fbpx
Wikipedia

Asymmetric hydrogenation

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information (what chemists refer to as chirality) to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being collectively awarded one half of the 2001 Nobel Prize in Chemistry.[1]

History edit

In 1956 a heterogeneous catalyst made of palladium deposited on silk was shown to effect asymmetric hydrogenation.[2] Later, in 1968, the groups of William Knowles and Leopold Horner independently published the examples of asymmetric hydrogenation using a homogeneous catalysts. While exhibiting only modest enantiomeric excesses, these early reactions demonstrated feasibility. By 1972, enantiomeric excess of 90% was achieved, and the first industrial synthesis of the Parkinson's drug L-DOPA commenced using this technology.[3][4]

 
L-DOPA

The field of asymmetric hydrogenation continued to experience a number of notable advances. Henri Kagan developed DIOP, an easily prepared C2-symmetric diphosphine that gave high ee's in certain reactions. Ryōji Noyori introduced the ruthenium-based catalysts for the asymmetric hydrogenated polar substrates, such as ketones and aldehydes. Robert H. Crabtree demonstrated the ability for Iridium compounds to catalyse asymmetric hydrogenation reactions in 1979 with the invention of Crabtree's catalyst.[5] In the early 1990's, the introduction of P,N ligands by several groups independently then further expanded the scope of the C2-symmetric ligands, although they are not fundamentally superior to chiral ligands lacking rotational symmetry.[6]

Today, asymmetric hydrogenation is a routine methodology in laboratory and industrial scale organic chemistry. The importance of asymmetric hydrogenation was recognized by the 2001 Nobel Prize in Chemistry awarded to William Standish Knowles and Ryōji Noyori.

Mechanism edit

Asymmetric hydrogenations operate by conventional mechanisms invoked for other hydrogenations. This includes inner sphere mechanisms, outer sphere mechanisms and the σ-bond metathesis mechanisms.[7] The type of mechanism employed by a catalyst is largely dependent on the ligands used in a system, which in turn leads to certain catalyst-substrate affinities.

Inner sphere mechanisms edit

The so-called inner sphere mechanism entails coordination of the alkene to the metal center.[8] Other characteristics of this mechanism include a tendency for a homolytic splitting of dihydrogen when more electron-rich, low-valent metals are present while electron-poor, high valent metals normally exhibit a heterolytic cleavage of dihydrogen assisted by a base.[9]

The diagram below depicts purposed mechanisms for catalytic hydrogenation with rhodium complexes which are inner sphere mechanisms. In the unsaturated mechanism, the chiral product formed will have the opposite mode compared to the catalyst used. While the thermodynamically favoured complex between the catalyst and the substrate is unable to undergo hydrogenation, the unstable, unfavoured complex undergoes hydrogenation rapidly.[10] The dihydride mechanism on the other hand sees the complex initially hydrogenated to the dihydride form. This subsequently allows for the coordination of the double bond on the non-hindered side. Through insertion and reductive elimination, the product's chirality matches that of the ligand.[11]

 
Proposed mechanisms for asymmetric hydrogenation

The preference for producing one enantiomer instead of another in these reactions is often explained in terms of steric interactions between the ligand and the prochiral substrate. Consideration of these interactions has led to the development of quadrant diagrams where "blocked" areas are denoted with a shaded box, while "open" areas are left unfilled. In the modeled reaction, large groups on an incoming olefin will tend to orient to fill the open areas of the diagram, while smaller groups will be directed to the blocked areas and hydrogen delivery will then occur to the back face of the olefin, fixing the stereochemistry. Note that only part of the chiral phosphine ligand is shown for the sake of clarity.

 
Quadrant model for asymmetric hydrogenation

Outer sphere mechanisms edit

Some catalysts operate by "outer sphere mechanisms" such that the substrate never bonds directly to the metal but rather interacts with its ligands, which is often a metal hydride and a protic hydrogen on a ligand. As such, in most cases dihydrogen is split heterolytically, with the metal acting as a Lewis acid and either an external or internal base "deprotonating" the hydride.[7]

 
Proposed intermediates in the outer sphere mechanisms for: heterolytic hydrogenation of the catalyst (left) and hydride transfer from the catalyst to the substate (right)

For an example of this mechanism we can consider the BINAP-Ru-diamine system. The dihalide form of the catalyst is converted to the catalysts by reaction of H2 in the presence of base:[12]

RuCl2(BINAP)(diamine) + 2 KOBu-t + 2 H2 → RuH2(BINAP)(diamine) + 2 KCl + 2 HOBu-t

The resulting catalysts have three kinds of ligands:

  • hydrides, which transfer to the unsaturated substrate
  • diamines, which interact with substrate and with base activator by the second coordination sphere
  • diphosphine, which confers asymmetry.

The "Noyori-class" of catalysts are often referred to as bifunctional catalysts to emphasize the fact that both the metal and the (amine) ligand are functional.[13]

In the hydrogenation of C=O containing substates, the mechanism was long assumed to operate by a six membered pericyclic transition state/intermediate whereby the hydrido ruthenium hydride center (HRu-NH) interacts with the carbonyl substrate R2C=O.[14] More recent DFT and experimental studies have shown that this model is largely incorrect. Instead, the amine backbone interacts strongly with the base activator, which often is used in large excess.[12] However in both cases, the substate does not bond directly with the metal centre, thus making it a great example of an outer sphere mechanism.

Metals edit

Practical AH employ platinum metal-based catalysts.[15][16][17]

Base metals edit

Iron is a popular research target for many catalytic processes, owing largely to its low cost and low toxicity relative to other transition metals.[18] Asymmetric hydrogenation methods using iron have been realized, although in terms of rates and selectivity, they are inferior to catalysts based on precious metals.[19] In some cases, structurally ill-defined nanoparticles have proven to be the active species in situ and the modest selectivity observed may result from their uncontrolled geometries.[20]

Ligand classes edit

Phosphine ligands edit

Chiral phosphine ligands, especially C2-symmetric ligands, are the source of chirality in most asymmetric hydrogenation catalysts. Of these the BINAP ligand is well-known, as a result of its Nobel Prize-winning application in the Noyori asymmetric hydrogenation.[3]

Chiral phosphine ligands can be generally classified as mono- or bidentate. They can be further classified according to the location of the stereogenic centre – phosphorus vs the organic substituents. Ligands with a C2 symmetry element have been particularly popular, in part because the presence of such an element reduces the possible binding conformations of a substrate to a metal-ligand complex dramatically (often resulting in exceptional enantioselectivity).[21]

Monodentate phosphines edit

Monophosphine-type ligands were among the first to appear in asymmetric hydrogenation, e.g., the ligand CAMP.[22] Continued research into these types of ligands has explored both P-alkyl and P-heteroatom bonded ligands, with P-heteroatom ligands like the phosphites and phosphoramidites generally achieving more impressive results.[23] Structural classes of ligands that have been successful include those based on the binapthyl structure of MonoPHOS [24] or the spiro ring system of SiPHOS.[25] Notably, these monodentate ligands can be used in combination with each other to achieve a synergistic improvement in enantioselectivity;[26] something that is not possible with the diphosphine ligands.[23]

 
A ferrocene derivative
 
The CAMP ligand
 
A BINOL derivative

Chiral diphosphine ligands edit

The diphosphine ligands have received considerably more attention than the monophosphines and, perhaps as a consequence, have a much longer list of achievement. This class includes the first ligand to achieve high selectivity (DIOP), the first ligand to be used in industrial asymmetric synthesis (DIPAMP[27][28][4]) and what is likely the best known chiral ligand (BINAP).[3] Chiral diphosphine ligands are now ubiquitous in asymmetric hydrogenation.

 


Historically important diphosphine ligands

P,N and P,O ligands edit

 
Generic PHOX ligand architecture
 
Effective ligand for various asymmetric-hydrogenation processes

The use of P,N ligands in asymmetric hydrogenation can be traced to the C2 symmetric bisoxazoline ligand.[29] However, these symmetric ligands were soon superseded by monooxazoline ligands whose lack of C2 symmetry has in no way limits their efficacy in asymmetric catalysis.[30] Such ligands generally consist of an achiral nitrogen-containing heterocycle that is functionalized with a pendant phosphorus-containing arm, although both the exact nature of the heterocycle and the chemical environment phosphorus center has varied widely. No single structure has emerged as consistently effective with a broad range of substrates, although certain privileged structures (like the phosphine-oxazoline or PHOX architecture) have been established.[31][30][32] Moreover, within a narrowly defined substrate class the performance of metallic complexes with chiral P,N ligands can closely approach perfect conversion and selectivity in systems otherwise very difficult to target.[33] Certain complexes derived from chelating P-O ligands have shown promising results in the hydrogenation of α,β-unsaturated ketones and esters.[34]

NHC ligands edit

 
Catalyst developed by Burgess for asymmetric hydrogenation

Simple N-heterocyclic carbene (NHC)-based ligands have proven impractical for asymmetrical hydrogenation.

Some C,N ligands combine an NHC with a chiral oxazoline to give a chelating ligand.[35][36] NHC-based ligands of the first type have been generated as large libraries from the reaction of smaller libraries of individual NHCs and oxazolines.[35][36] NHC-based catalysts featuring a bulky seven-membered metallocycle on iridium have been applied to the catalytic hydrogenation of unfunctionalized olefins[35] and vinyl ether alcohols with conversions and ee's in the high 80s or 90s.[37] The same system has been applied to the synthesis of a number of aldol,[38] vicinal dimethyl[39] and deoxypolyketide[40] motifs, and to the deoxypolyketides themselves.[41]

C2-symmetric NHCs have shown themselves to be highly useful ligands for the asymmetric hydrogenation.[42]

Acyclic substrates edit

Substrates can be classified according to their polarity. Nonpolar substrates are dominated by alkenes. Polar substrates include ketones, enamines ketimines.

Nonpolar substrates edit

[43]

Alkenes that are particularly amenable to asymmetric hydrogenation often feature a polar functional group adjacent to the site to be hydrogenated. In the absence of this functional group, catalysis often results in low ee's. For some unfunctionalized olefins, iridium with P,N-based ligands) have proven effective, however. Alkene substrates are often classified according to their substituents, e.g., 1,1-disubstituted, 1,2-diaryl trisubstituted, 1,1,2-trialkyl and tetrasubstituted olefins.[44][45] and even within these classes variations may exist that make different solutions optimal.[46]

 


Example of asymmetric hydrogenation of unfunctionalized olefins


 
Chiral phosphoramidite and phosphonite ligands used in the asymmetric hydrogenation of enamines

Conversely to the case of olefins, asymmetric hydrogenation of enamines has favoured diphosphine-type ligands; excellent results have been achieved with both iridium- and rhodium-based systems. However, even the best systems often suffer from low ee's and a lack of generality. Certain pyrrolidine-derived enamines of aromatic ketones are amenable to asymmetrically hydrogenation with cationic rhodium(I) phosphonite systems, and I2 and acetic acid system with ee values usually above 90% and potentially as high as 99.9%.[47] A similar system using iridium(I) and a very closely related phosphoramidite ligand is effective for the asymmetric hydrogenation of pyrrolidine-type enamines where the double bond was inside the ring: in other words, of dihydropyrroles.[48] In both cases, the enantioselectivity dropped substantially when the ring size was increased from five to six.

Imines and ketones edit

 
Noyori catalyst for asymmetric hydrogenation of ketones

Ketones and imines are related functional groups, and effective technologies for the asymmetric hydrogenation of each are also closely related. Early examples are Noyori's ruthenium-chiral diphosphine-diamine system.[49][50]

For carbonyl and imine substrates, end-on, η1 coordination can compete with η2 mode. For η1-bound substrates, the hydrogen-accepting carbon is removed from the catalyst and resists hydrogenation.[51]

Iridium/P,N ligand-based systems have been effective for some ketones and imines. For example, a consistent system for benzylic aryl imines uses the P,N ligand SIPHOX in conjunction with iridium(I) in a cationic complex to achieve asymmetric hydrogenation with ee >90%.[52] An efficient catalyst for ketones, (turnover number (TON) up to 4,550,000 and ee up to 99.9%) is an iridium(I) system with a closely related tridentate ligand.[53]

 


Highly effective system for the asymmetric hydrogenation of ketones

The BINAP/diamine-Ru catalyst is effective for the asymmetric reduction of both functionalized and simple ketones,[54] and BINAP/diamine-Ru catalyst can catalyze aromatic, heteroaromatic, and olefinic ketones enantioselectively.[55] Better stereoselectivity is achieved when one substituent is larger than the other (see Flippin-Lodge angle).

 
BINAP/diamine-Ru catalyst scope

Aromatic substrates edit

The asymmetric hydrogenation of aromatic (especially heteroaromatic), substrates is a very active field of ongoing research. Catalysts in this field must contend with a number of complicating factors, including the tendency of highly stable aromatic compounds to resist hydrogenation, the potential coordinating (and therefore catalyst-poisoning) abilities of both substrate and product, and the great diversity in substitution patterns that may be present on any one aromatic ring.[56] Of these substrates the most consistent success has been seen with nitrogen-containing heterocycles, where the aromatic ring is often activated either by protonation or by further functionalization of the nitrogen (generally with an electron-withdrawing protecting group). Such strategies are less applicable to oxygen- and sulfur-containing heterocycles, since they are both less basic and less nucleophilic; this additional difficulty may help to explain why few effective methods exist for their asymmetric hydrogenation.

Quinolines, isoquinolines and quinoxalines edit

Two systems exist for the asymmetric hydrogenation of 2-substituted quinolines with isolated yields generally greater than 80% and ee values generally greater than 90%. The first is an iridium(I)/chiral phosphine/I2 system, first reported by Zhou et al..[57] While the first chiral phosphine used in this system was MeOBiPhep, newer iterations have focused on improving the performance of this ligand. To this end, systems use phosphines (or related ligands) with improved air stability,[58] recyclability,[58] ease of preparation,[59] lower catalyst loading[60][61] and the potential role of achiral phosphine additives.[62] As of October 2012 no mechanism appears to have been proposed, although both the necessity of I2 or a halogen surrogate and the possible role of the heteroaromatic N in assisting reactivity have been documented.[56]

The second is an organocatalytic transfer hydrogenation system based on Hantzsch esters and a chiral Brønsted acid. In this case, the authors envision a mechanism where the isoquinoline is alternately protonated in an activating step, then reduced by conjugate addition of hydride from the Hantzsch ester.[63]

 

Proposed organocatalytic mechanism

Much of the asymmetric hydrogenation chemistry of quinoxalines is closely related to that of the structurally similar quinolines. Effective (and efficient) results can be obtained with an Ir(I)/phophinite/I2 system[64] and a Hantzsh ester-based organocatalytic system,[65] both of which are similar to the systems discussed earlier with regards to quinolines.

Pyridines edit

Pyridines are highly variable substrates for asymmetric reduction (even compared to other heteroaromatics), in that five carbon centers are available for differential substitution on the initial ring. As of October 2012 no method seems to exist that can control all five, although at least one reasonably general method exists.

The most-general method of asymmetric pyridine hydrogenation is actually a heterogeneous method, where asymmetry is generated from a chiral oxazolidinone bound to the C2 position of the pyridine. Hydrogenating such functionalized pyridines over a number of different heterogeneous metal catalysts gave the corresponding piperidine with the substituents at C3, C4, and C5 positions in an all-cis geometry, in high yield and excellent enantioselectivity. The oxazolidinone auxiliary is also conveniently cleaved under the hydrogenation conditions.[66]

 


Asymmetric hydrogenation of pyridines with heterogeneous catalyst

Methods designed specifically for 2-substituted pyridine hydrogenation can involve asymmetric systems developed for related substrates like 2-substituted quinolines and quinoxalines. For example, an iridium(I)\chiral phosphine\I2 system is effective in the asymmetric hydrogenation of activated (alkylated) 2-pyridiniums[67] or certain cyclohexanone-fused pyridines.[68] Similarly, chiral Brønsted acid catalysis with a Hantzsh ester as a hydride source is effective for some 2-alkyl pyridines with additional activating substitution.[69]

Indoles and pyrroles edit

The asymmetric hydrogenation of indoles has been established with N-Boc protection.[70]

 


Method for asymmetric hydrogenation of Boc-protected indoles

A Pd(TFA)2/H8-BINAP system achieves the enantioselective cis-hydrogenation of 2,3- and 2-substituted indoles.[71][72]

 


Sequential alkylation and asymmetric hydrogenation of 2-substituted indoles

Akin to the behavior of indoles, pyrroles can be converted to pyrrolidines by asymmetric hydrogenation.[73]

 


The asymmetric hydrogenation of 2,3,5-substituted N-Boc pyrroles

Oxygen- and sulfur-containing heterocycles edit

The asymmetric hydrogenation of furans and benzofurans is challenging.[74]

 


The asymmetric hydrogenation of furans and benzofurans

Asymmetric hydrogenation of thiophenes and benzothiophenes has been catalyzed by some ruthenium(II) complexes of N-heterocyclic carbenes (NHC). This system appears to possess superb selectivity (ee > 90%) and perfect diastereoselectivity (all cis) if the substrate has a fused (or directly bound) phenyl ring but yields only racemic product in all other tested cases.[75]

 


The asymmetric hydrogenation of thiophenes and benzothiophenes

Heterogeneous catalysis edit

No heterogeneous catalyst has been commercialized for asymmetric hydrogenation.

The first asymmetric hydrogenation focused on palladium deposited on a silk support. Cinchona alkaloids have been used as chiral modifiers for enantioselectivity hydrogenation.[76]

 
Cinchonidine, one of the cinchona alkaloids

An alternative technique and one that allows more control over the structural and electronic properties of active catalytic sites is the immobilization of catalysts that have been developed for homogeneous catalysis on a heterogeneous support. Covalent bonding of the catalyst to a polymer or other solid support is perhaps most common, although immobilization of the catalyst may also be achieved by adsorption onto a surface, ion exchange, or even physical encapsulation. One drawback of this approach is the potential for the proximity of the support to change the behaviour of the catalyst, lowering the enantioselectivity of the reaction. To avoid this, the catalyst is often bound to the support by a long linker though cases are known where the proximity of the support can actually enhance the performance of the catalyst.[76]

The final approach involves the construction of MOFs that incorporate chiral reaction sites from a number of different components, potentially including chiral and achiral organic ligands, structural metal ions, catalytically active metal ions, and/or preassembled catalytically active organometallic cores.[77] One of these involved ruthenium-based catalysts. As little as 0.005 mol% of such catalysts proved sufficient to achieve the asymmetric hydrogenation of aryl ketones, although the usual conditions featured 0.1 mol % of catalyst and resulted in an enantiomeric excess of 90.6–99.2%.[78]

 


The active site of a heterogeneous zirconium phosphonate catalyst for asymmetric hydrogenation

Industrial applications edit

 
(S,S)-Ro 67-8867

Asymmetric hydrogenations are used in the production of several drugs, such as the antibacterial levofloxin, the antibiotic carbapenem, and the antipsychotic agent BMS181100.[15][16][17]

 
Noyori asymmetric hydrogenation

Knowles' research into asymmetric hydrogenation and its application to the production scale synthesis of L-Dopa[4] gave asymmetric hydrogenation a strong start in the industrial world. A 2001 review indicated that asymmetric hydrogenation accounted for 50% of production scale, 90% of pilot scale, and 74% of bench scale catalytic, enantioselective processes in industry, with the caveat that asymmetric catalytic methods in general were not yet widely used.[79]

Asymmetric hydrogenation has replaced kinetic resolution based methods has resulted in substantial improvements in the process's efficiency.[12] can be seen in a number of specific cases where the For example, Roche's Catalysis Group was able to achieve the synthesis of (S,S)-Ro 67-8867 in 53% overall yield, a dramatic increase above the 3.5% that was achieved in the resolution based synthesis.[80] Roche's synthesis of mibefradil was likewise improved by replacing resolution with asymmetric hydrogenation, reducing the step count by three and increasing the yield of a key intermediate to 80% from the original 70%.[81]

 


Asymmetric hydrogenation in the industrial synthesis of mibefradil

Noyori-inspired hydrogenation catalysts have been applied to the commercial synthesis of number of fine chemicals. (R)-1,2-Propandiol, precursor to the antibacterial levofloxacin, can be efficiently synthesized from hydroxyacetone using Noyori asymmetric hydrogenation:[17]

 
levofloxaxin synthesis

Newer routes focus on the hydrogenation of (R)-methyl lactate.[12]

An antibiotic carbapenem is also prepared using Noyori asymmetric hydrogenation via (2S,3R)-methyl 2-(benzamidomethyl)-3-hydroxybutanoate, which is synthesized from racemic methyl 2-(benzamidomethyl)-3-oxobutanoate by dynamic kinetic resolution.

 
carbapenem synthesis

An antipsychotic agent BMS 181100 is synthesized using BINAP/diamine-Ru catalyst.

 
BMS181110 synthesis

References edit

  1. ^ "The Nobel Prize in Chemistry 2001". 2001-10-10.
  2. ^ Akabori, S.; Sakurai, S.; Izumi, Y.; Fujii, Y. (1956). "An Asymmetric Catalyst". Nature. 178 (4528): 323. Bibcode:1956Natur.178..323A. doi:10.1038/178323b0. PMID 13358737. S2CID 4221816.
  3. ^ a b c Noyori, R. (2003). "Asymmetric Catalysis: Science and Opportunities (Nobel Lecture 2001)". Advanced Synthesis & Catalysis. 345 (12): 15–41. doi:10.1002/adsc.200390002.
  4. ^ a b c Knowles, W. S. (2002). "Asymmetric Hydrogenations (Nobel Lecture)". Angewandte Chemie International Edition. 41 (12): 1998–2007. doi:10.1002/1521-3773(20020617)41:12<1998::AID-ANIE1998>3.0.CO;2-8. PMID 19746594.
  5. ^ Crabtree, Robert (1979). "Iridium compounds in catalysis". Accounts of Chemical Research. 12 (9): 331–337. doi:10.1021/ar50141a005.
  6. ^ Pfaltz, A. (2004). "Asymmetric Catalysis Special Feature Part II: Design of chiral ligands for asymmetric catalysis: From C2-symmetric P,P- and N,N-ligands to sterically and electronically nonsymmetrical P,N-ligands". Proceedings of the National Academy of Sciences. 101 (16): 5723–5726. Bibcode:2004PNAS..101.5723P. doi:10.1073/pnas.0307152101. PMC 395974. PMID 15069193.
  7. ^ a b de Vries, Johannes G.; Elsevier, Cornelis J., eds. (2006-10-20). The Handbook of Homogeneous Hydrogenation (1 ed.). Wiley. doi:10.1002/9783527619382. ISBN 978-3-527-31161-3.
  8. ^ Gridnev, I. D.; Imamoto, T. (2004). "On the Mechanism of Stereoselection in Rh-Catalyzed Asymmetric Hydrogenation: A General Approach for Predicting the Sense of Enantioselectivity". Accounts of Chemical Research. 37 (9): 633–644. doi:10.1021/ar030156e. PMID 15379579.
  9. ^ Wen, Jialin; Wang, Fangyuan; Zhang, Xumu (2021). "Asymmetric hydrogenation catalyzed by first-row transition metal complexes". Chemical Society Reviews. 50 (5): 3211–3237. doi:10.1039/D0CS00082E. ISSN 0306-0012.
  10. ^ Gridnev, Ilya D.; Imamoto, Tsuneo (2004-09-01). "On the Mechanism of Stereoselection in Rh-Catalyzed Asymmetric Hydrogenation: A General Approach for Predicting the Sense of Enantioselectivity". Accounts of Chemical Research. 37 (9): 633–644. doi:10.1021/ar030156e. ISSN 0001-4842.
  11. ^ Imamoto, Tsuneo; Tamura, Ken; Zhang, Zhenfeng; Horiuchi, Yumi; Sugiya, Masashi; Yoshida, Kazuhiro; Yanagisawa, Akira; Gridnev, Ilya D. (2012-01-25). "Rigid P-Chiral Phosphine Ligands with tert -Butylmethylphosphino Groups for Rhodium-Catalyzed Asymmetric Hydrogenation of Functionalized Alkenes". Journal of the American Chemical Society. 134 (3): 1754–1769. doi:10.1021/ja209700j. ISSN 0002-7863.
  12. ^ a b c d Dub, Pavel A.; Gordon, John C. (2018). "The role of the metal-bound N–H functionality in Noyori-type molecular catalysts". Nature Reviews Chemistry. 2 (12): 396–408. doi:10.1038/s41570-018-0049-z. S2CID 106394152.
  13. ^ Noyori, Ryōji; Masashi Yamakawa; Shohei Hashiguchi (2001-11-01). "Metal−Ligand Bifunctional Catalysis: A Nonclassical Mechanism for Asymmetric Hydrogen Transfer between Alcohols and Carbonyl Compounds". The Journal of Organic Chemistry. 66 (24): 7931–7944. doi:10.1021/jo010721w. PMID 11722188.
  14. ^ Ohkuma, T.; Ooka, H.; Ikariya, T.; Noyori, R. (1995), "Preferential hydrogenation of aldehydes and ketones", Journal of the American Chemical Society, 117 (41): 10417–10418, doi:10.1021/ja00146a041
  15. ^ a b Ikariya, T.; Blacker, A. J. (2007). "Asymmetric Transfer Hydrogenation of Ketones with Bifunctional Transition Metal-Based Molecular Catalysts". Accounts of Chemical Research. 40 (12): 1300–1308. doi:10.1021/ar700134q. PMID 17960897.
  16. ^ a b Pilkington, C.; Lennon, I. (2003). "The Application of Asymmetric Hydrogenation for the Manufacture of Pharmaceutical Intermediates:The Need for Catalyst Diversity". Synthesis. 2003 (11): 1639. doi:10.1055/s-2003-40871.
  17. ^ a b c Noyori, R. (2002), "Asymmetric Catalysis: Science and Opportunities (Nobel Lecture)", Angewandte Chemie International Edition, 41 (12): 2008–22, doi:10.1002/1521-3773(20020617)41:12<2008::aid-anie2008>3.0.co;2-4, PMID 19746595
  18. ^ Enthaler, S.; Junge, K.; Beller, M. (2008). "Sustainable Metal Catalysis with Iron: From Rust to a Rising Star?". Angewandte Chemie International Edition. 47 (18): 3317–21. doi:10.1002/anie.200800012. PMID 18412184.
  19. ^ Mikhailine, A.; Lough, A. J.; Morris, R. H. (2009). "Efficient Asymmetric Transfer Hydrogenation of Ketones Catalyzed by an Iron Complex Containing a P−N−N−P Tetradentate Ligand Formed by Template Synthesis". Journal of the American Chemical Society. 131 (4): 1394–1395. doi:10.1021/ja809493h. PMID 19133772.
  20. ^ Sonnenberg, J. F.; Coombs, N.; Dube, P. A.; Morris, R. H. (2012). "Iron Nanoparticles Catalyzing the Asymmetric Transfer Hydrogenation of Ketones". Journal of the American Chemical Society. 134 (13): 5893–5899. doi:10.1021/ja211658t. PMID 22448656.
  21. ^ Whitesell, J. K. (1989). "C2 symmetry and asymmetric induction". Chemical Reviews. 89 (7): 1581–1590. doi:10.1021/cr00097a012.
  22. ^ Knowles, W. S.; Sabacky, M. J.; Vineyard, B. D. (1972). "Catalytic asymmetric hydrogenation". Journal of the Chemical Society, Chemical Communications. 214 (1): 119–124. doi:10.1039/C39720000010. PMID 4270504.
  23. ^ a b Jerphagnon, T.; Renaud, J. L.; Bruneau, C. (2004). "Chiral monodentate phosphorus ligands for rhodium-catalyzed asymmetric hydrogenation". Tetrahedron: Asymmetry. 15 (14): 2101. doi:10.1016/j.tetasy.2004.04.037.
  24. ^ Van Den Berg, M.; Minnaard, A. J.; Schudde, E. P.; Van Esch, J.; De Vries, A. H. M.; De Vries, J. G.; Feringa, B. L. (2000). "Highly Enantioselective Rhodium-Catalyzed Hydrogenation with Monodentate Ligands" (PDF). Journal of the American Chemical Society. 122 (46): 11539. doi:10.1021/ja002507f. hdl:11370/3c92d080-f024-45fe-b997-b100634bd612. S2CID 95403641.
  25. ^ Fu, Y.; Xie, J. H.; Hu, A. G.; Zhou, H.; Wang, L. X.; Zhou, Q. L. (2002). "Novel monodentate spiro phosphorus ligands for rhodium-catalyzed hydrogenation reactions". Chemical Communications (5): 480–481. doi:10.1039/B109827F. PMID 12120551.
  26. ^ Reetz, M. T.; Sell, T.; Meiswinkel, A.; Mehler, G. (2003). "A New Principle in Combinatorial Asymmetric Transition-Metal Catalysis: Mixtures of Chiral Monodentate P Ligands". Angewandte Chemie International Edition. 42 (7): 790–3. doi:10.1002/anie.200390209. PMID 12596201.
  27. ^ Vineyard, B. D.; Knowles, W. S.; Sabacky, M. J.; Bachman, G. L.; Weinkauff, D. J. (1977). "Asymmetric hydrogenation. Rhodium chiral bisphosphine catalyst". Journal of the American Chemical Society. 99 (18): 5946. doi:10.1021/ja00460a018.
  28. ^ Knowles, W. S.; Sabacky, M. J.; Vineyard, B. D.; Weinkauff, D. J. (1975). "Asymmetric hydrogenation with a complex of rhodium and a chiral bisphosphine". Journal of the American Chemical Society. 97 (9): 2567. doi:10.1021/ja00842a058.
  29. ^ Müller, D.; Umbricht, G.; Weber, B.; Pfaltz, A. (1991). "C2-Symmetric 4,4',5,5'-Tetrahydrobi(oxazoles) and 4,4',5,5'-Tetrahydro-2,2'-methylenebis[oxazoles] as Chiral Ligands for Enantioselective Catalysis Preliminary Communication". Helvetica Chimica Acta. 74: 232–240. doi:10.1002/hlca.19910740123.
  30. ^ a b Helmchen, G. N.; Pfaltz, A. (2000). "PhosphinooxazolinesA New Class of Versatile, Modular P,N-Ligands for Asymmetric Catalysis". Accounts of Chemical Research. 33 (6): 336–345. doi:10.1021/ar9900865. PMID 10891051.
  31. ^ Lightfoot, A.; Schnider, P.; Pfaltz, A. (1998). "Enantioselective Hydrogenation of Olefins with Iridium-Phosphanodihydrooxazole Catalysts". Angewandte Chemie International Edition. 37 (20): 2897–2899. doi:10.1002/(SICI)1521-3773(19981102)37:20<2897::AID-ANIE2897>3.0.CO;2-8. PMID 29711115.
  32. ^ Franzke, A.; Pfaltz, A. (2011). "Zwitterionic Iridium Complexes with P,N-Ligands as Catalysts for the Asymmetric Hydrogenation of Alkenes". Chemistry: A European Journal. 17 (15): 4131–44. doi:10.1002/chem.201003314. PMID 21381140.
  33. ^ Maurer, F.; Huch, V.; Ullrich, A.; Kazmaier, U. (2012). "Development of Catalysts for the Stereoselective Hydrogenation of α,β-Unsaturated Ketones". The Journal of Organic Chemistry. 77 (11): 5139–5143. doi:10.1021/jo300246c. PMID 22571628.
  34. ^ Rageot, D.; Woodmansee, D. H.; Pugin, B. T.; Pfaltz, A. (2011). "Proline-Based P,O Ligand/Iridium Complexes as Highly Selective Catalysts: Asymmetric Hydrogenation of Trisubstituted Alkenes". Angewandte Chemie International Edition. 50 (41): 9598–601. doi:10.1002/anie.201104105. PMID 21882320.
  35. ^ a b c Perry, M. C.; Cui, X.; Powell, M. T.; Hou, D. R.; Reibenspies, J. H.; Burgess, K. (2003). "Optically Active Iridium Imidazol-2-ylidene-oxazoline Complexes: Preparation and Use in Asymmetric Hydrogenation of Arylalkenes". Journal of the American Chemical Society. 125 (1): 113–123. doi:10.1021/ja028142b. PMID 12515512.
  36. ^ a b Nanchen, S.; Pfaltz, A. (2006). "Synthesis and Application of Chiral N-Heterocyclic Carbene–Oxazoline Ligands: Iridium-Catalyzed Enantioselective Hydrogenation". Chemistry: A European Journal. 12 (17): 4550–8. doi:10.1002/chem.200501500. PMID 16557626.
  37. ^ Zhu, Y.; Burgess, K. (2008). "Iridium-Catalyzed Asymmetric Hydrogenation of Vinyl Ethers". Advanced Synthesis & Catalysis. 350 (7–8): 979. doi:10.1002/adsc.200700546.
  38. ^ Zhao, J.; Burgess, K. (2009). "Aldol-Type Chirons from Asymmetric Hydrogenations of Trisubstituted Alkenes". Organic Letters. 11 (10): 2053–2056. doi:10.1021/ol900308w. PMID 19368378.
  39. ^ Zhao, J.; Burgess, K. (2009). "Synthesis of Vicinal Dimethyl Chirons by Asymmetric Hydrogenation of Trisubstituted Alkenes". Journal of the American Chemical Society. 131 (37): 13236–13237. doi:10.1021/ja905458n. PMID 19719102.
  40. ^ Zhou, J.; Burgess, K. (2007). "Α,ω-Functionalized 2,4-Dimethylpentane Dyads and 2,4,6-Trimethylheptane Triads through Asymmetric Hydrogenation". Angewandte Chemie International Edition. 46 (7): 1129–31. doi:10.1002/anie.200603635. PMID 17200966.
  41. ^ Zhou, J.; Zhu, Y.; Burgess, K. (2007). "Synthesis of (S,R,R,S,R,S)-4,6,8,10,16,18- Hexamethyldocosane from Antitrogus parvulus via Diastereoselective Hydrogenations". Organic Letters. 9 (7): 1391–1393. doi:10.1021/ol070298z. PMID 17338543.
  42. ^ Urban, S.; Ortega, N.; Glorius, F. (2011). "Ligand-Controlled Highly Regioselective and Asymmetric Hydrogenation of Quinoxalines Catalyzed by Ruthenium N-Heterocyclic Carbene Complexes". Angewandte Chemie International Edition. 50 (16): 3803–6. doi:10.1002/anie.201100008. PMID 21442699.
  43. ^ Cui, X.; Burgess, K. (2005). "Catalytic Homogeneous Asymmetric Hydrogenations of Largely Unfunctionalized Alkenes". Chemical Reviews. 105 (9): 3272–3296. doi:10.1021/cr0500131. PMID 16159153.
  44. ^ Pàmies, O.; Andersson, P. G.; Diéguez, M. (2010). "Asymmetric Hydrogenation of Minimally Functionalised Terminal Olefins: An Alternative Sustainable and Direct Strategy for Preparing Enantioenriched Hydrocarbons". Chemistry: A European Journal. 16 (48): 14232–40. doi:10.1002/chem.201001909. PMID 21140401.
  45. ^ Woodmansee, D. H.; Pfaltz, A. (2011). "Asymmetric hydrogenation of alkenes lacking coordinating groups". Chemical Communications. 47 (28): 7912–7916. doi:10.1039/c1cc11430a. PMID 21556431.
  46. ^ Mazuela, J.; Verendel, J. J.; Coll, M.; SchäFfner, B. N.; BöRner, A.; Andersson, P. G.; PàMies, O.; DiéGuez, M. (2009). "Iridium Phosphite−Oxazoline Catalysts for the Highly Enantioselective Hydrogenation of Terminal Alkenes". Journal of the American Chemical Society. 131 (34): 12344–12353. doi:10.1021/ja904152r. PMID 19658416.
  47. ^ Hou, G. H.; Xie, J. H.; Wang, L. X.; Zhou, Q. L. (2006). "Highly Efficient Rh(I)-Catalyzed Asymmetric Hydrogenation of Enamines Using Monodente Spiro Phosphonite Ligands". Journal of the American Chemical Society. 128 (36): 11774–11775. doi:10.1021/ja0644778. PMID 16953614.
  48. ^ Hou, G. H.; Xie, J. H.; Yan, P. C.; Zhou, Q. L. (2009). "Iridium-Catalyzed Asymmetric Hydrogenation of Cyclic Enamines". Journal of the American Chemical Society. 131 (4): 1366–1367. doi:10.1021/ja808358r. PMID 19132836.
  49. ^ Noyori, R.; Ohkuma, T. (2001). "Asymmetric Catalysis by Architectural and Functional Molecular Engineering: Practical Chemo- and Stereoselective Hydrogenation of Ketones". Angewandte Chemie International Edition. 40 (1): 40–73. doi:10.1002/1521-3773(20010105)40:1<40::AID-ANIE40>3.0.CO;2-5. PMID 11169691.
  50. ^ Hems, W. P.; Groarke, M.; Zanotti-Gerosa, A.; Grasa, G. A. (2007). "[(Bisphosphine) Ru(II) Diamine] Complexes in Asymmetric Hydrogenation: Expanding the Scope of the Diamine Ligand". Accounts of Chemical Research. 40 (12): 1340–1347. doi:10.1021/ar7000233. PMID 17576143.
  51. ^ Noyori, R.; Yamakawa, M.; Hashiguchi, S. (2001). "Metal−Ligand Bifunctional Catalysis: A Nonclassical Mechanism for Asymmetric Hydrogen Transfer between Alcohols and Carbonyl Compounds". The Journal of Organic Chemistry. 66 (24): 7931–7944. doi:10.1021/jo010721w. PMID 11722188.
  52. ^ Zhu, S. F.; Xie, J. B.; Zhang, Y. Z.; Li, S.; Zhou, Q. L. (2006). "Well-Defined Chiral Spiro Iridium/Phosphine−Oxazoline Cationic Complexes for Highly Enantioselective Hydrogenation of Imines at Ambient Pressure". Journal of the American Chemical Society. 128 (39): 12886–12891. doi:10.1021/ja063444p. PMID 17002383.
  53. ^ Xie, J. H.; Liu, X. Y.; Xie, J. B.; Wang, L. X.; Zhou, Q. L. (2011). "An Additional Coordination Group Leads to Extremely Efficient Chiral Iridium Catalysts for Asymmetric Hydrogenation of Ketones". Angewandte Chemie International Edition. 50 (32): 7329–32. doi:10.1002/anie.201102710. PMID 21751315.
  54. ^ Ohkuma, T.; Ooka, H.; Yamakawa, M.; Ikariya, T.; Noyori, R. (1996), "Stereoselective Hydrogenation of Simple Ketones Catalyzed by Ruthenium(II) Complexes", The Journal of Organic Chemistry, 61 (15): 4872–4873, doi:10.1021/jo960997h
  55. ^ Noyori, R.; Ohkuma, T. (2001), "Asymmetric Catalysis by Architectural and Functional Molecular Engineering: Practical Chemo- and Stereoselective Hydrogenation of Ketones", Angewandte Chemie International Edition, 40 (1): 40–73, doi:10.1002/1521-3773(20010105)40:1<40::aid-anie40>3.0.co;2-5, PMID 11169691
  56. ^ a b Zhou, Y. G. (2007). "Asymmetric Hydrogenation of Heteroaromatic Compounds". Accounts of Chemical Research. 40 (12): 1357–1366. CiteSeerX 10.1.1.653.5495. doi:10.1021/ar700094b. PMID 17896823.
  57. ^ Wang, W. B.; Lu, S. M.; Yang, P. Y.; Han, X. W.; Zhou, Y. G. (2003). "Highly Enantioselective Iridium-Catalyzed Hydrogenation of Heteroaromatic Compounds, Quinolines". Journal of the American Chemical Society. 125 (35): 10536–10537. CiteSeerX 10.1.1.651.3119. doi:10.1021/ja0353762. PMID 12940733.
  58. ^ a b Xu, L.; Lam, K. H.; Ji, J.; Wu, J.; Fan, Q. H.; Lo, W. H.; Chan, A. S. C. (2005). "Air-stable Ir-(P-Phos) complex for highly enantioselective hydrogenation of quinolines and their immobilization in poly(ethylene glycol) dimethyl ether (DMPEG)". Chemical Communications (11): 1390–2. doi:10.1039/B416397D. hdl:10397/8906. PMID 15756313.
  59. ^ Lam, K. H.; Xu, L.; Feng, L.; Fan, Q. H.; Lam, F. L.; Lo, W. H.; Chan, A. S. C. (2005). "Highly Enantioselective Iridium-Catalyzed Hydrogenation of Quinoline Derivatives Using Chiral Phosphinite H8-BINAPO". Advanced Synthesis & Catalysis. 347 (14): 1755. doi:10.1002/adsc.200505130. hdl:10397/26878.
  60. ^ Wang, Z. J.; Deng, G. J.; Li, Y.; He, Y. M.; Tang, W. J.; Fan, Q. H. (2007). "Enantioselective Hydrogenation of Quinolines Catalyzed by Ir(BINAP)-Cored Dendrimers: Dramatic Enhancement of Catalytic Activity". Organic Letters. 9 (7): 1243–1246. doi:10.1021/ol0631410. PMID 17328554.
  61. ^ Qiu, L.; Kwong, F. Y.; Wu, J.; Lam, W. H.; Chan, S.; Yu, W. Y.; Li, Y. M.; Guo, R.; Zhou, Z.; Chan, A. S. C. (2006). "A New Class of Versatile Chiral-Bridged Atropisomeric Diphosphine Ligands: Remarkably Efficient Ligand Syntheses and Their Applications in Highly Enantioselective Hydrogenation Reactions". Journal of the American Chemical Society. 128 (17): 5955–5965. doi:10.1021/ja0602694. hdl:10397/60397. PMID 16637664.
  62. ^ Reetz, M. T.; Li, X. (2006). "Asymmetric hydrogenation of quinolines catalyzed by iridium complexes of BINOL-derived diphosphonites". Chemical Communications (20): 2159–60. doi:10.1039/b602320g. PMID 16703140.
  63. ^ Rueping; Antonchick, A.; Theissmann, T. (2006). "A highly enantioselective Brønsted acid catalyzed cascade reaction: organocatalytic transfer hydrogenation of quinolines and their application in the synthesis of alkaloids". Angewandte Chemie International Edition in English. 45 (22): 3683–3686. doi:10.1002/anie.200600191. PMID 16639754.
  64. ^ Tang, W.; Xu, L.; Fan, Q. H.; Wang, J.; Fan, B.; Zhou, Z.; Lam, K. H.; Chan, A. S. C. (2009). "Asymmetric Hydrogenation of Quinoxalines with Diphosphinite Ligands: A Practical Synthesis of Enantioenriched, Substituted Tetrahydroquinoxalines". Angewandte Chemie International Edition. 48 (48): 9135–8. doi:10.1002/anie.200904518. hdl:10397/20432. PMID 19876991.
  65. ^ Rueping, M.; Tato, F.; Schoepke, F. R. (2010). "The First General, Efficient and Highly Enantioselective Reduction of Quinoxalines and Quinoxalinones". Chemistry: A European Journal. 16 (9): 2688–91. doi:10.1002/chem.200902907. PMID 20140920.
  66. ^ Glorius, F.; Spielkamp, N.; Holle, S.; Goddard, R.; Lehmann, C. W. (2004). "Efficient Asymmetric Hydrogenation of Pyridines". Angewandte Chemie International Edition. 43 (21): 2850–2. doi:10.1002/anie.200453942. PMID 15150766.
  67. ^ Ye, Z. S.; Chen, M. W.; Chen, Q. A.; Shi, L.; Duan, Y.; Zhou, Y. G. (2012). "Iridium-Catalyzed Asymmetric Hydrogenation of Pyridinium Salts". Angewandte Chemie International Edition. 51 (40): 10181–4. doi:10.1002/anie.201205187. PMID 22969060.
  68. ^ Tang, W. J.; Tan, J.; Xu, L. J.; Lam, K. H.; Fan, Q. H.; Chan, A. S. C. (2010). "Highly Enantioselective Hydrogenation of Quinoline and Pyridine Derivatives with Iridium-(P-Phos) Catalyst". Advanced Synthesis & Catalysis. 352 (6): 1055. doi:10.1002/adsc.200900870. hdl:10397/22884.
  69. ^ Rueping, M.; Antonchick, A. P. (2007). "Organocatalytic Enantioselective Reduction of Pyridines". Angewandte Chemie International Edition. 46 (24): 4562–5. doi:10.1002/anie.200701158. PMID 17492817.
  70. ^ Baeza, A.; Pfaltz, A. (2010). "Iridium-Catalyzed Asymmetric Hydrogenation of N-Protected Indoles". Chemistry: A European Journal. 16 (7): 2036–9. doi:10.1002/chem.200903105. PMID 20104554.
  71. ^ Xiao, Y. C.; Wang, C.; Yao, Y.; Sun, J.; Chen, Y. C. (2011). "Direct Asymmetric Hydrosilylation of Indoles: Combined Lewis Base and Brønsted Acid Activation". Angewandte Chemie International Edition. 50 (45): 10661–4. doi:10.1002/anie.201105341. PMID 21932274.
  72. ^ Duan, Y.; Chen, M. W.; Ye, Z. S.; Wang, D. S.; Chen, Q. A.; Zhou, Y. G. (2011). "An Enantioselective Approach to 2,3-Disubstituted Indolines through Consecutive Brønsted Acid/Pd-Complex-Promoted Tandem Reactions". Chemistry: A European Journal. 17 (26): 7193–7. doi:10.1002/chem.201100576. PMID 21567504.
  73. ^ Wang, D. S.; Ye, Z. S.; Chen, Q. A.; Zhou, Y. G.; Yu, C. B.; Fan, H. J.; Duan, Y. (2011). "Highly Enantioselective Partial Hydrogenation of Simple Pyrroles: A Facile Access to Chiral 1-Pyrrolines". Journal of the American Chemical Society. 133 (23): 8866–8869. doi:10.1021/ja203190t. PMID 21591641.
  74. ^ Wang, D. S.; Chen, Q. A.; Lu, S. M.; Zhou, Y. G. (2012). "Asymmetric Hydrogenation of Heteroarenes and Arenes". Chemical Reviews. 112 (4): 2557–2590. doi:10.1021/cr200328h. PMID 22098109.
  75. ^ Urban, S.; Beiring, B.; Ortega, N.; Paul, D.; Glorius, F. (2012). "Asymmetric Hydrogenation of Thiophenes and Benzothiophenes". Journal of the American Chemical Society. 134 (37): 15241–15244. doi:10.1021/ja306622y. PMID 22934527.
  76. ^ a b Heitbaum, M.; Glorius, F.; Escher, I. (2006). "Asymmetric Heterogeneous Catalysis". Angewandte Chemie International Edition. 45 (29): 4732–62. doi:10.1002/anie.200504212. PMID 16802397.
  77. ^ Yoon, M.; Srirambalaji, R.; Kim, K. (2012). "Homochiral Metal–Organic Frameworks for Asymmetric Heterogeneous Catalysis". Chemical Reviews. 112 (2): 1196–1231. doi:10.1021/cr2003147. PMID 22084838.
  78. ^ Hu, A.; Ngo, H. L.; Lin, W. (2003). "Chiral Porous Hybrid Solids for Practical Heterogeneous Asymmetric Hydrogenation of Aromatic Ketones". Journal of the American Chemical Society. 125 (38): 11490–11491. doi:10.1021/ja0348344. PMID 13129339.
  79. ^ Blaser, H. U.; Spindler, F.; Studer, M. (2001). "Enantioselective catalysis in fine chemicals production". Applied Catalysis A: General. 221 (1–2): 119–143. doi:10.1016/S0926-860X(01)00801-8. PMID 12613584.
  80. ^ Blaser, Hans-Ulrich; Federsel, Hans-Jürgen, eds. (2010). Asymmetric Catalysis on Industrial Scale. Weinheim: Wiley-VCH. pp. 13–16. doi:10.1002/9783527630639. ISBN 978-3-527-63063-9.
  81. ^ Jacobsen, E.N.; Pfaltz, Andreas; Yamamato, H., eds. (1999). Comprehensive Asymmetric Catalysis. Berlin; New York: Springer. pp. 1443–1445. ISBN 978-3-540-64336-4.

asymmetric, hydrogenation, chemical, reaction, that, adds, atoms, hydrogen, target, substrate, molecule, with, three, dimensional, spatial, selectivity, critically, this, selectivity, does, come, from, target, molecule, itself, from, other, reagents, catalysts. Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target substrate molecule with three dimensional spatial selectivity Critically this selectivity does not come from the target molecule itself but from other reagents or catalysts present in the reaction This allows spatial information what chemists refer to as chirality to transfer from one molecule to the target forming the product as a single enantiomer The chiral information is most commonly contained in a catalyst and in this case the information in a single molecule of catalyst may be transferred to many substrate molecules amplifying the amount of chiral information present Similar processes occur in nature where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer such as amino acids that a cell needs to function By imitating this process chemists can generate many novel synthetic molecules that interact with biological systems in specific ways leading to new pharmaceutical agents and agrochemicals The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers William Standish Knowles and Ryōji Noyori being collectively awarded one half of the 2001 Nobel Prize in Chemistry 1 Contents 1 History 2 Mechanism 2 1 Inner sphere mechanisms 2 2 Outer sphere mechanisms 3 Metals 3 1 Base metals 4 Ligand classes 4 1 Phosphine ligands 4 1 1 Monodentate phosphines 4 1 2 Chiral diphosphine ligands 4 1 3 P N and P O ligands 4 1 4 NHC ligands 5 Acyclic substrates 5 1 Nonpolar substrates 5 2 Imines and ketones 6 Aromatic substrates 6 1 Quinolines isoquinolines and quinoxalines 6 2 Pyridines 6 3 Indoles and pyrroles 6 4 Oxygen and sulfur containing heterocycles 7 Heterogeneous catalysis 8 Industrial applications 9 ReferencesHistory editIn 1956 a heterogeneous catalyst made of palladium deposited on silk was shown to effect asymmetric hydrogenation 2 Later in 1968 the groups of William Knowles and Leopold Horner independently published the examples of asymmetric hydrogenation using a homogeneous catalysts While exhibiting only modest enantiomeric excesses these early reactions demonstrated feasibility By 1972 enantiomeric excess of 90 was achieved and the first industrial synthesis of the Parkinson s drug L DOPA commenced using this technology 3 4 nbsp L DOPAThe field of asymmetric hydrogenation continued to experience a number of notable advances Henri Kagan developed DIOP an easily prepared C2 symmetric diphosphine that gave high ee s in certain reactions Ryōji Noyori introduced the ruthenium based catalysts for the asymmetric hydrogenated polar substrates such as ketones and aldehydes Robert H Crabtree demonstrated the ability for Iridium compounds to catalyse asymmetric hydrogenation reactions in 1979 with the invention of Crabtree s catalyst 5 In the early 1990 s the introduction of P N ligands by several groups independently then further expanded the scope of the C2 symmetric ligands although they are not fundamentally superior to chiral ligands lacking rotational symmetry 6 Today asymmetric hydrogenation is a routine methodology in laboratory and industrial scale organic chemistry The importance of asymmetric hydrogenation was recognized by the 2001 Nobel Prize in Chemistry awarded to William Standish Knowles and Ryōji Noyori Mechanism editAsymmetric hydrogenations operate by conventional mechanisms invoked for other hydrogenations This includes inner sphere mechanisms outer sphere mechanisms and the s bond metathesis mechanisms 7 The type of mechanism employed by a catalyst is largely dependent on the ligands used in a system which in turn leads to certain catalyst substrate affinities Inner sphere mechanisms edit The so called inner sphere mechanism entails coordination of the alkene to the metal center 8 Other characteristics of this mechanism include a tendency for a homolytic splitting of dihydrogen when more electron rich low valent metals are present while electron poor high valent metals normally exhibit a heterolytic cleavage of dihydrogen assisted by a base 9 The diagram below depicts purposed mechanisms for catalytic hydrogenation with rhodium complexes which are inner sphere mechanisms In the unsaturated mechanism the chiral product formed will have the opposite mode compared to the catalyst used While the thermodynamically favoured complex between the catalyst and the substrate is unable to undergo hydrogenation the unstable unfavoured complex undergoes hydrogenation rapidly 10 The dihydride mechanism on the other hand sees the complex initially hydrogenated to the dihydride form This subsequently allows for the coordination of the double bond on the non hindered side Through insertion and reductive elimination the product s chirality matches that of the ligand 11 nbsp Proposed mechanisms for asymmetric hydrogenation The preference for producing one enantiomer instead of another in these reactions is often explained in terms of steric interactions between the ligand and the prochiral substrate Consideration of these interactions has led to the development of quadrant diagrams where blocked areas are denoted with a shaded box while open areas are left unfilled In the modeled reaction large groups on an incoming olefin will tend to orient to fill the open areas of the diagram while smaller groups will be directed to the blocked areas and hydrogen delivery will then occur to the back face of the olefin fixing the stereochemistry Note that only part of the chiral phosphine ligand is shown for the sake of clarity nbsp Quadrant model for asymmetric hydrogenation Outer sphere mechanisms edit Some catalysts operate by outer sphere mechanisms such that the substrate never bonds directly to the metal but rather interacts with its ligands which is often a metal hydride and a protic hydrogen on a ligand As such in most cases dihydrogen is split heterolytically with the metal acting as a Lewis acid and either an external or internal base deprotonating the hydride 7 nbsp Proposed intermediates in the outer sphere mechanisms for heterolytic hydrogenation of the catalyst left and hydride transfer from the catalyst to the substate right For an example of this mechanism we can consider the BINAP Ru diamine system The dihalide form of the catalyst is converted to the catalysts by reaction of H2 in the presence of base 12 RuCl2 BINAP diamine 2 KOBu t 2 H2 RuH2 BINAP diamine 2 KCl 2 HOBu tThe resulting catalysts have three kinds of ligands hydrides which transfer to the unsaturated substrate diamines which interact with substrate and with base activator by the second coordination sphere diphosphine which confers asymmetry The Noyori class of catalysts are often referred to as bifunctional catalysts to emphasize the fact that both the metal and the amine ligand are functional 13 In the hydrogenation of C O containing substates the mechanism was long assumed to operate by a six membered pericyclic transition state intermediate whereby the hydrido ruthenium hydride center HRu NH interacts with the carbonyl substrate R2C O 14 More recent DFT and experimental studies have shown that this model is largely incorrect Instead the amine backbone interacts strongly with the base activator which often is used in large excess 12 However in both cases the substate does not bond directly with the metal centre thus making it a great example of an outer sphere mechanism Metals editPractical AH employ platinum metal based catalysts 15 16 17 Base metals edit Iron is a popular research target for many catalytic processes owing largely to its low cost and low toxicity relative to other transition metals 18 Asymmetric hydrogenation methods using iron have been realized although in terms of rates and selectivity they are inferior to catalysts based on precious metals 19 In some cases structurally ill defined nanoparticles have proven to be the active species in situ and the modest selectivity observed may result from their uncontrolled geometries 20 Ligand classes editPhosphine ligands edit Chiral phosphine ligands especially C2 symmetric ligands are the source of chirality in most asymmetric hydrogenation catalysts Of these the BINAP ligand is well known as a result of its Nobel Prize winning application in the Noyori asymmetric hydrogenation 3 Chiral phosphine ligands can be generally classified as mono or bidentate They can be further classified according to the location of the stereogenic centre phosphorus vs the organic substituents Ligands with a C2 symmetry element have been particularly popular in part because the presence of such an element reduces the possible binding conformations of a substrate to a metal ligand complex dramatically often resulting in exceptional enantioselectivity 21 Monodentate phosphines edit Monophosphine type ligands were among the first to appear in asymmetric hydrogenation e g the ligand CAMP 22 Continued research into these types of ligands has explored both P alkyl and P heteroatom bonded ligands with P heteroatom ligands like the phosphites and phosphoramidites generally achieving more impressive results 23 Structural classes of ligands that have been successful include those based on the binapthyl structure of MonoPHOS 24 or the spiro ring system of SiPHOS 25 Notably these monodentate ligands can be used in combination with each other to achieve a synergistic improvement in enantioselectivity 26 something that is not possible with the diphosphine ligands 23 nbsp A ferrocene derivative nbsp The CAMP ligand nbsp A BINOL derivative Chiral diphosphine ligands edit The diphosphine ligands have received considerably more attention than the monophosphines and perhaps as a consequence have a much longer list of achievement This class includes the first ligand to achieve high selectivity DIOP the first ligand to be used in industrial asymmetric synthesis DIPAMP 27 28 4 and what is likely the best known chiral ligand BINAP 3 Chiral diphosphine ligands are now ubiquitous in asymmetric hydrogenation nbsp Historically important diphosphine ligands P N and P O ligands edit nbsp Generic PHOX ligand architecture nbsp Effective ligand for various asymmetric hydrogenation processes The use of P N ligands in asymmetric hydrogenation can be traced to the C2 symmetric bisoxazoline ligand 29 However these symmetric ligands were soon superseded by monooxazoline ligands whose lack of C2 symmetry has in no way limits their efficacy in asymmetric catalysis 30 Such ligands generally consist of an achiral nitrogen containing heterocycle that is functionalized with a pendant phosphorus containing arm although both the exact nature of the heterocycle and the chemical environment phosphorus center has varied widely No single structure has emerged as consistently effective with a broad range of substrates although certain privileged structures like the phosphine oxazoline or PHOX architecture have been established 31 30 32 Moreover within a narrowly defined substrate class the performance of metallic complexes with chiral P N ligands can closely approach perfect conversion and selectivity in systems otherwise very difficult to target 33 Certain complexes derived from chelating P O ligands have shown promising results in the hydrogenation of a b unsaturated ketones and esters 34 NHC ligands edit nbsp Catalyst developed by Burgess for asymmetric hydrogenationSimple N heterocyclic carbene NHC based ligands have proven impractical for asymmetrical hydrogenation Some C N ligands combine an NHC with a chiral oxazoline to give a chelating ligand 35 36 NHC based ligands of the first type have been generated as large libraries from the reaction of smaller libraries of individual NHCs and oxazolines 35 36 NHC based catalysts featuring a bulky seven membered metallocycle on iridium have been applied to the catalytic hydrogenation of unfunctionalized olefins 35 and vinyl ether alcohols with conversions and ee s in the high 80s or 90s 37 The same system has been applied to the synthesis of a number of aldol 38 vicinal dimethyl 39 and deoxypolyketide 40 motifs and to the deoxypolyketides themselves 41 C2 symmetric NHCs have shown themselves to be highly useful ligands for the asymmetric hydrogenation 42 Acyclic substrates editSubstrates can be classified according to their polarity Nonpolar substrates are dominated by alkenes Polar substrates include ketones enamines ketimines Nonpolar substrates edit 43 Alkenes that are particularly amenable to asymmetric hydrogenation often feature a polar functional group adjacent to the site to be hydrogenated In the absence of this functional group catalysis often results in low ee s For some unfunctionalized olefins iridium with P N based ligands have proven effective however Alkene substrates are often classified according to their substituents e g 1 1 disubstituted 1 2 diaryl trisubstituted 1 1 2 trialkyl and tetrasubstituted olefins 44 45 and even within these classes variations may exist that make different solutions optimal 46 nbsp Example of asymmetric hydrogenation of unfunctionalized olefins nbsp Chiral phosphoramidite and phosphonite ligands used in the asymmetric hydrogenation of enaminesConversely to the case of olefins asymmetric hydrogenation of enamines has favoured diphosphine type ligands excellent results have been achieved with both iridium and rhodium based systems However even the best systems often suffer from low ee s and a lack of generality Certain pyrrolidine derived enamines of aromatic ketones are amenable to asymmetrically hydrogenation with cationic rhodium I phosphonite systems and I2 and acetic acid system with ee values usually above 90 and potentially as high as 99 9 47 A similar system using iridium I and a very closely related phosphoramidite ligand is effective for the asymmetric hydrogenation of pyrrolidine type enamines where the double bond was inside the ring in other words of dihydropyrroles 48 In both cases the enantioselectivity dropped substantially when the ring size was increased from five to six Imines and ketones edit nbsp Noyori catalyst for asymmetric hydrogenation of ketonesKetones and imines are related functional groups and effective technologies for the asymmetric hydrogenation of each are also closely related Early examples are Noyori s ruthenium chiral diphosphine diamine system 49 50 For carbonyl and imine substrates end on h1 coordination can compete with h2 mode For h1 bound substrates the hydrogen accepting carbon is removed from the catalyst and resists hydrogenation 51 Iridium P N ligand based systems have been effective for some ketones and imines For example a consistent system for benzylic aryl imines uses the P N ligand SIPHOX in conjunction with iridium I in a cationic complex to achieve asymmetric hydrogenation with ee gt 90 52 An efficient catalyst for ketones turnover number TON up to 4 550 000 and ee up to 99 9 is an iridium I system with a closely related tridentate ligand 53 nbsp Highly effective system for the asymmetric hydrogenation of ketones The BINAP diamine Ru catalyst is effective for the asymmetric reduction of both functionalized and simple ketones 54 and BINAP diamine Ru catalyst can catalyze aromatic heteroaromatic and olefinic ketones enantioselectively 55 Better stereoselectivity is achieved when one substituent is larger than the other see Flippin Lodge angle nbsp BINAP diamine Ru catalyst scopeAromatic substrates editThe asymmetric hydrogenation of aromatic especially heteroaromatic substrates is a very active field of ongoing research Catalysts in this field must contend with a number of complicating factors including the tendency of highly stable aromatic compounds to resist hydrogenation the potential coordinating and therefore catalyst poisoning abilities of both substrate and product and the great diversity in substitution patterns that may be present on any one aromatic ring 56 Of these substrates the most consistent success has been seen with nitrogen containing heterocycles where the aromatic ring is often activated either by protonation or by further functionalization of the nitrogen generally with an electron withdrawing protecting group Such strategies are less applicable to oxygen and sulfur containing heterocycles since they are both less basic and less nucleophilic this additional difficulty may help to explain why few effective methods exist for their asymmetric hydrogenation Quinolines isoquinolines and quinoxalines edit Two systems exist for the asymmetric hydrogenation of 2 substituted quinolines with isolated yields generally greater than 80 and ee values generally greater than 90 The first is an iridium I chiral phosphine I2 system first reported by Zhou et al 57 While the first chiral phosphine used in this system was MeOBiPhep newer iterations have focused on improving the performance of this ligand To this end systems use phosphines or related ligands with improved air stability 58 recyclability 58 ease of preparation 59 lower catalyst loading 60 61 and the potential role of achiral phosphine additives 62 As of October 2012 no mechanism appears to have been proposed although both the necessity of I2 or a halogen surrogate and the possible role of the heteroaromatic N in assisting reactivity have been documented 56 The second is an organocatalytic transfer hydrogenation system based on Hantzsch esters and a chiral Bronsted acid In this case the authors envision a mechanism where the isoquinoline is alternately protonated in an activating step then reduced by conjugate addition of hydride from the Hantzsch ester 63 nbsp Proposed organocatalytic mechanism Much of the asymmetric hydrogenation chemistry of quinoxalines is closely related to that of the structurally similar quinolines Effective and efficient results can be obtained with an Ir I phophinite I2 system 64 and a Hantzsh ester based organocatalytic system 65 both of which are similar to the systems discussed earlier with regards to quinolines Pyridines edit Pyridines are highly variable substrates for asymmetric reduction even compared to other heteroaromatics in that five carbon centers are available for differential substitution on the initial ring As of October 2012 no method seems to exist that can control all five although at least one reasonably general method exists The most general method of asymmetric pyridine hydrogenation is actually a heterogeneous method where asymmetry is generated from a chiral oxazolidinone bound to the C2 position of the pyridine Hydrogenating such functionalized pyridines over a number of different heterogeneous metal catalysts gave the corresponding piperidine with the substituents at C3 C4 and C5 positions in an all cis geometry in high yield and excellent enantioselectivity The oxazolidinone auxiliary is also conveniently cleaved under the hydrogenation conditions 66 nbsp Asymmetric hydrogenation of pyridines with heterogeneous catalyst Methods designed specifically for 2 substituted pyridine hydrogenation can involve asymmetric systems developed for related substrates like 2 substituted quinolines and quinoxalines For example an iridium I chiral phosphine I2 system is effective in the asymmetric hydrogenation of activated alkylated 2 pyridiniums 67 or certain cyclohexanone fused pyridines 68 Similarly chiral Bronsted acid catalysis with a Hantzsh ester as a hydride source is effective for some 2 alkyl pyridines with additional activating substitution 69 Indoles and pyrroles edit The asymmetric hydrogenation of indoles has been established with N Boc protection 70 nbsp Method for asymmetric hydrogenation of Boc protected indoles A Pd TFA 2 H8 BINAP system achieves the enantioselective cis hydrogenation of 2 3 and 2 substituted indoles 71 72 nbsp Sequential alkylation and asymmetric hydrogenation of 2 substituted indoles Akin to the behavior of indoles pyrroles can be converted to pyrrolidines by asymmetric hydrogenation 73 nbsp The asymmetric hydrogenation of 2 3 5 substituted N Boc pyrroles Oxygen and sulfur containing heterocycles edit The asymmetric hydrogenation of furans and benzofurans is challenging 74 nbsp The asymmetric hydrogenation of furans and benzofurans Asymmetric hydrogenation of thiophenes and benzothiophenes has been catalyzed by some ruthenium II complexes of N heterocyclic carbenes NHC This system appears to possess superb selectivity ee gt 90 and perfect diastereoselectivity all cis if the substrate has a fused or directly bound phenyl ring but yields only racemic product in all other tested cases 75 nbsp The asymmetric hydrogenation of thiophenes and benzothiophenesHeterogeneous catalysis editNo heterogeneous catalyst has been commercialized for asymmetric hydrogenation The first asymmetric hydrogenation focused on palladium deposited on a silk support Cinchona alkaloids have been used as chiral modifiers for enantioselectivity hydrogenation 76 nbsp Cinchonidine one of the cinchona alkaloidsAn alternative technique and one that allows more control over the structural and electronic properties of active catalytic sites is the immobilization of catalysts that have been developed for homogeneous catalysis on a heterogeneous support Covalent bonding of the catalyst to a polymer or other solid support is perhaps most common although immobilization of the catalyst may also be achieved by adsorption onto a surface ion exchange or even physical encapsulation One drawback of this approach is the potential for the proximity of the support to change the behaviour of the catalyst lowering the enantioselectivity of the reaction To avoid this the catalyst is often bound to the support by a long linker though cases are known where the proximity of the support can actually enhance the performance of the catalyst 76 The final approach involves the construction of MOFs that incorporate chiral reaction sites from a number of different components potentially including chiral and achiral organic ligands structural metal ions catalytically active metal ions and or preassembled catalytically active organometallic cores 77 One of these involved ruthenium based catalysts As little as 0 005 mol of such catalysts proved sufficient to achieve the asymmetric hydrogenation of aryl ketones although the usual conditions featured 0 1 mol of catalyst and resulted in an enantiomeric excess of 90 6 99 2 78 nbsp The active site of a heterogeneous zirconium phosphonate catalyst for asymmetric hydrogenationIndustrial applications edit nbsp S S Ro 67 8867Asymmetric hydrogenations are used in the production of several drugs such as the antibacterial levofloxin the antibiotic carbapenem and the antipsychotic agent BMS181100 15 16 17 nbsp Noyori asymmetric hydrogenationKnowles research into asymmetric hydrogenation and its application to the production scale synthesis of L Dopa 4 gave asymmetric hydrogenation a strong start in the industrial world A 2001 review indicated that asymmetric hydrogenation accounted for 50 of production scale 90 of pilot scale and 74 of bench scale catalytic enantioselective processes in industry with the caveat that asymmetric catalytic methods in general were not yet widely used 79 Asymmetric hydrogenation has replaced kinetic resolution based methods has resulted in substantial improvements in the process s efficiency 12 can be seen in a number of specific cases where the For example Roche s Catalysis Group was able to achieve the synthesis of S S Ro 67 8867 in 53 overall yield a dramatic increase above the 3 5 that was achieved in the resolution based synthesis 80 Roche s synthesis of mibefradil was likewise improved by replacing resolution with asymmetric hydrogenation reducing the step count by three and increasing the yield of a key intermediate to 80 from the original 70 81 nbsp Asymmetric hydrogenation in the industrial synthesis of mibefradilNoyori inspired hydrogenation catalysts have been applied to the commercial synthesis of number of fine chemicals R 1 2 Propandiol precursor to the antibacterial levofloxacin can be efficiently synthesized from hydroxyacetone using Noyori asymmetric hydrogenation 17 nbsp levofloxaxin synthesisNewer routes focus on the hydrogenation of R methyl lactate 12 An antibiotic carbapenem is also prepared using Noyori asymmetric hydrogenation via 2S 3R methyl 2 benzamidomethyl 3 hydroxybutanoate which is synthesized from racemic methyl 2 benzamidomethyl 3 oxobutanoate by dynamic kinetic resolution nbsp carbapenem synthesisAn antipsychotic agent BMS 181100 is synthesized using BINAP diamine Ru catalyst nbsp BMS181110 synthesisReferences edit The Nobel Prize in Chemistry 2001 2001 10 10 Akabori S Sakurai S Izumi Y Fujii Y 1956 An Asymmetric Catalyst Nature 178 4528 323 Bibcode 1956Natur 178 323A doi 10 1038 178323b0 PMID 13358737 S2CID 4221816 a b c Noyori R 2003 Asymmetric Catalysis Science and Opportunities Nobel Lecture 2001 Advanced Synthesis amp Catalysis 345 12 15 41 doi 10 1002 adsc 200390002 a b c Knowles W S 2002 Asymmetric Hydrogenations Nobel Lecture Angewandte Chemie International Edition 41 12 1998 2007 doi 10 1002 1521 3773 20020617 41 12 lt 1998 AID ANIE1998 gt 3 0 CO 2 8 PMID 19746594 Crabtree Robert 1979 Iridium compounds in catalysis Accounts of Chemical Research 12 9 331 337 doi 10 1021 ar50141a005 Pfaltz A 2004 Asymmetric Catalysis Special Feature Part II Design of chiral ligands for asymmetric catalysis From C2 symmetric P P and N N ligands to sterically and electronically nonsymmetrical P N ligands Proceedings of the National Academy of Sciences 101 16 5723 5726 Bibcode 2004PNAS 101 5723P doi 10 1073 pnas 0307152101 PMC 395974 PMID 15069193 a b de Vries Johannes G Elsevier Cornelis J eds 2006 10 20 The Handbook of Homogeneous Hydrogenation 1 ed Wiley doi 10 1002 9783527619382 ISBN 978 3 527 31161 3 Gridnev I D Imamoto T 2004 On the Mechanism of Stereoselection in Rh Catalyzed Asymmetric Hydrogenation A General Approach for Predicting the Sense of Enantioselectivity Accounts of Chemical Research 37 9 633 644 doi 10 1021 ar030156e PMID 15379579 Wen Jialin Wang Fangyuan Zhang Xumu 2021 Asymmetric hydrogenation catalyzed by first row transition metal complexes Chemical Society Reviews 50 5 3211 3237 doi 10 1039 D0CS00082E ISSN 0306 0012 Gridnev Ilya D Imamoto Tsuneo 2004 09 01 On the Mechanism of Stereoselection in Rh Catalyzed Asymmetric Hydrogenation A General Approach for Predicting the Sense of Enantioselectivity Accounts of Chemical Research 37 9 633 644 doi 10 1021 ar030156e ISSN 0001 4842 Imamoto Tsuneo Tamura Ken Zhang Zhenfeng Horiuchi Yumi Sugiya Masashi Yoshida Kazuhiro Yanagisawa Akira Gridnev Ilya D 2012 01 25 Rigid P Chiral Phosphine Ligands with tert Butylmethylphosphino Groups for Rhodium Catalyzed Asymmetric Hydrogenation of Functionalized Alkenes Journal of the American Chemical Society 134 3 1754 1769 doi 10 1021 ja209700j ISSN 0002 7863 a b c d Dub Pavel A Gordon John C 2018 The role of the metal bound N H functionality in Noyori type molecular catalysts Nature Reviews Chemistry 2 12 396 408 doi 10 1038 s41570 018 0049 z S2CID 106394152 Noyori Ryōji Masashi Yamakawa Shohei Hashiguchi 2001 11 01 Metal Ligand Bifunctional Catalysis A Nonclassical Mechanism for Asymmetric Hydrogen Transfer between Alcohols and Carbonyl Compounds The Journal of Organic Chemistry 66 24 7931 7944 doi 10 1021 jo010721w PMID 11722188 Ohkuma T Ooka H Ikariya T Noyori R 1995 Preferential hydrogenation of aldehydes and ketones Journal of the American Chemical Society 117 41 10417 10418 doi 10 1021 ja00146a041 a b Ikariya T Blacker A J 2007 Asymmetric Transfer Hydrogenation of Ketones with Bifunctional Transition Metal Based Molecular Catalysts Accounts of Chemical Research 40 12 1300 1308 doi 10 1021 ar700134q PMID 17960897 a b Pilkington C Lennon I 2003 The Application of Asymmetric Hydrogenation for the Manufacture of Pharmaceutical Intermediates The Need for Catalyst Diversity Synthesis 2003 11 1639 doi 10 1055 s 2003 40871 a b c Noyori R 2002 Asymmetric Catalysis Science and Opportunities Nobel Lecture Angewandte Chemie International Edition 41 12 2008 22 doi 10 1002 1521 3773 20020617 41 12 lt 2008 aid anie2008 gt 3 0 co 2 4 PMID 19746595 Enthaler S Junge K Beller M 2008 Sustainable Metal Catalysis with Iron From Rust to a Rising Star Angewandte Chemie International Edition 47 18 3317 21 doi 10 1002 anie 200800012 PMID 18412184 Mikhailine A Lough A J Morris R H 2009 Efficient Asymmetric Transfer Hydrogenation of Ketones Catalyzed by an Iron Complex Containing a P N N P Tetradentate Ligand Formed by Template Synthesis Journal of the American Chemical Society 131 4 1394 1395 doi 10 1021 ja809493h PMID 19133772 Sonnenberg J F Coombs N Dube P A Morris R H 2012 Iron Nanoparticles Catalyzing the Asymmetric Transfer Hydrogenation of Ketones Journal of the American Chemical Society 134 13 5893 5899 doi 10 1021 ja211658t PMID 22448656 Whitesell J K 1989 C2 symmetry and asymmetric induction Chemical Reviews 89 7 1581 1590 doi 10 1021 cr00097a012 Knowles W S Sabacky M J Vineyard B D 1972 Catalytic asymmetric hydrogenation Journal of the Chemical Society Chemical Communications 214 1 119 124 doi 10 1039 C39720000010 PMID 4270504 a b Jerphagnon T Renaud J L Bruneau C 2004 Chiral monodentate phosphorus ligands for rhodium catalyzed asymmetric hydrogenation Tetrahedron Asymmetry 15 14 2101 doi 10 1016 j tetasy 2004 04 037 Van Den Berg M Minnaard A J Schudde E P Van Esch J De Vries A H M De Vries J G Feringa B L 2000 Highly Enantioselective Rhodium Catalyzed Hydrogenation with Monodentate Ligands PDF Journal of the American Chemical Society 122 46 11539 doi 10 1021 ja002507f hdl 11370 3c92d080 f024 45fe b997 b100634bd612 S2CID 95403641 Fu Y Xie J H Hu A G Zhou H Wang L X Zhou Q L 2002 Novel monodentate spiro phosphorus ligands for rhodium catalyzed hydrogenation reactions Chemical Communications 5 480 481 doi 10 1039 B109827F PMID 12120551 Reetz M T Sell T Meiswinkel A Mehler G 2003 A New Principle in Combinatorial Asymmetric Transition Metal Catalysis Mixtures of Chiral Monodentate P Ligands Angewandte Chemie International Edition 42 7 790 3 doi 10 1002 anie 200390209 PMID 12596201 Vineyard B D Knowles W S Sabacky M J Bachman G L Weinkauff D J 1977 Asymmetric hydrogenation Rhodium chiral bisphosphine catalyst Journal of the American Chemical Society 99 18 5946 doi 10 1021 ja00460a018 Knowles W S Sabacky M J Vineyard B D Weinkauff D J 1975 Asymmetric hydrogenation with a complex of rhodium and a chiral bisphosphine Journal of the American Chemical Society 97 9 2567 doi 10 1021 ja00842a058 Muller D Umbricht G Weber B Pfaltz A 1991 C2 Symmetric 4 4 5 5 Tetrahydrobi oxazoles and 4 4 5 5 Tetrahydro 2 2 methylenebis oxazoles as Chiral Ligands for Enantioselective Catalysis Preliminary Communication Helvetica Chimica Acta 74 232 240 doi 10 1002 hlca 19910740123 a b Helmchen G N Pfaltz A 2000 PhosphinooxazolinesA New Class of Versatile Modular P N Ligands for Asymmetric Catalysis Accounts of Chemical Research 33 6 336 345 doi 10 1021 ar9900865 PMID 10891051 Lightfoot A Schnider P Pfaltz A 1998 Enantioselective Hydrogenation of Olefins with Iridium Phosphanodihydrooxazole Catalysts Angewandte Chemie International Edition 37 20 2897 2899 doi 10 1002 SICI 1521 3773 19981102 37 20 lt 2897 AID ANIE2897 gt 3 0 CO 2 8 PMID 29711115 Franzke A Pfaltz A 2011 Zwitterionic Iridium Complexes with P N Ligands as Catalysts for the Asymmetric Hydrogenation of Alkenes Chemistry A European Journal 17 15 4131 44 doi 10 1002 chem 201003314 PMID 21381140 Maurer F Huch V Ullrich A Kazmaier U 2012 Development of Catalysts for the Stereoselective Hydrogenation of a b Unsaturated Ketones The Journal of Organic Chemistry 77 11 5139 5143 doi 10 1021 jo300246c PMID 22571628 Rageot D Woodmansee D H Pugin B T Pfaltz A 2011 Proline Based P O Ligand Iridium Complexes as Highly Selective Catalysts Asymmetric Hydrogenation of Trisubstituted Alkenes Angewandte Chemie International Edition 50 41 9598 601 doi 10 1002 anie 201104105 PMID 21882320 a b c Perry M C Cui X Powell M T Hou D R Reibenspies J H Burgess K 2003 Optically Active Iridium Imidazol 2 ylidene oxazoline Complexes Preparation and Use in Asymmetric Hydrogenation of Arylalkenes Journal of the American Chemical Society 125 1 113 123 doi 10 1021 ja028142b PMID 12515512 a b Nanchen S Pfaltz A 2006 Synthesis and Application of Chiral N Heterocyclic Carbene Oxazoline Ligands Iridium Catalyzed Enantioselective Hydrogenation Chemistry A European Journal 12 17 4550 8 doi 10 1002 chem 200501500 PMID 16557626 Zhu Y Burgess K 2008 Iridium Catalyzed Asymmetric Hydrogenation of Vinyl Ethers Advanced Synthesis amp Catalysis 350 7 8 979 doi 10 1002 adsc 200700546 Zhao J Burgess K 2009 Aldol Type Chirons from Asymmetric Hydrogenations of Trisubstituted Alkenes Organic Letters 11 10 2053 2056 doi 10 1021 ol900308w PMID 19368378 Zhao J Burgess K 2009 Synthesis of Vicinal Dimethyl Chirons by Asymmetric Hydrogenation of Trisubstituted Alkenes Journal of the American Chemical Society 131 37 13236 13237 doi 10 1021 ja905458n PMID 19719102 Zhou J Burgess K 2007 A w Functionalized 2 4 Dimethylpentane Dyads and 2 4 6 Trimethylheptane Triads through Asymmetric Hydrogenation Angewandte Chemie International Edition 46 7 1129 31 doi 10 1002 anie 200603635 PMID 17200966 Zhou J Zhu Y Burgess K 2007 Synthesis of S R R S R S 4 6 8 10 16 18 Hexamethyldocosane from Antitrogus parvulus via Diastereoselective Hydrogenations Organic Letters 9 7 1391 1393 doi 10 1021 ol070298z PMID 17338543 Urban S Ortega N Glorius F 2011 Ligand Controlled Highly Regioselective and Asymmetric Hydrogenation of Quinoxalines Catalyzed by Ruthenium N Heterocyclic Carbene Complexes Angewandte Chemie International Edition 50 16 3803 6 doi 10 1002 anie 201100008 PMID 21442699 Cui X Burgess K 2005 Catalytic Homogeneous Asymmetric Hydrogenations of Largely Unfunctionalized Alkenes Chemical Reviews 105 9 3272 3296 doi 10 1021 cr0500131 PMID 16159153 Pamies O Andersson P G Dieguez M 2010 Asymmetric Hydrogenation of Minimally Functionalised Terminal Olefins An Alternative Sustainable and Direct Strategy for Preparing Enantioenriched Hydrocarbons Chemistry A European Journal 16 48 14232 40 doi 10 1002 chem 201001909 PMID 21140401 Woodmansee D H Pfaltz A 2011 Asymmetric hydrogenation of alkenes lacking coordinating groups Chemical Communications 47 28 7912 7916 doi 10 1039 c1cc11430a PMID 21556431 Mazuela J Verendel J J Coll M SchaFfner B N BoRner A Andersson P G PaMies O DieGuez M 2009 Iridium Phosphite Oxazoline Catalysts for the Highly Enantioselective Hydrogenation of Terminal Alkenes Journal of the American Chemical Society 131 34 12344 12353 doi 10 1021 ja904152r PMID 19658416 Hou G H Xie J H Wang L X Zhou Q L 2006 Highly Efficient Rh I Catalyzed Asymmetric Hydrogenation of Enamines Using Monodente Spiro Phosphonite Ligands Journal of the American Chemical Society 128 36 11774 11775 doi 10 1021 ja0644778 PMID 16953614 Hou G H Xie J H Yan P C Zhou Q L 2009 Iridium Catalyzed Asymmetric Hydrogenation of Cyclic Enamines Journal of the American Chemical Society 131 4 1366 1367 doi 10 1021 ja808358r PMID 19132836 Noyori R Ohkuma T 2001 Asymmetric Catalysis by Architectural and Functional Molecular Engineering Practical Chemo and Stereoselective Hydrogenation of Ketones Angewandte Chemie International Edition 40 1 40 73 doi 10 1002 1521 3773 20010105 40 1 lt 40 AID ANIE40 gt 3 0 CO 2 5 PMID 11169691 Hems W P Groarke M Zanotti Gerosa A Grasa G A 2007 Bisphosphine Ru II Diamine Complexes in Asymmetric Hydrogenation Expanding the Scope of the Diamine Ligand Accounts of Chemical Research 40 12 1340 1347 doi 10 1021 ar7000233 PMID 17576143 Noyori R Yamakawa M Hashiguchi S 2001 Metal Ligand Bifunctional Catalysis A Nonclassical Mechanism for Asymmetric Hydrogen Transfer between Alcohols and Carbonyl Compounds The Journal of Organic Chemistry 66 24 7931 7944 doi 10 1021 jo010721w PMID 11722188 Zhu S F Xie J B Zhang Y Z Li S Zhou Q L 2006 Well Defined Chiral Spiro Iridium Phosphine Oxazoline Cationic Complexes for Highly Enantioselective Hydrogenation of Imines at Ambient Pressure Journal of the American Chemical Society 128 39 12886 12891 doi 10 1021 ja063444p PMID 17002383 Xie J H Liu X Y Xie J B Wang L X Zhou Q L 2011 An Additional Coordination Group Leads to Extremely Efficient Chiral Iridium Catalysts for Asymmetric Hydrogenation of Ketones Angewandte Chemie International Edition 50 32 7329 32 doi 10 1002 anie 201102710 PMID 21751315 Ohkuma T Ooka H Yamakawa M Ikariya T Noyori R 1996 Stereoselective Hydrogenation of Simple Ketones Catalyzed by Ruthenium II Complexes The Journal of Organic Chemistry 61 15 4872 4873 doi 10 1021 jo960997h Noyori R Ohkuma T 2001 Asymmetric Catalysis by Architectural and Functional Molecular Engineering Practical Chemo and Stereoselective Hydrogenation of Ketones Angewandte Chemie International Edition 40 1 40 73 doi 10 1002 1521 3773 20010105 40 1 lt 40 aid anie40 gt 3 0 co 2 5 PMID 11169691 a b Zhou Y G 2007 Asymmetric Hydrogenation of Heteroaromatic Compounds Accounts of Chemical Research 40 12 1357 1366 CiteSeerX 10 1 1 653 5495 doi 10 1021 ar700094b PMID 17896823 Wang W B Lu S M Yang P Y Han X W Zhou Y G 2003 Highly Enantioselective Iridium Catalyzed Hydrogenation of Heteroaromatic Compounds Quinolines Journal of the American Chemical Society 125 35 10536 10537 CiteSeerX 10 1 1 651 3119 doi 10 1021 ja0353762 PMID 12940733 a b Xu L Lam K H Ji J Wu J Fan Q H Lo W H Chan A S C 2005 Air stable Ir P Phos complex for highly enantioselective hydrogenation of quinolines and their immobilization in poly ethylene glycol dimethyl ether DMPEG Chemical Communications 11 1390 2 doi 10 1039 B416397D hdl 10397 8906 PMID 15756313 Lam K H Xu L Feng L Fan Q H Lam F L Lo W H Chan A S C 2005 Highly Enantioselective Iridium Catalyzed Hydrogenation of Quinoline Derivatives Using Chiral Phosphinite H8 BINAPO Advanced Synthesis amp Catalysis 347 14 1755 doi 10 1002 adsc 200505130 hdl 10397 26878 Wang Z J Deng G J Li Y He Y M Tang W J Fan Q H 2007 Enantioselective Hydrogenation of Quinolines Catalyzed by Ir BINAP Cored Dendrimers Dramatic Enhancement of Catalytic Activity Organic Letters 9 7 1243 1246 doi 10 1021 ol0631410 PMID 17328554 Qiu L Kwong F Y Wu J Lam W H Chan S Yu W Y Li Y M Guo R Zhou Z Chan A S C 2006 A New Class of Versatile Chiral Bridged Atropisomeric Diphosphine Ligands Remarkably Efficient Ligand Syntheses and Their Applications in Highly Enantioselective Hydrogenation Reactions Journal of the American Chemical Society 128 17 5955 5965 doi 10 1021 ja0602694 hdl 10397 60397 PMID 16637664 Reetz M T Li X 2006 Asymmetric hydrogenation of quinolines catalyzed by iridium complexes of BINOL derived diphosphonites Chemical Communications 20 2159 60 doi 10 1039 b602320g PMID 16703140 Rueping Antonchick A Theissmann T 2006 A highly enantioselective Bronsted acid catalyzed cascade reaction organocatalytic transfer hydrogenation of quinolines and their application in the synthesis of alkaloids Angewandte Chemie International Edition in English 45 22 3683 3686 doi 10 1002 anie 200600191 PMID 16639754 Tang W Xu L Fan Q H Wang J Fan B Zhou Z Lam K H Chan A S C 2009 Asymmetric Hydrogenation of Quinoxalines with Diphosphinite Ligands A Practical Synthesis of Enantioenriched Substituted Tetrahydroquinoxalines Angewandte Chemie International Edition 48 48 9135 8 doi 10 1002 anie 200904518 hdl 10397 20432 PMID 19876991 Rueping M Tato F Schoepke F R 2010 The First General Efficient and Highly Enantioselective Reduction of Quinoxalines and Quinoxalinones Chemistry A European Journal 16 9 2688 91 doi 10 1002 chem 200902907 PMID 20140920 Glorius F Spielkamp N Holle S Goddard R Lehmann C W 2004 Efficient Asymmetric Hydrogenation of Pyridines Angewandte Chemie International Edition 43 21 2850 2 doi 10 1002 anie 200453942 PMID 15150766 Ye Z S Chen M W Chen Q A Shi L Duan Y Zhou Y G 2012 Iridium Catalyzed Asymmetric Hydrogenation of Pyridinium Salts Angewandte Chemie International Edition 51 40 10181 4 doi 10 1002 anie 201205187 PMID 22969060 Tang W J Tan J Xu L J Lam K H Fan Q H Chan A S C 2010 Highly Enantioselective Hydrogenation of Quinoline and Pyridine Derivatives with Iridium P Phos Catalyst Advanced Synthesis amp Catalysis 352 6 1055 doi 10 1002 adsc 200900870 hdl 10397 22884 Rueping M Antonchick A P 2007 Organocatalytic Enantioselective Reduction of Pyridines Angewandte Chemie International Edition 46 24 4562 5 doi 10 1002 anie 200701158 PMID 17492817 Baeza A Pfaltz A 2010 Iridium Catalyzed Asymmetric Hydrogenation of N Protected Indoles Chemistry A European Journal 16 7 2036 9 doi 10 1002 chem 200903105 PMID 20104554 Xiao Y C Wang C Yao Y Sun J Chen Y C 2011 Direct Asymmetric Hydrosilylation of Indoles Combined Lewis Base and Bronsted Acid Activation Angewandte Chemie International Edition 50 45 10661 4 doi 10 1002 anie 201105341 PMID 21932274 Duan Y Chen M W Ye Z S Wang D S Chen Q A Zhou Y G 2011 An Enantioselective Approach to 2 3 Disubstituted Indolines through Consecutive Bronsted Acid Pd Complex Promoted Tandem Reactions Chemistry A European Journal 17 26 7193 7 doi 10 1002 chem 201100576 PMID 21567504 Wang D S Ye Z S Chen Q A Zhou Y G Yu C B Fan H J Duan Y 2011 Highly Enantioselective Partial Hydrogenation of Simple Pyrroles A Facile Access to Chiral 1 Pyrrolines Journal of the American Chemical Society 133 23 8866 8869 doi 10 1021 ja203190t PMID 21591641 Wang D S Chen Q A Lu S M Zhou Y G 2012 Asymmetric Hydrogenation of Heteroarenes and Arenes Chemical Reviews 112 4 2557 2590 doi 10 1021 cr200328h PMID 22098109 Urban S Beiring B Ortega N Paul D Glorius F 2012 Asymmetric Hydrogenation of Thiophenes and Benzothiophenes Journal of the American Chemical Society 134 37 15241 15244 doi 10 1021 ja306622y PMID 22934527 a b Heitbaum M Glorius F Escher I 2006 Asymmetric Heterogeneous Catalysis Angewandte Chemie International Edition 45 29 4732 62 doi 10 1002 anie 200504212 PMID 16802397 Yoon M Srirambalaji R Kim K 2012 Homochiral Metal Organic Frameworks for Asymmetric Heterogeneous Catalysis Chemical Reviews 112 2 1196 1231 doi 10 1021 cr2003147 PMID 22084838 Hu A Ngo H L Lin W 2003 Chiral Porous Hybrid Solids for Practical Heterogeneous Asymmetric Hydrogenation of Aromatic Ketones Journal of the American Chemical Society 125 38 11490 11491 doi 10 1021 ja0348344 PMID 13129339 Blaser H U Spindler F Studer M 2001 Enantioselective catalysis in fine chemicals production Applied Catalysis A General 221 1 2 119 143 doi 10 1016 S0926 860X 01 00801 8 PMID 12613584 Blaser Hans Ulrich Federsel Hans Jurgen eds 2010 Asymmetric Catalysis on Industrial Scale Weinheim Wiley VCH pp 13 16 doi 10 1002 9783527630639 ISBN 978 3 527 63063 9 Jacobsen E N Pfaltz Andreas Yamamato H eds 1999 Comprehensive Asymmetric Catalysis Berlin New York Springer pp 1443 1445 ISBN 978 3 540 64336 4 Retrieved from https en wikipedia org w index php title Asymmetric hydrogenation amp oldid 1189389640, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.