]
  66. Grelier, G.; Rey-Rodriguez, R.; Darses, B.; Retailleau, P.; Dauban, P. Catalytic Intramolecular C(sp3)–H Amination of Carbamimidates. Eur. J. Org. Chem. 2017, 2017, 1880–1883. [Google Scholar] [CrossRef]
  67. Milczek, E.; Boudet, N.; Blakey, S. Enantioselective C-H Amination Using Cationic Ruthenium(II)–pybox Catalysts. Angew. Chem. Int. Ed. 2008, 47, 6825–6828. [Google Scholar] [CrossRef] [PubMed]
  68. Yamawaki, M.; Kitagaki, S.; Anada, M.; Hashimoto, S. Enantioselective Intramolecular C–H Amidation of Sulfamate Esters Catalyzed by Chiral Dirhodium(II) Carboxylates. Heterocycles 2006, 69, 527–537. [Google Scholar] [CrossRef]
  69. Reddy, R.P.; Davies, H.M.L. Dirhodium Tetracarboxylates Derived from Adamantylglycine as Chiral Catalysts for Enantioselective C-H Aminations. Org. Lett. 2006, 8, 5013–5016. [Google Scholar] [CrossRef]
  70. Zhou, Z.; Tan, Y.; Shen, X.; Ivlev, S.; Meggers, E. Catalytic enantioselective synthesis of β-amino alcohols by nitrene insertion. Sci. China Chem. 2021, 64, 452–458. [Google Scholar] [CrossRef]
  71. Hu, Q.; Guo, C. Platinum/Copper Dual-Catalyzed Asymmetric Vinylogous Addition Reactions for the Synthesis of Functionalized Indoles. Angew. Chem. Int. Ed. 2023, 62, e202305638. [Google Scholar] [CrossRef]
  72. Buchsteiner, M.; Martinez-Rodriguez, L.; Jerabek, P.; Pozo, I.; Patzer, M.; Nöthling, N.; Lehmann, C.W.; Fürstner, A. Catalytic Asymmetric Fluorination of Copper Carbene Complexes: Preparative Advances and a Mechanistic Rationale. Chem. Eur. J. 2020, 26, 2509–2515. [Google Scholar] [CrossRef]
  73. Liu, C.; Yi, J.-C.; Zheng, Z.-B.; Tang, Y.; Dai, L.-X.; You, S.-L. Enantioselective Synthesis of 3a-Amino-Pyrroloindolines by Copper-Catalyzed Direct Asymmetric Dearomative Amination of Tryptamines. Angew. Chem. Int. Ed. 2016, 55, 751–754. [Google Scholar] [CrossRef] [PubMed]
  74. Zeng, J.; Sun, G.; Yao, W.; Zhu, Y.; Wang, R.; Cai, L.; Liu, K.; Zhang, Q.; Liu, X.-W.; Wan, Q. 3-Aminodeoxypyranoses in Glycosylation: Diversity-Oriented Synthesis and Assembly in Oligosaccharides. Angew. Chem. Int. Ed. 2017, 56, 5227–5231. [Google Scholar] [CrossRef]
  75. Allison, B.D.; Mani, N.S. Gram-Scale Synthesis of a β-Secretase 1 (BACE 1) Inhibitor. ACS Omega 2017, 2, 397–408. [Google Scholar] [CrossRef]
  76. Zeng, J.; Wang, R.; Yao, W.; Zhang, S.; Sun, G.; Liao, Z.; Menga, L.; Wan, Q. Diversified synthesis and α-selective glycosylation of 3-amino-2,3,6-trideoxy sugars. Org. Chem. Front. 2018, 5, 3391–3395. [Google Scholar] [CrossRef]
  77. Strambeanu, I.I.; White, M.C. Catalyst-Controlled C−O versus C−N Allylic Functionalization of Terminal Olefins. J. Am. Chem. Soc. 2013, 135, 12032–12037. [Google Scholar] [CrossRef] [PubMed]
  78. Cochran, B.M.; Michael, F.E. Metal-Free Oxidative Cyclization of Urea-Tethered Alkenes with Hypervalent Iodine. Org. Lett. 2008, 10, 5039–5042. [Google Scholar] [CrossRef] [PubMed]
  79. Lakshman, M.K.; Chaturvedi, S.; Zajc, B.; Gibson, D.T.; Resnick, S.M. A General Chemoenzymatic Synthesis of Enantiopure cis β-Amino Alcohols from Microbially Derived cis-Glycols. Synthesis 1998, 1998, 1352–1356. [Google Scholar] [CrossRef]
  80. Boyd, D.R.; Sharma, N.D.; Bowers, N.L.; Goodrich, P.A.; Groocock, M.R.; Blacker, A.J.; Clarke, D.A.; Howard, T.; Dalton, H. Stereoselective Dioxygenase-Catalysed Benzylic Hydroxylation at Prochiral Methylene Groups in the Chemoenzymatic Synthesis of Enantiopure Vicinal Aminoindanols. Tetrahedron Asymmetry 1996, 7, 1559–1562. [Google Scholar] [CrossRef]
  81. Ogasawara, K.; Takahashi, M. Lipase-Mediated Resolution of trans-1-Azidoindan-2-ol: A New Route to Optically Pure cis-1-Aminoindan-2-ol. Synthesis 1996, 1996, 954–958. [Google Scholar]
  82. Kajiro, H.; Mitamura, S.; Mori, A.; Hiyama, T.A. Practical Synthesis of (1S,2R)-1-Amino-2-indanol, a Key Component of an HIV Protease Inhibitor, Indinavir. Bull. Chem. Soc. Jpn. 1999, 72, 1093–1100. [Google Scholar] [CrossRef]
  83. Ayhan, S.; Demir, A.S.; Hamamci, H.; Doganel, F.; Ozgul, E. Chemoenzymatic synthesis of (1S,2 R)-1-amino-2-indanol, a key intermediate of HIV protease inhibitor, indinavir. J. Mol. Catal. B Enzym. 2000, 9, 157–161. [Google Scholar]
  84. Komiya, N.; Noji, S.; Murahashi, S.-I. Manganese Catalyzed Asymmetric Oxidation of Alkanes to Optically Active Ketones Bearing Asymmetric Center at the α-Position. Tetrahedron Lett. 1998, 39, 7921–7924. [Google Scholar] [CrossRef]
  85. Nguyen, T.H.; Ma, E. Efficient diastereoselective synthesis of cis-2-amino-1-indanol derivatives and cis- and trans-1-amino-2-indanol via Pd-catalyzed hydrogenation. Synth. Commun. 2021, 51, 3717–3728. [Google Scholar] [CrossRef]
  86. Yadav, J.S.; Reddy, G.S.K.K.; Sabitha, G.; Krishna, A.D.; Prasad, A.R.; Rahaman, H.U.R.; Rao, K.V.; Rao, A.B. Daucus carota and baker’s yeast mediated bio-reduction of prochiral ketones. Tetrahedron Asymmetry 2007, 18, 717–723. [Google Scholar] [CrossRef]
  87. Kim, E.J.; An, K.M.; Ko, S.Y. Enantioselective Synthesis of Cyclic Amino Alcohols: Cis-1-Amino-2-indanol. Bull. Korean Chem. Soc. 2006, 27, 2019–2022. [Google Scholar] [CrossRef]
  88. Ko, S.Y. Unusual Regioselection in the Mitsunobu Reactions of syn-2,3-Dihydroxy Esters:  Synthesis of Statine and Its Diastereomer. J. Org. Chem. 2002, 67, 2689–2691. [Google Scholar] [CrossRef] [PubMed]
  89. Rosa-Barrales, A.; Pérez-Bautista, J.A.; Cruz-Gregorio, S.; Meza-Léon, R.L.; Quintero, L.; Cortezano-Arellano, O.; Sartillo-Piscil, F. Chiron approach and the [4 + 2] Diels-Alder cycloaddition of 2-pyrones for the synthesis of cis-(-)-aminoindan-2-ol. Results Chem. 2021, 3, 100174–100177. [Google Scholar] [CrossRef]
  90. Ramirez, E.; Sánchez, M.; Meza-Léon, R.L.; Quintero, L.; Sartillo-Piscil, F. Selective Wittig olefination in aqueous media for the rapid preparation of unsaturated 7,3-lactone-α-d-xylofuranose derivatives. Tetrahedron Lett. 2010, 51, 2178–2180. [Google Scholar] [CrossRef]
  91. Kajiro, J.; Mitamura, S.-I.; Mori, A.; Hiyama, T. A Practical Synthesis of (1S, 2R)-1-Amino-2-indanol, a Key Component of HIV Protease Inhibitor, Indinavir. Synlett 1998, 1, 51–52. [Google Scholar] [CrossRef]
  92. Urban, F.J.; Moore, B.S.J. Synthesis of optically active 2-benzyldihydrobenzopyrans for the hypoglycemic agent englitazone. Heterocyclic. Chem. 1992, 29, 431–438. [Google Scholar] [CrossRef]
  93. Thompson, W.J.; Fitzgerald, P.M.D.; Holloway, M.K.; Emini, E.A.; Darke, P.L.; McKeever, B.M.; Schleif, W.A.; Quintero, J.C.; Zugay, J.A.; Tucker, T.J.; et al. Synthesis and Antiviral Activity of a Series of HIV-1 Protease Inhibitors with Functionality Tethered to the Pt or P/ Phenyl Substituents: X-ray Crystal Structure Assisted Design. Med. Chem. 1992, 35, 1685–1701. [Google Scholar] [CrossRef] [PubMed]
  94. Kawabata, T.; Yamamoto, K.; Momose, Y.; Yoshida, H.; Nagaoka, Y.; Fuji, K. Kinetic resolution of amino alcohol derivatives with a chiral nucleophilic catalyst: Access to enantiopure cyclic cis-amino alcohols. Chem. Commun. 2001, 2700–2701. [Google Scholar] [CrossRef]
  95. Pelotier, B.; Priem, G.; Macdonald, S.J.F.; Anson, M.S.; Uptonc, R.J.; Campbell, I.B. Convenient method for the kinetic resolution of β-aminoalcohols. Tetrahedron Lett. 2005, 46, 9005–9007. [Google Scholar] [CrossRef]
  96. Sakuraia, R.; Sakaib, K. Resolution of racemic cis-1-amino-2-indanol by diastereomeric salt formation with (S)-2-phenylpropionic acid. Tetrahedron Asymmetry 2003, 14, 411–413. [Google Scholar] [CrossRef]
  97. Askin, D.; Cianciosi, S.J.; Hoerrner, R.S. Reductive Amination Process. U.S. Patent 5,508,404, 6 April 1996. [Google Scholar]
  98. Luna, A.; Maestro, A.; Astorga, C.; Gotor, V. Enzymatic resolution of (±)-cis- and (±)-trans-1-aminoindan-2-ol and (±)-cis- and (±)-trans-2-aminoindan-1-ol. Tetrahedron Asymmetry 1999, 10, 1969–1977. [Google Scholar] [CrossRef]
  99. Anilkumar, A.T.; Goto, K.; Takahashi, T.; Ishizaki, K.; Kaga, H. Lipase-mediated resolution of cis-1-amino-2-indanol, the key component of the HIV protease inhibitor indinavir. Tetrahedron Asymmetry 1999, 10, 2501–2503. [Google Scholar] [CrossRef]
  100. Kim, Y.-J.; Choi, Y.-S.; Yang, S.; Yang, W.R.; Jeong, J.-H. Continuous-Flow Kinetic Resolution of (±)-cis-1-Amino-2-indanol by Lipase-Catalyzed N-Acetylation. Synlett 2015, 26, 1981–1984. [Google Scholar]
  101. Chen, S.; Wu, L.; Shao, Q.; Yang, G.; Zhang, W. Pd(II)-Catalyzed Asymmetric 1,6-Conjugate Addition of Arylboronic Acids to Meldrum’s Acid-Derived Dienes. Chem. Commun. 2018, 54, 2522–2525. [Google Scholar] [CrossRef]
  102. Cao, S.; Hong, W.; Ye, Z.; Gong, L. Photocatalytic three-component asymmetric sulfonylation via direct C(sp3)–H functionalization. Nat. Commun. 2021, 12, 2377–2387. [Google Scholar] [CrossRef]
  103. Singh, R.; Bhamidipati, S. Tricyclic Carbamate JAK Inhibitors. WO 2010/039518 A2, 8 April 2010. [Google Scholar]
  104. Xu, H.; Wolf, C. Asymmetric Synthesis of Chiral 1,3-Diaminopropanols: Bisoxazolidine-Catalyzed C–C Bond Formation with α-Keto Amides. Angew. Chem. Int. Ed. 2011, 50, 12249–12252. [Google Scholar] [CrossRef] [PubMed]
  105. Tanaka, K.; Kobayashi, T.; Mori, H.; Katsumura, S. Development of Highly Stereoselective Asymmetric 6π-Azaelectrocyclization of Conformationally Flexible Linear 1-Azatrienes. From Determination of Multifunctional Chiral Amines, 7-Alkyl cis-1-Amino-2-indanols, to Application as a New Synthetic Strategy: Formal Synthesis of 20-Epiuleine. J. Org. Chem. 2004, 69, 5906–5925. [Google Scholar] [PubMed]
  106. Kobayashi, T.; Tanaka, K.; Miwa, J.; Katsumura, S. Synthesis of new chiral auxiliaries for 6p-azaelectrocyclization: 4- and 7-alkyl substituted cis-1-amino-2-indanols. Tetrahedron Asymmetry 2004, 15, 185–188. [Google Scholar] [CrossRef]
  107. Liu, S.Y.; Katsumura, S. An improved synthesis of (-)-7-isopropyl-cis-1-amino-2-indanolthe chiral auxiliary for 6π-azaelectrocyclization. Chin. Chem. Lett. 2009, 20, 1204–1206. [Google Scholar] [CrossRef]
  108. Liang, F.; Gibney, M.; Yeh, V.; Li, X.; Molteni, V.; Shaw, D.; Berman, A.M.; Lewis, S. Compounds and Compositions as PDGFR Kinase Inhibitors. WO 2013/033620 A1, 7 March 2013. [Google Scholar]
  109. Alley, K.A.; Clarkson, A.J.; Uehara, A.; Johnson, J.S. A Site-Specific Synthetic Route to Substituted Inda(box) Ligands. Org. Lett. 2023, 25, 9108–9113. [Google Scholar] [CrossRef]
  110. Marino, S.T.; Stachurska-Buczek, D.; Huggins, D.A.; Krywult, B.M.; Sheehan, C.S.; Nguyen, T.; Choi, N.; Parsons, J.G.; Griffiths, P.G.; James, I.W.; et al. Synthesis of Chiral Building Blocks for Use in Drug Discovery. Molecules 2004, 9, 405–426. [Google Scholar] [CrossRef]
  111. Marcin, L.R.; Good, A.C.; Wu, Y.-J.; Zuev, D.S.; Olson, R.E.; Wang, N. Diaminopropane Derived Macrocycles as Inhibitors of Beta Amyloid Production. U.S. Patent 2008/0194535 A1, 10 August 2008. [Google Scholar]
  112. Gross, M.F.; Beaudoin, S.; McNaughton-Smith, G.; Amato, G.S.; Castle, N.A.; Huang, C.; Zou, A.; Yu, W. Aryl sulfonamido indane inhibitors of the Kv1.5 ion channel. Bioorg. Med. Chem. Lett. 2007, 17, 2849–2853. [Google Scholar] [CrossRef]
  113. McCarron, A.S.; Sells, T.B.; Stirling, M.; Stroud, S.G. Inhibitors of NEDD8-Activating Enzyme. WO 2013/028832 A2, 19 August 2013. [Google Scholar]
  114. Graham, S.L.; Hungate, R.W.; Huff, J.L.; Lyle, T.A.; Young, S.D.; Britcher, S.F.; Scholz, T.H.; Payne, L.S.; Dorsey, B.D. HIV Protease Inhibitors Having Polyether Substituents. U.S. Patent 747156, 27 May 1992. [Google Scholar]
  115. Kobayashi, Y.; Kinbara, K.; Sato, M.; Saigo, K. Synthesis, Absolute Configuration, and Application of Enantiopure trans-1-Aminobenz[f]indan-2-ol. Chirality. 2005, 17, 108–112. [Google Scholar] [CrossRef]
  116. Kobayashi, Y.; Kurasawa, T.; Kinbara, K.; Saigo, K. Rational Design of CH/π Interaction Sites in a Basic Resolving Agent. J. Org. Chem. 2004, 69, 7436–7441. [Google Scholar] [CrossRef]
  117. Jin, Y.X.; Tan, G.; Choi, H.J.; Hyun, M.H. A Facile Synthesis of Optically Active Four Stereoisomers of 1-Aminobenz[f]indan-2-ol. Bull. Korean Chem. Soc. 2006, 27, 755–758. [Google Scholar] [CrossRef]
  118. Rendina, V.L.; Goetz, S.A.; Neitzel, A.E.; Kaplan, H.Z.; Kingsbury, J.S. Scalable synthesis of a new enantiomerically pure π-extended rigid amino indanol. Tetrahedron Lett. 2012, 53, 15–18. [Google Scholar] [CrossRef]
  119. Powers, T.; Steeneck, C.; Biesinger, R.; Bluhm, H.; Deng, H.; Dodd, R.; Gallagher JR, B.M.; Gege, C.; Hochgurtel, M.; Kiely, A.; et al. Multicyclic Bis-Amide MMP Inhibitors. U.S. Patent 2006/0173183 A, 3 August 2006. [Google Scholar]
Scheme 1. cis-1-amino-2-indanol structure and examples of applications.
Scheme 1. cis-1-amino-2-indanol structure and examples of applications.
Molecules 29 02442 sch001
Scheme 2. Main strategies developed for the synthesis of cis-1-amino-2-indanol.
Scheme 2. Main strategies developed for the synthesis of cis-1-amino-2-indanol.
Molecules 29 02442 sch002
Scheme 3. Ritter type reaction from indene oxide.
Scheme 3. Ritter type reaction from indene oxide.
Molecules 29 02442 sch003
Scheme 4. Enantioselective synthesis of (1S,2R)-1-amino-2-indanol via the Ritter process.
Scheme 4. Enantioselective synthesis of (1S,2R)-1-amino-2-indanol via the Ritter process.
Molecules 29 02442 sch004
Scheme 5. Proposed mechanism for Ritter type reaction from epoxide.
Scheme 5. Proposed mechanism for Ritter type reaction from epoxide.
Molecules 29 02442 sch005
Scheme 6. Ritter type reaction from diols.
Scheme 6. Ritter type reaction from diols.
Molecules 29 02442 sch006
Scheme 7. Ritter type reaction from indene under electrophotocatalytic conditions.
Scheme 7. Ritter type reaction from indene under electrophotocatalytic conditions.
Molecules 29 02442 sch007
Scheme 8. Proposed mechanism for Ritter type reaction from indene under electrophotocatalytic conditions.
Scheme 8. Proposed mechanism for Ritter type reaction from indene under electrophotocatalytic conditions.
Molecules 29 02442 sch008
Scheme 9. Racemic synthesis of 1 through intramolecular amide cyclization.
Scheme 9. Racemic synthesis of 1 through intramolecular amide cyclization.
Molecules 29 02442 sch009
Scheme 10. Sepracor synthesis of enantiopure (1S,2R)-1-amino-2-indanol 1.
Scheme 10. Sepracor synthesis of enantiopure (1S,2R)-1-amino-2-indanol 1.
Molecules 29 02442 sch010
Scheme 11. Ghosh synthesis of the two enantiomers of cis-1-amino-2-indanol.
Scheme 11. Ghosh synthesis of the two enantiomers of cis-1-amino-2-indanol.
Molecules 29 02442 sch011
Scheme 12. Didier synthesis of the two enantiomers of cis-1-amino-2-indanol.
Scheme 12. Didier synthesis of the two enantiomers of cis-1-amino-2-indanol.
Molecules 29 02442 sch012
Scheme 13. cis-1-amino-2-indanol from indene by iodoamidation.
Scheme 13. cis-1-amino-2-indanol from indene by iodoamidation.
Molecules 29 02442 sch013
Scheme 14. Telluric anhydride-induced formation of oxazolidinone.
Scheme 14. Telluric anhydride-induced formation of oxazolidinone.
Molecules 29 02442 sch014
Scheme 15. Recent approaches to the amino-oxygenation of indene.
Scheme 15. Recent approaches to the amino-oxygenation of indene.
Molecules 29 02442 sch015
Scheme 16. General mechanism of indene amino-oxygenation.
Scheme 16. General mechanism of indene amino-oxygenation.
Molecules 29 02442 sch016
Scheme 17. Radical-mediated C–N bond formation.
Scheme 17. Radical-mediated C–N bond formation.
Molecules 29 02442 sch017
Scheme 18. First enantioselective radical-mediated C1-N functionalization.
Scheme 18. First enantioselective radical-mediated C1-N functionalization.
Molecules 29 02442 sch018
Scheme 19. MOF-supported copper-catalyzed oxazoline formation.
Scheme 19. MOF-supported copper-catalyzed oxazoline formation.
Molecules 29 02442 sch019
Scheme 20. Copper-mediated intramolecular Csp3-H amination via radical pathway.
Scheme 20. Copper-mediated intramolecular Csp3-H amination via radical pathway.
Molecules 29 02442 sch020
Scheme 21. Approaches involving a nitrene intermediate.
Scheme 21. Approaches involving a nitrene intermediate.
Molecules 29 02442 sch021
Scheme 22. Meggers’ approach to the synthesis of cis-1-amino-2-indanol.
Scheme 22. Meggers’ approach to the synthesis of cis-1-amino-2-indanol.
Molecules 29 02442 sch022
Scheme 23. Resnick’s synthesis of cis-1-amino-2-indanol.
Scheme 23. Resnick’s synthesis of cis-1-amino-2-indanol.
Molecules 29 02442 sch023
Scheme 24. Synthesis of enantiopure cis-1-amino-2-indanol via resolution of the corresponding azido-alcohol.
Scheme 24. Synthesis of enantiopure cis-1-amino-2-indanol via resolution of the corresponding azido-alcohol.
Molecules 29 02442 sch024
Scheme 25. Synthesis of (1S,2R)-1 from 2-hydroxy-1-indanone.
Scheme 25. Synthesis of (1S,2R)-1 from 2-hydroxy-1-indanone.
Molecules 29 02442 sch025
Scheme 26. Synthesis of 1 from 1,2-indanedion-1-oxime.
Scheme 26. Synthesis of 1 from 1,2-indanedion-1-oxime.
Molecules 29 02442 sch026
Scheme 27. Ko’s synthesis of cis-1-amino-2-indanol.
Scheme 27. Ko’s synthesis of cis-1-amino-2-indanol.
Molecules 29 02442 sch027
Scheme 28. Synthesis of cis-1-amino-2-indanol from 7,3-xylofuranose derivative.
Scheme 28. Synthesis of cis-1-amino-2-indanol from 7,3-xylofuranose derivative.
Molecules 29 02442 sch028
Scheme 29. Synthesis of cis-1-amino-2-indanol from D-phenylalanine.
Scheme 29. Synthesis of cis-1-amino-2-indanol from D-phenylalanine.
Molecules 29 02442 sch029
Scheme 30. Chemical resolution of racemic cis-1-amino-2-indanol.
Scheme 30. Chemical resolution of racemic cis-1-amino-2-indanol.
Molecules 29 02442 sch030
Scheme 31. Kawabata’s kinetic resolution via enantioselective acylation.
Scheme 31. Kawabata’s kinetic resolution via enantioselective acylation.
Molecules 29 02442 sch031
Scheme 32. Resolution of cis-1-amino-2-indanol via diastereomeric salt formation.
Scheme 32. Resolution of cis-1-amino-2-indanol via diastereomeric salt formation.
Molecules 29 02442 sch032
Scheme 33. R selective enzymatic acylation of N-Cbz protected racemic cis-1-amino-2-indanol.
Scheme 33. R selective enzymatic acylation of N-Cbz protected racemic cis-1-amino-2-indanol.
Molecules 29 02442 sch033
Scheme 34. R selective enzymatic hydrolysis of N,O-Ac protected racemic cis-1-amino-2-indanol.
Scheme 34. R selective enzymatic hydrolysis of N,O-Ac protected racemic cis-1-amino-2-indanol.
Molecules 29 02442 sch034
Scheme 35. Continuous-flow kinetic resolution of cis-1-amino-2-indanol with Novozym 435.
Scheme 35. Continuous-flow kinetic resolution of cis-1-amino-2-indanol with Novozym 435.
Molecules 29 02442 sch035
Scheme 36. General strategy of post-functionalization.
Scheme 36. General strategy of post-functionalization.
Molecules 29 02442 sch036
Scheme 37. General strategy of pre-functionalization.
Scheme 37. General strategy of pre-functionalization.
Molecules 29 02442 sch037
Table 1. Molecules obtained via the post-functionalization strategy.
Table 1. Molecules obtained via the post-functionalization strategy.
EntryR4R5R6R7Strategy
1HHBrHBromination [71,72,101,102]
2HHtBuHFriedel–Crafts [71]
3HHCyHBromination/Coupling [71,72,101]
4HHNO2HNitration [103]
5HHAr and derivativesHBromination/Coupling [72,101,102]
6HHiPrHBromination/Coupling [72,73]
7HAdAdHBromination/Coupling [72]
Table 2. Molecules obtained via the pre-functionalization strategy.
Table 2. Molecules obtained via the pre-functionalization strategy.
EntryR4R5R6R7Strategy
1HHHMeStrategy A—Ritter [104,105,106]
Strategy C—Reduction [86]
2HHHiPrStrategy A—Ritter [105,106,107]
Strategy C—Reduction [86]
3HHFHStrategy A—Ritter [108]
4HHBrHStrategy A—Ritter [109]
5HHOMeHStrategy A—Ritter [110]
Strategy B—Mitsunobu [111]
6HHNO2HStrategy A—Ritter [112]
7HClHHStrategy A—Ritter [113]
8HFHHStrategy A—Ritter [108]
9HBrHHStrategy C—Reduction [86,119]
10MeHHHStrategy A—Ritter [106]
Strategy C—Reduction [86]
11FHHHStrategy C—Reduction [114]
12HBrBrHStrategy C—Reduction [86]
13HΠ-extended (biaryl)HStrategy A—Ritter [115,116]
Strategy B—Mitsunobu [117]
14HHΠ-extended (biaryl)Strategy A—Ritter [118]
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

Mendas, I.; Gastaldi, S.; Suppo, J.-S. Strategies for Accessing cis-1-Amino-2-Indanol. Molecules 2024, 29, 2442. https://doi.org/10.3390/molecules29112442

AMA Style

Mendas I, Gastaldi S, Suppo J-S. Strategies for Accessing cis-1-Amino-2-Indanol. Molecules. 2024; 29(11):2442. https://doi.org/10.3390/molecules29112442

Chicago/Turabian Style

Mendas, Inès, Stéphane Gastaldi, and Jean-Simon Suppo. 2024. "Strategies for Accessing cis-1-Amino-2-Indanol" Molecules 29, no. 11: 2442. https://doi.org/10.3390/molecules29112442

Article Metrics

Back to TopTop