Lukas
Lückemeier†
,
Marco
Pierau†
and
Frank
Glorius
*
Universität Münster, Organisch-Chemisches Institut, Corrensstraße 36, 48149 Münster, Germany. E-mail: glorius@uni-muenster.de
First published on 10th July 2023
(Hetero)aromatic compounds are vastly available and easy to functionalise building blocks in the chemical industry. Asymmetric arene hydrogenation enables direct access to complex three-dimensional scaffolds with (multiple) defined stereocentres in a single catalytic process and, by this, the rapid installation of molecular complexity. The potential usage of hydrogen from renewable sources and perfect atom economy bears the potential for sustainable and broadly applicable transformations to valuable products. The aim of this review is to present the state-of-the-art in transition-metal catalysed asymmetric hydrogenation of (hetero)arenes, to highlight recent advances and important trends and to provide a broad overview for the reader.
The saturated aromatic analogues are less prominently represented, however they are at least as important for i.e. drug discovery. Beneficial correlations have been shown between important structural elements, such as high Fsp3, increased molecular complexity or aliphatic ring count, and the clinical success of drug candidates.7–9 In contrast to that, a high aromatic ring count can be associated with certain parameters like poor solubility in aqueous media, high lipophilicity or decreased protein binding.10,11 For this reason, saturated (hetero)cyclic scaffolds are featured in a vast number of top-selling small molecule drugs, frequently with precise stereochemistry.12 However, the synthesis of aliphatic rings can be challenging and step-intensive. Consequently, it has been of high interest in recent years to develop efficient pathways to access these scaffolds in an efficient manner.13–15
Arguably the most sustainable and powerful way to synthesise stereochemically well-defined aliphatic (hetero)cycles is the asymmetric hydrogenation of (hetero)arenes. Starting from broadly accessible and easily modifiable (hetero)arenes, multiple stereocentres can be set in a single synthetic operation giving rise to complex and enantioenriched (hetero)cyclic scaffolds (Fig. 1). Since this catalytic transformation only adds molecular hydrogen to a molecule, the atom economy is perfect rendering hydrogenation as an intrinsically green reaction.16 Further features such as the use of renewable feedstocks and the reduction in waste and step-count emphasise the environmentally benign character. Taking advantage of this strategic value, asymmetric (hetero)arene hydrogenation is employed in numerous industrial processes and total syntheses. Beautiful recent examples are the total syntheses of (−)-jorunnamycin A and (−)-jorumycin 1 by the Stoltz group17 and the multikilogram synthesis of a promising diabetes drug candidate 2 by chemists of Boehringer Ingelheim (Fig. 1).18 Both routes utilise an asymmetric heteroarene hydrogenation as a key step to construct the cores of the complex target structures. This tutorial aims to convey the reader important methods and principles of asymmetric transition metal catalysed arene hydrogenation, highlight historically significant and recently developed catalyst systems and give an outlook into the future of the field. Notably, asymmetric organocatalytic hydrogenation of arenes is also an important emerging field. As this tutorial cannot be comprehensive, recent reviews have given a detailed overview.13,14,19,20
Fig. 1 Hydrogenation as an environmentally benign method to provide direct access to diverse saturated carbo- and heterocyclic motifs. |
Fig. 2 Major challenges in asymmetric arene hydrogenation: high kinetic activation barrier, stereoselectivity and chemoselectivity. |
Three main strategies have been developed to overcome these obstacles: catalyst activation, substrate activation and relay catalysis. Catalysts can be activated by introducing additives that tune the steric and/or electronic environment of the catalyst. Primarily, the development of new ligands influenced the success of enantioselective arene hydrogenation.13,14,25–28 Moreover, it was often observed that the addition of iodine drastically increased the activity and enantioselectivity of metal complexes in hydrogenation reaction. It is proposed that iodine prevents the formation of inactive dinuclear metal complexes, thus increasing the catalytic rate.29–31 Substrates can be activated by installing protecting groups to the heteroatoms of the arene that could serve as a directing group for the catalyst and/or protect the catalyst from being poisoned.32–35 Also protonation of N-heteroarenes is a common strategy to lower the aromaticity of substrates.36 In relay catalysis, the aromaticity of a substrate is broken by an achiral catalyst, that however hydrogenates the substrate only partially. Then, a chiral catalyst sets the stereocentre to yield the final product.37
Various catalyst systems have been developed since the first reports in the late 90s. Most of these systems follow two major approaches to set the stereocentres, either homogeneous enantioselective or asymmetric heterogeneous hydrogenation. The following sections will give a deeper insight into both areas, discussing the advantages and disadvantages of these methods.
One requirement for these catalysts is to remain stable under reductive reaction conditions. Depending on the external conditions, homogeneous metal complexes can form heterogeneous metal nanoparticles or clusters that can lead to a competing racemic hydrogenation reaction.44–46 For stereoinduction, the chiral environment at the metal centre needs to distinguish between two enantiotopic sides of the flat (hetero)arenes. The selectivity is often highly dependent on the solvent, temperature, additives and counterions present in the reaction medium. So far, many enantioselective hydrogenations for different (hetero)arenes including nitrogen-, oxygen-, sulphur-containing and carbocyclic arenes were published.
Most enantioselective arene hydrogenations target bicyclic heteroarenes and partially hydrogenate one ring. The aromatic stabilisation of bicyclic arenes is in general lower compared to two monocyclic arenes. In substrates with one heterocyclic ring and an annulated benzene, the latter one preferably remains intact because of the high stabilisation energy of benzene.
Fig. 4 The alkaloid (–)-Galipine (left) containing a tetrahydroquinoline was accessed via enantioselective hydrogenation by Zhou and co-workers.51 The two ligands enabling high selectivities in asymmetric quinoline hydrogenation by Zhou and co-workers are shown on the right.49,50 |
A remarkable enantiodivergent hydrogenation of 2-aryl substituted quinolines 4 was recently published by Dong, Zhang and co-workers (Scheme 1).52 An iridium catalyst with a thiourea based N-Me-Zhaophos ligand L3 facilitated efficient hydrogenation under acidic conditions.
Scheme 1 Enantiodivergent hydrogenation of 2-arylated quinolines by Dong, Zhang and co-workers.52 |
Simple alteration of the solvent enabled access to both tetrahydroquinoline enantiomers 5 with only one stereoisomer of the catalyst. Highly polar solvents, especially protic solvents, gave the (S)-enantiomer with a TON of up to 1680 while less polar solvents switched the selectivity to the (R)-stereoisomer with a TON of up to 680. Deuterium labelling experiments indicate a difference in mechanism, that needs further investigation. Functionalised phenyl groups including halides, trifluoromethyl and dimethylamine as well as thiophene were tolerated well. 2,3-Disubstituted derivatives with an additional alkyl group yielded the cis-product 6 in high diastereoselectivity under slightly adjusted reaction conditions.
An arene hydrogenation is most of the times a multistep reaction involving different catalytic operations. A typical mechanism for the hydrogenation of quinolines is shown in Scheme 2. After substrate activation a 1,4-hydride addition gives the enamine 9 which can tautomerise to the imine 10. 1,2-Hydride addition to the activated iminium ion under stereocontrol of the catalyst yields the chiral tetrahydroquinoline 12 (THQ).52
Scheme 2 Outer sphere mechanism of the enantioselective hydrogenation of 2-substituted pyridines. After protonation the quinolinium cation undergoes a 1,4-hydride addition forming an enamine intermediate. Acid catalysed tautomerisation and subsequent 1,2-hydride addition gives the chiral tetrahydroquinoline cation.52 |
Sun and co-workers prepared a water stable iridium complex C1 with a simple N,N-ligand containing a benzimidazole scaffold for the transfer hydrogenation of 2-substituted quinolines 13 in an aqueous reaction medium under mild conditions (Scheme 3).53 Formic acid serves a dual role as hydrogen donor and to protonate C1 to increase its solubility. This catalytic hydrogenation appears to be appealing in particular because of its high TON of up to 33000 and therefore low amounts of catalyst are required. Although 2-methylated quinolines could be hydrogenated in good ees, the enantioselective outcome proofed to be sensitive when switching to a different 2-substituent.
Scheme 3 Transfer hydrogenation of 2-substituted quinolines with a water-soluble chiral iridium complex by Sun and co-workers.53 |
Chiral ruthenium diamine complexes are a privileged motive in asymmetric arene hydrogenation. A widely used example is the Ru-DPEN system introduced by Fan.54 More recently Fan and co-workers demonstrated the versatility of this catalyst class by using a chiral 1,2-diaminocyclohexane derived ligand for the efficient preparation of a new chiral terpyridine type class of N,N,N-ligands (Scheme 4).55 The tridentate Lewis basic substrates and resulting products make this transformation particularly challenging since such motives can strongly coordinate to the metal centre resulting in deactivation of the catalyst. The 2,6-bis(tetrahydroquinolin-2-yl)pyridines 16 (PyBTHQ) were obtained in high yields and excellent ees. These newly accessed PyBTHQ were successfully employed as ligands in a copper catalysed enantioselective Friedel–Crafts reaction.
Scheme 4 Synthesis of chiral terpyridine N,N,N-ligands via an enantioselective hydrogenation with a chiral Ru-diamine catalyst by Fan and co-workers.55 |
The vast majority of asymmetric arene hydrogenations requires the use of precious and rare metals. Thus, switching to readily available and inexpensive 3d-metals is a more sustainable and highly desired method, but yet scarcely presented. A very remarkable example for its utility was recently published by Lan, Liu and co-workers who employed a manganese catalyst for the enantioselective hydrogenation of quinolines 17 (Scheme 5).56 By strategic alteration of the N,N,P-pincer ligand L4 they investigated the role of the benzimidazole scaffold and nitrogen substitution pattern. Free N–H groups in the linker and the imidazole proofed to be crucial for the reactivity and selectivity of the complex. DFT calculations of the transition state in the stereo determining step reveal the attractive π–π-interaction of the substrate with the benzimidazole favours one enantiomer. This synthetic method features an outstanding chemoselectivity for reductively labile groups. Tri- and disubstituted alkenes and alkynes are tolerated well. Low catalyst loadings were demonstrated on a gram scale with a TON up to 3840. An interesting application is the enantio- and diastereoselective hydrogenation of bis(quinoline-2-yl)methanes 21 followed by a simple condensation to access enantioenriched six membered NHC-prescursors 22.
Scheme 5 Manganese catalysed enantioselective hydrogenation of 2-substituted quinolines published by Lan, Liu and co-workers.56 |
Scheme 6 Cascade hydrogenation-condensation strategy for the synthesis of chiral indolizidines and quinolizidines involving an enantioselective hydrogenation published by Fan and co-workers.57 |
One-pot cascade procedures are an emerging synthetic strategy.58,59 Instead of conducting a hydrogenation with a certain arene directly, the substrate is formed in situ by a sequence of reactions and then hydrogenated. This strategy is especially useful when the starting materials to form the arene are cheap and easily available. A tandem strategy involving two catalysts was used by Yu, Fan and co-workers to prepare chiral 1H-isochromenes 30 (Scheme 7).60Ortho-(alkynyl)arylketones 29 were used to generate reactive isochromenylium cations 31 by copper catalysis in situ, which were hydrogenated with a Ru-DPEN catalyst C3. It was further shown that the stereocentre of the 1H-isochromenes 30 can be used to induce stereoselectivity in a subsequent heterogenous hydrogenation to access the cis-configurated 1,3-substituted isochromanes.
Scheme 7 In situ formation of aromatic isochromenylium cations followed by enantioselective hydrogenation for the synthesis of chiral 1H-isochromenes by Yu, Fan and co-workers.60 |
Scheme 8 (A) Synthesis of chiral cis-configurated 1,3-disubstituted tetrahydroisoquinolines via enantioselective hydrogenation published by Stoltz and co-workers.62 (B) Application as key step in the total syntheses of the natural products (−)-jorunnamycin A and (−)-jorumycin by Stoltz and co-workers.17 |
An impressive example for the synthetic power of asymmetric arene hydrogenation was shown by Stoltz and co-workers in the total syntheses of the complex natural products (−)-jorunnamycin A and (−)-jorumycin (Scheme 8B).17 In a key step of their route the bis-isoquinoline 37 was hydrogenated creating four new stereocentres under excellent control of diastereo- and enantioselectivity in a single step and setting the stage for the following lactam condensation to form the highly substituted pentacyclic scaffold 40. Similar to the above described method an iridium catalyst with a modified JosiPhos ligand L6 was used. This non-biomimetic approach enabled the synthesis of more electron rich partially deoxygenated derivatives for the strategic investigation of the location of active sites for the cytotoxicity.
Arene hydrogenation reactions are mainly cis-selective because the hydrogen is transferred from one catalytic face. trans-Selective arene hydrogenations are very rare and until recently no example for an asymmetric version was known.63 Mechanistically a π-facial exchange in between the stepwise hydrogenation is required.
Stoltz and co-workers recently reported that the diastereoselectivity for the enantioselective hydrogenation of 1,3-disubstituted isoquinolines 41 can be switched to a trans-selective method when the coordinating solvent THF was exchanged for the non-coordinating solvent DCE (Scheme 9).63 The right choice of additive turned out to be crucial for the selectivity. TBABr and TBACl provided trans-selectivity whereas TBAI led to a favoured formation of the cis-diastereomer. Although the trans-selectivity for the chloride donor was slightly better the enantioselectivity was diminished compared to the bromide donor. This method requires a 1-hydroxymethyl group as coordinating group limiting the scope of accessible products. Other coordinating groups could not switch the selectivity in favour of the trans-product. In general, the selectivity for the trans-diastereomer is certainly lower than in the cis-selective reaction, still the enantioselectivity is high and the ability to access both chiral diastereomers with the same catalyst by simply switching the reaction medium and additive is synthetically a very useful methodology.
Scheme 9 trans-Selective directed enantioselective hydrogenation of 1,3-disubstituted isoquinolines with subsequent formation of an oxazolidine-2-one by Stoltz and co-workers.63 |
Scheme 10 Iridium catalysed enantioselective hydrogenation of 2- and 3-substituted indoles and benzofurans by Han, Ding and co-workers.64 |
Designing a catalytic system which is capable of performing asymmetric hydrogenations of various aromatic scaffolds in high stereoselectivity while maintaining a broad functional group tolerance is a challenging task. The ruthenium-SINpEt catalyst reported by Glorius and co-workers proved to be exceptionally versatile for the asymmetric hydrogenation of multiple heterocycles (Fig. 5).65 Different frameworks including five membered monocycles, 6,5-bicycles and 6,6-bicycles and different heteroatoms namely oxygen, nitrogen and even sulphur gave reliably good selectivity and provided efficient hydrogenation. The heteroarenes can be electronically rich or poor, with a single or multiple heteroatom(s) in various positions including nitrogen bridged scaffolds.66–71 The range of substrates is presented in more detail in a previous review.65
Fig. 5 Overview of the scope of diverse heterocycles for the efficient enantioselective hydrogenation with the privileged Ru-SINpEt catalyst by Glorius and co-workers.65 |
In most enantioselective hydrogenation reactions of polycyclic heteroarenes the ring containing the heteroatom(s) is the one that is hydrogenated more readily. Carbocyclic arenes usually have higher aromaticity and are less polarised compared to heteroarenes.48 However, with the Ru-SINpEt system Glorius and co-workers for the first time were able to selectively hydrogenate the carbocylic benzene ring of quinoxalines enantioselectively with retention of the pyrazine motif.72
Fig. 6 Overview of the suitable (hetero)arenes for the efficient enantioselective hydrogenation with the privileged Ru-PhTRAP catalyst established by Kuwano and co-workers.47,73–76 |
Recently Zhou and co-workers reported the asymmetric partial hydrogenation of different polycyclic all-carbon arenes to access axially chiral product motifs with a rhodium-phosphine catalytic system (Scheme 11).77 A broad scope of 9-phenyl substituted anthracenes 51 bearing a protected amine or ether directing group were desymmetrised and also two examples without a directing group were shown (Scheme 11A).
Scheme 11 Asymmetric hydrogenation of carbocylic motifs to access axial-chiral and central-chiral products by Zhou and co-workers.77 (A) Desymmetrisation of 9-phenyl substituted anthracenes. (B) Kinetic resolution of phenyl-substituted naphtalenes. (C) Enantioselective hydrogenation of 9-subtituted phenanthracenes. (D) Structures of the ligands used for these reported hydrogenations. |
Crucial for the enantioselectivity is the strongly electron donating and sterically very demanding WingPhos ligand L8, the choice of counterion of the rhodium salt and the solvent. A weakly coordinating SbF6− anion and non-coordinating dichloromethane proved to be beneficial. The utility of this desymmetrisation was shown for the synthesis of axial chiral phosphine ligands by hydrogenating the corresponding phosphine oxide substrates followed by reduction to the phosphines. This procedure was successfully applied for the kinetic resolution of racemates of phenyl-substituted naphthalenes 53 (Scheme 11B). The matching atropisomer was partially hydrogenated at the naphthalene because of its lower aromaticity per ring. Both the enantioenriched mismatched starting material 54 and the partially hydrogenated product 55 were obtained in good selectivity. Moreover, phenanthracenes 56 bearing a coordinating amide in 9-position were enantioselectively hydrogenated at the central ring to yield central-chiral partially saturated products 57 in high selectivity and good yields (Scheme 11C). For this hydrogenation a JosiPhos ligand L9 was found to be the best and the reaction conditions were adjusted. Its noteworthy to comment that although also 9,10-disubstituted motifs gave high ees, the necessity of the directing group and a scope which is limited to non- or methyl-substituted external rings limits its applicability.77 In a previously reported similar method by Zhou and co-workers 9-acetamide substituted phenanthracenes could also be hydrogenated in high selectivity.78
An intriguing new catalyst design for the enantioselective and diastereoselective complete hydrogenation of quinolines and naphthalenes was recently published by Chirik and co-workers (Scheme 12).79 Earth-abundant molybdenum was chosen as more sustainable transition metal and strategic evaluation of ligands led to an oxazoline imino(pyridine) motif. A bulky 4-tert-butyl group on the pyridine ring turned out to be crucial for high enantioselectivity, although facing away from the metal centre. The added steric repulsion might prevent hydrogenation of the pyridine and finally erosion of the catalyst leading to racemic background reactions lowering the ee.
Scheme 12 Molybdenum catalysed enantioselective full hydrogenation of quinolines and naphthalenes to the corresponding decahydroquinolines and decalines by Chirik and co-workers.79 |
Nitrogen-containing and carbocyclic fused bicycles were investigated as substrates. Similar to quinolines the hydrogenation of naphthalenes usually yields tetralins because a highly stabilised benzene ring is conserved.48 2,6-Dialkylated naphthalenes provided the fully reduced cis-configurated decalines in high enantioselecitivity. For the 1,7-dimethylated naphthalene 60 the trans-product 61 was observed as major diastereomer. Different substitution patterns including 1,8-disubstituted and monosubstituted naphthalenes resulted in reduced selectivity, although the monosubstituted 2-methyl derivative gave a moderate ee. Interestingly, while 2,6-disubstituted quinolines were hydrogenated to the corresponding decahydroquinolines, monosubstituted quinolines, isoquinolines, quinoxalines and 2,8-disubstituted quinolines only yielded mixtures of partially saturated products. The chemoselectivity for whether the carbocyclic ring or the nitrogen containing ring was hydrogenated was found out to be highly dependent on the steric environment and location of the substituent. The enantioselectivity on the other side, was also high for 3- and 4-substituted quinolines. Mechanistic investigations support a mode of enantioinduction based on buried volume without a coordination of the nitrogen to the catalyst. The development of a molybdenum-based enantioselective catalyst for arene hydrogenation with unique reactivity marks a progress towards sustainability. However, relatively high catalyst loadings, low yields for heteroarenes and a scope restricted to alkyl substituents limit the applicability of this new system and demand further studies.79
Fig. 7 Top: Overview of the susceptible pyridine motifs for enantioselective hydrogenation.80–90 Middle: Common activation strategies of pyridines for homogenous asymmetric hydrogenation. Bottom: Examples for successfully used ligands in enantioselective pyridine hydrogenation. |
Chiral piperazines are common scaffolds in pharmaceutically active compounds. From a strategic perspective the most straight-forward way to access them is a single-step enantioselective hydrogenation. Still, there are only few reports using this pathway.92 Compared to piperidines the piperazines contain an additional amine that can lead to catalyst poisoning. Zhou and co-workers developed methods to hydrogenate 3-monosubstituted, 2,3- and 3,5-disubstituted pyrazines with iridium-phosphine catalysts in good selectivities after activation by benzylation to the corresponding pyrazinium salt.93
The choice of counterion turned out to be crucial for the selectivity and depending on the substitution pattern different ligands and solvent systems were used. The authors demonstrated the utility of their transformation by synthesising the NK1 receptor antagonist Vestipitant 77 from the benzylated piperazinium in three steps. Zhou, Shi and co-workers recently reported the enantioselective hydrogenation of 5,6-arylated pyrazin-2-ols 72 using a palladium-TolBINAP catalyst.94 The enantioselectivity is high but the method is mainly limited to substrates with identical 5- and 6-substituents.
Another important class of N-heterocycles are undoubtedly pyrimidines. Besides the above-mentioned challenges of overcoming aromatic stabilisation and formation of strongly coordinating products, the fully hydrogenated 1,3-diazinanes are rather sensitive to hydrolysis because of the cyclic aminal moiety. Partial enantioselective hydrogenation to the 1,4,5,6-tetrahydropyrimidines is feasible when a stabilising 2-substituent is introduced preventing the formation of an aminal. Kuwano and co-workers converted 2-arylated pyrimidines 69 to the corresponding cyclic amidines 73 with an iridium-JosiPhos catalyst.95 Substrate activation with iodine and ytterbium triflate promoted high enantioselectivity for a broad scope of 4-alkylated or arylated substrates 73. 2-Hydroxpyrimidines 70 are suitable substrates for the synthesis of chiral cyclic ureas 74via hydrogenation as shown by Shi, Zhou and co-workers.96,97 The 2-hydroxypyrimidines 70 have a lower aromaticity and exist in an equilibrium with the tautomeric pyrimidone. Activation with organic acids or TCCA enabled effective hydrogenation of mono- and multi-substituted 2-hydroxypyrimidines in overall good selectivities (Fig. 8).
Fig. 8 Top: Overview of pyrimidines and pyrazines successfully employed in an enantioselective hydrogenation reaction.93–97 Middle: Corresponding product motifs after hydrogenation. Bottom: Examples for successfully used ligands in these hydrogenation reactions. |
The field of asymmetric heterogeneous hydrogenation can be divided into three subareas. First, immobilisation of privileged homogeneous catalyst systems on a surface. Therefore, chiral ligands or preformed metal complexes are bound covalently or non-covalently (i.e. ionic interactions) to a support such as silica (Fig. 9A).102,103 Second, hydrogenation over surfaces modified with chiral molecules. By the addition of chiral complexing molecules that can interact with the substrate, a chiral environment on the metal surface is created (Fig. 9B).101 The best known and studied modified surface is the Orito system. His group used cinchona alkaloids to modify a platinum surface for the enantioselective hydrogenation of pyruvates and other α-functionalised ketones.104 Despite their success in other areas, these two methods have produced only moderate to bad ees in the field of asymmetric arene hydrogenation. More promising results can be found in the third subarea, the diastereoselective hydrogenation of chiral molecules. These chiral substrates either have steric repulsive or electronic attractive interactions with the metal surface so that one diastereotopic face of the arene is preferentially hydrogenated (Fig. 9C).105
First attempts of diastereoselective hydrogenation were performed by Lemaire in 1994. Employing menthoxyacetic acid as chiral auxiliary to o-cresol, he was able to achieve a promising but low 10% ee.106 Four years later Besson and co-workers used pyroglutamic acid methyl ester as a more rigid chiral auxiliary in the hydrogenation of o-toluic acid.107 A highly diastereoselective arene hydrogenation resulted, with an impressive de of 95% at 49% conversion. The Glorius group developed a diastereoselective pyridine hydrogenation using the Evans’ auxiliary as chiral unit and Pd(OH)2/C as achiral catalyst (Scheme 14A).108 Crucial for the success of the reaction is the acidic medium which serves three important functions. The protonation of the pyridine facilitates a hydride attack and lowers the kinetic barrier, additionally it negates the Lewis basicity of the free electron pair on the nitrogen, preventing it from poisoning the catalyst. More importantly, it is believed that a hydrogen bond between the protonated nitrogen and the carbonyl group locks the conformation of the auxiliary, assuring the efficient shielding of one diastereotopic face and consequently yielding high ees (ϕ 94% ee). With this method they were able to set up to three stereocentres in a single operation. Another aspect that renders this reaction environmentally benign is the recyclability of the auxiliary. Under the reaction conditions it is cleaved after hydrogenation – serving as a traceless auxiliary – and can be recovered and reused without any loss in enantioselectivity. This protocol found application in the enantioselective synthesis of (−)-isooncinotine.109
Scheme 14 Chiral auxiliary assisted asymmetric hydrogenations of pyridines and pyridinium salts.108,110,112 |
Recently, Glorius and co-workers disclosed an asymmetric interrupted pyridine hydrogenation which builds upon these results (Scheme 14B).110 During the hydrogenation cycle the imine intermediate 86 accumulates and can be intercepted by a nucleophile. By switching the solvent from AcOH to a mixture of THF/H2O they could utilise water as a nucleophile to attack the imine intermediate. The product 88, enantioenriched δ-lactams, could be synthesised in high ees (ϕ 90% ee) and yields (ϕ 72%) and it could be shown that this method can be applied in the synthesis of enantiopure δ-amino acids. Another application was demonstrated in the elegant total synthesis of (−)-senepodine F by Ishikawa and co-workers.111 As one of the key steps, one piperidine moiety was constructed by the asymmetric interrupted pyridine hydrogenation protocol reported by Glorius. The final product was obtained in 13% overall yield over 17 steps.
Another interrupted hydrogenation process making use of a chiral auxiliary is the recently published reductive amination of pyridinium salts reported by Xiao and co-workers (Scheme 14C).112 This protocol describes an asymmetric transfer hydrogenation employing [Cp*RhCl2]2 as achiral catalyst and formic acid as the hydrogen source. The authors discovered that if chiral amine 90 is added to the reaction mixture it is incorporated into the product, giving rise to enantiomerically pure piperidines 91. After intensive mechanistic investigations it is proposed that as a key step the dihydropyridinium intermediate 92 is intercepted by water, which leads to a ring opening and an expulsion of ethylamine. After condensation with chiral amine 90 and ring closure, a final diastereoselective hydrogenation of the imine leads to the enantiomerically pure piperidine 98. Remarkably, this protocol tolerates many reductively labile groups that are usually not stable under normal hydrogenation conditions. Easily reducible groups such as nitro, cyano or carbonyl groups as well as halogens, which are prone to hydrodefunctionalisation, are well tolerated. This underlines the mildness of the reaction by not employing high hydrogen pressure, but only transfer hydrogenation conditions. The products are usually obtained in high diastereoselectivities and yields.
Besides these recent examples, there has been also a few other highly diastereoselective (≥90% ee) hydrogenations of furanes,113,114 pyrroles115 and quinolines116 utilising chiral auxiliaries. Nonetheless, introducing chiral auxiliaries that possess enough conformational rigidity to shield one diastereotopic face efficiently is a challenging task. Additionally, the auxiliaries need to be readily synthesised from the chiral pool and the facile installation, cleavage and recyclability is essential for the protocol to be environmentally benign. These drawbacks limit the area of application for this strategy.
Scheme 15 Investigation of substituent effects on the diastereoselective hydrogenation of 1-substituted indanes.117,118 |
The value of this strategy was demonstrated multiple times in the syntheses of natural products and promising drug candidates. In 1968 Cooke and Fodor published the synthesis of (−)-sedridine 103.119 They took advantage of a hydroxyl group adjacent to the pyridine ring, inducing a high diastereoselectivity. By IR studies they discovered that the hydroxyl group is completely intramolecularly hydrogen bonded to the pyridine nitrogen. The authors propose that this leads to a rigid ring structure in which the methyl group shields one face of the ring, leading to a high d.r. (Scheme 16A).
Scheme 16 Substrate induced asymmetric diastereoselective hydrogenations in total syntheses of natural products and drug candidates.119,120,122,123 |
In the synthesis of Perindopril 107, an inhibitor of Angiotensin Converting Enzyme (ACE), the optically pure octahydroindole core was constructed by a diastereoselective hydrogenation of the corresponding indoline 105, yielding the product 106 with a d.r. = 9:1.120 After converting 106 to the corresponding tert-butyl ester, the optical purity was determined as >98% ee. Further synthetic transformations yielded the desired drug (Scheme 16B).
Tashiromine is an indolizidine alkaloid, a group of natural products that can be extracted from various plants and animals with potent biological activity and pharmacological effects.121 Shortly after the first extraction of natural tashiromine, Branchaud and co-workers synthesised (+)- and (−)-tashiromine to address the stereochemistry question.122 After assembling the tetrahydroindolizine core 108via an enantioselective pyrrole/cobaloxime π-cation cyclisation, a diastereoselective hydrogenation with Rh/Al2O3 yielded the indolizidine 109 with a d.r. = 2:1. Further cleavage of the protecting group and epimerisation yielded (+)-tashiromine 110 (Scheme 16C).
Another interesting synthesis of an indolizidine via a diastereoselective hydrogenation was reported by Daïch and co-workers.123 Starting from the furoindolizidinone 111, a direct hydrogenation of the ketone and the furan resulted in a mixture of the four different diastereomers (113-116), albeit the d.r. of the reaction was highly dependent on the employed catalyst. While Pd/C, Ru/C and RANEY®-Ni produced poor d.r., Rh/Al2O3 gave 113 as almost the sole product (d.r. = 96:0:1:3). The major contributor to the diastereoselectivity is the steric hindrance of the pyrrolidone ring junction which guides the hydrogen to the exo face of the molecule. However, if the ketone was reduced with NaBH4 to the ancillary alcohol 112 prior to the hydrogenation of the furan, the diastereoselective outcome was predominantly determined by haptophilic effects. The alcohol and the furan oxygen coordinate to the catalyst guiding it and consequently the hydrogen to the same face of the ring system. Therefore, 115 is formed as major diastereomer with an excellent d.r. = 93:7 and RANEY®-Ni as the most selective catalyst (Scheme 16D).
Bach and co-workers disclosed highly diastereoselective hydrogenations of 2-oxindoles, 3,4-dihydroquinolones (117) and aromatic 2,5-diketopiperazines 121 in 2020 and 2022 (Scheme 17).124,125 As a catalyst was employed a rhodium complex, bearing a cyclic (amino)(alkyl)carbene (Rh-CAAC, C7). This complex decomposes under hydrogenation conditions into nanoparticles which are active for arene hydrogenation, but tolerate a plethora of reductively labile motifs such as ketones,126 fluorine,127–129 boron130 or silicon.131 Additionally, it displays a high diastereoselectivity towards the all-cis configurated product. Also, Bach and co-workers observed great functional group compatibility (fluorine, boron, silicon, amides, alcohols, amines, ethers, esters), good yields (ϕ 72%) and high levels of diastereoselectivity for the all-cis product 118 (for 2-oxindoles and 3,4-dihydroquinolones ϕ d.r. 89:11, for 2,5-diketopiperazines ϕ d.r. > 99:1). The great facial selectivity can be attributed to the steric repulsive effects of the substituents, shielding one side efficiently. Furthermore, this method was used in the synthesis of enantiomerically pure compounds 119 and 120. Especially, optically pure 2,5-diketopiperazines like 123 are a reoccurring scaffold in various natural and biologically active products.132 This underlines the utility of this method for its application in total synthesis of biologically relevant compounds.
Scheme 17 Asymmetric diastereoselective hydrogenations of (A) 2-oxindoles, 3,4-dihydroquinolones and (B) aromatic 2,5-diketopiperazines with a highly selective Rh-CAAC catalyst.124,125 |
The same Rh-complex was used by Glorius and co-workers in a novel relay catalysis approach to construct enantiomerically pure octahydrobenzofurans 125.133 In a one-pot reaction, first the furan ring was reduced by chiral Ru-SINpEt C5 to set the stereocentre. For this step it is crucial that it proceeds at low temperature to (a) achieve a high ee and (b) prevent a racemic background reaction. At 25 °C the Rh-CAAC complex C7 does not form nanoparticles, which would hydrogenate the benzofuran completely to give a racemic mixture. After 3 h the temperature was elevated to 60 °C and the pressure to 70 bar to induce the formation of nanoparticles and the diastereoselective hydrogenation of the remaining benzene ring. The facial discrimination of the second step was obtained by the pre-set stereocentre. With this relay catalysis the authors showed a broad substrate scope, in general high yields (ϕ 85%) and diastereoselectivities (ϕ 93:7 d.r.), thus combining the great chemo- and diastereoselectivity of the Rh-CAAC complex C7 with the high enantioselectivity of the Ru-SINpEt C5. Additionally, it was possible to set up to six stereocentres in a single synthetic operation (Scheme 18A).
Scheme 18 Hydrogenations employing relay catalysis to first set a stereocentre and then perform a diastereoselective hydrogenation.133,134 |
In a similar manner, Andersson and co-workers used relay catalysis for the asymmetric full saturation of vinyl arenes 128.134 They discovered that [Rh(COD)2]SbF6 as Rh-precursor can aggregate into nanoparticles potent enough to reduce benzene rings. Simultaneously, it can also form an irreversible complex with bisphosphine ligands that are known to hydrogenate enamides with a high degree of reactivity and enantioselectivity. Kinetic studies confirmed that the Rh-bisphosphine-complex fully hydrogenates the enamide in under 20 min and after an induction period of ∼1 h Rh-nanoparticles are formed that then hydrogenate the benzene ring. Essential for the success is a metal to ligand ratio of ∼2:1. In general, good yields, ees and d.r.s were obtained whereas the in situ set stereocentre efficiently shields one face of the benzene ring. Further utilisation of the method has been shown in the synthesis of a saturated analogue of Rasagiline, an anti-Parkinson's therapeutic (Scheme 18B).
Fig. 10 Illustration of future objectives to improve the sustainability and broaden the applicability of asymmetric arene hydrogenation. |
Additionally, highly active catalysts with a high TON and TOF are desired because these are usually expensive and laborious to recycle.139,140 Sophisticated ligand synthesis increases step-count, waste production and cost of the overall catalyst. Simple ligands that can be easily synthesised from the chiral pool are advantageous.141 Most asymmetric arene hydrogenations require catalysts based on rare and expensive metals like iridium, rhodium and palladium.13 Much more desirable and environmentally benign are catalytic systems based on cheap and highly available earth-abundant metals including 3d-metals.142 However, often these complexes are sensitive to air and not reactive enough to promote efficient arene hydrogenation.143 The investigation of new ligand designs, especially pincer ligands for 3d-metals, is crucial.144 Only very recently, Lan, Liu and co-workers reported their pioneering work of the first manganese catalysed enantioselective arene hydrogenation.56
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2023 |