Nitro-Mannich reaction

Last updated

The nitro-Mannich reaction (or aza-Henry reaction) is the nucleophilic addition of a nitroalkane (or the corresponding nitronate anion) to an imine, resulting in the formation of a beta-nitroamine. [1] With the reaction involving the addition of an acidic carbon nucleophile to a carbon-heteroatom double bond, the nitro-Mannich reaction is related to some of the most fundamental carbon-carbon bond forming reactions in organic chemistry, including the aldol reaction, [2] Henry reaction (nitro-aldol reaction) [3] and Mannich reaction. [4]

Contents

Summary of enolate and nitronate nucleophilic additions to imines and aldehydes/ketones. Summary of nucleophilic additions to imines and aldehydes or ketones.svg
Summary of enolate and nitronate nucleophilic additions to imines and aldehydes/ketones.

Although extensive research has been conducted into the aforementioned reactions, the nitro-Mannich reaction has been studied to a far lesser extent even though it has been known for well over 100 years. [5] Significant attention only started to develop after the report of Anderson and co-workers at the turn of the century, [6] and has since resulted in a wide range of novel methodologies. The interest into the nitro-Mannich reaction stems from the synthetic utility of the beta-nitroamine products. They can be further manipulated by various methods, including reductive removal of the nitro group allowing access to monoamines, reduction of the nitro group affords 1,2-diamines and conversion of the nitro group into a carbonyl functionality furnishes beta-aminocarbonyl compounds.

Selected examples of the synthetic utility of beta-nitroamines synthesised via nitro- Mannich reactions. Examples of the synthetic utility of beta-nitroamines.svg
Selected examples of the synthetic utility of beta-nitroamines synthesised via nitro- Mannich reactions.

History

Early Examples of the Nitro-Mannich Reaction

The first nitro-Mannich reaction was reported by Henry in 1896. [5] In this report, Henry described the addition of nitroalkanes to an imine derived from hemiaminal. Elimination of water forms in-situ an imine, which then reacts with the nitro group (as a nitronate ion) to form a beta-nitroamine that can subsequently react further forming one of the two adducts. Although this is the first report of the nitro-Mannich reaction, no yields of the products were given.

First report of the nitro-Mannich reaction.svg

After Henry’s seminal report, Mousset [7] and Duden [8] made contributions to the field by studying the addition of branched nitroalkanes to hemiaminals using the same procedures reported by Henry. An example of nitro group reduction to an amine using SnCl2 and HCl was also disclosed by Duden and co-workers, thus representing the first use of the nitro-Mannich reaction to prepare polyamines. The next report did not appear until 1931, [9] when Cerf de Mauny conducted a thorough study of Henry’s original work using hemiaminals. The scope of the reaction was extended to higher order nitroalkanes affording a beta-nitroamine in excellent yields.

Cerf de Mauny's nitro-Mannich reaction.svg

The next contributions appeared in 1946, when Senkus and Johnson independently reported their studies into the nitro-Mannich reaction. Senkus and co-workers [10] illustrated that nitroalkanes may react with methanal (formaldehyde) and substituted primary amines in the presence of sodium sulfate (Na2SO4) to afford a variety of substituted beta-nitroamines in moderate to good yields. When using primary nitroalkane substrates, double addition of the nitroalkane to the imine was observed, but this could be avoided by employing secondary nitroalkanes. The study reported by Johnson and co-workers [11] also employed formaldehyde, but this was used in conjunction with a selection of secondary amines, furnishing the corresponding beta-nitroamines in moderate to good yields. Both authors also reduced the nitro group to an amine functionality using Raney Nickel.

Nitro-Mannich reactions reported by Senkus and Johnson.svg

Up until this point, all of the nitro-Mannich methodologies reported had used imines that were formed in situ from an aldehyde and an amine. In 1950, Hurd and Strong reported [12] the first nitro-Mannich reaction using a preformed imine. Exposing an imine to a nitroalkane afforded a substituted beta-nitroamines in moderate yields. The moderate yields obtained when using the preformed imine could possibly be attributed to a competing decomposition pathway of the imine or the product.

First nitro-Mannich reactions using preformed imines.svg

These early nitro-Mannich methodologies have been used by a number of groups for the synthesis of a variety of heterocyclic products, conjugated nitroalkenes (via elimination of the amino group) [13] [14] and dinitroamines. [15]

Non-Enantioselective Nitro-Mannich Reactions

Although the nitro-Mannich reaction enables access to synthetically useful beta-nitroamine motifs, the lack of selectivity in their synthesis remained a significant problem. Interest in the field started to increase considerably after Anderson and co-workers reported the first diastereoselective acyclic nitro-Mannich reaction. [6] A nitroalkane and n-butyllithium (nBuLi) were combined at -78 °C to give the corresponding nitronate ions. A selection of N-PMB imines were then added to the reaction mixture and after quenching with acetic acid, the beta-nitroamine products were afforded in good yields with moderate to good diastereoselectivities.

Anderson's diastereoselective nitro-Mannich reaction.svg

The authors then converted the beta-nitroamines into unprotected 1,2-diamines via a two step procedure. Firstly, the nitro group was reduced to amines using samarium iodide, followed by PMB removal in the presence of ceric ammonium nitrate (CAN). The same group later reported improvements to this methodology and expanded these preliminary results in further publications. [16] [17]

In 2000, Anderson and co-workers reported the racemic nitro-Mannich reaction of TMS-protected nitronate with N-PMB or N-PMP imines catalysed by Sc(OTf)3. [18] The authors first attempted the nitro-Mannich reaction using lithium-nitronates, however no product was formed using these conditions. As a result, the TMS-protected nitronate was used in conjunction with Scandium(III) trifluoromethanesulfonate [Sc(OTf)3] (4 mol%) to afford the beta-nitroamine products in moderate to excellent yields for a range of alkyl and aryl N-PMB and N-PMP protected imines.

Anderson's Sc(OTf)3 catalysed nitro-Mannich reaction.svg

Following Anderson’s report, Qian and co-workers described the Ytterbium(III) trifluoromethanesulfonate [Yb(OiPr)3] catalysed nitro-Mannich reaction of N-sulfonyl imines and nitromethane. [19] Using mild reactions conditions, the β-nitroamines bearing electron-rich and electron-poor aryl substituents were furnished in excellent yields after short reaction times.

Qian's Yb(OiPr)3 catalysed nitro-Mannich reaction.svg

Direct Metal Catalysed Enantioselective Nitro-Mannich Reactions

The first enantioselective metal catalysed nitro-Mannich reaction was reported by Shibasaki and co-workers in 1999. [20] The authors used a binaphthol ligated Yb/K heterobimetallic complex to induce enantiocontrol in the reaction, furnishing β-nitroamines in moderate to good yields with good enantioselectivities. However, nitromethane was the only nitroalkane that could be used with the heterobimetallic complex and the reactions were very slow (2.5–7 days) even when using a relatively high catalyst loading of 20 mol%.

Shibasaki's enantioselective nitro-Mannich reaction.svg

Building on the work of Shibasaki, Jørgensen and co-workers reported the asymmetric nitro-Mannich reaction of nitroalkanes and a N-PMP-α-iminoesters. [21] Catalysed by Cu(II)-BOX 52 and triethylamine (Et3N), the reaction afforded β-nitro-α-aminoesters in good yields with excellent enantiocontrol (up to 99% ee). The reaction tolerates a selection of nitroalkanes but is limited exclusively to N-PMP-α-iminoesters. The authors propose that the reaction proceeds via the chair-like transition structure, where both the N-PMP-α-iminoester and the nitronate anion bind to the Cu(II)-BOX complex.

Jorgensen's enantioselective nitro-Mannich reaction.svg

In 2007, Feng and co-workers reported that CuOTf used in conjunction with the shown chiral N-oxide ligand and DIPEA is an efficient catalytic system for the enantioselective nitro-Mannich reaction of nitromethane with N-sulfonyl imines. [22] Combining all of the reagents in THF at –40 °C resulted in the formation of β-nitroamines in excellent yields (up to 99%) and good enantioselectivities for a variety of substituted aryls groups. The postulated intermediate complex is similar to the transition structure proposed by Jørgensen and co-workers, where the ligated copper species binds to the N-sulfonyl imine. A hydrogen bonding interaction is proposed to exist between the amide NH and the nitronate species.

Feng's enantioselective copper catalysed nitro-Mannich reaction.svg

Around the same time as the report of Feng, Shibasaki and co-workers reported one of the most successful enantioselective nitro-Mannich reactions, catalysed by the shown Cu/Sm heterobimetallic complex. [23] Combining N-Boc protected imines and nitroalkanes resulted in moderate to excellent yields and good to excellent enantioselectivities of the products. Interestingly, the nitro-Mannich reaction catalysed by complex affords syn-β-nitroamines, whereas most other enantioselective methodologies favour anti-β-nitroamines. The authors later reported an improved version of the protocol and proposed a mechanistic rational to account for the observed syn diastereoselectivity. [24]

Shibasaki's Cu Sm heterobimetallic complex for enantioselective nitro-Mannich reactions.svg

Organocatalysed Enantioselective Nitro-Mannich Reactions

Since the inception of organocatalysis, numerous accounts of organocatalysed enantioselective nitro-Mannich reactions have been reported. [1] These include examples using Brønsted base catalysts, Brønsted acid catalysts, bifunctional Brønsted base/H-bond donor catalysts and phase-transfer catalysts.

Bifunctional Brønsted Base/H-Bond Donor Organocatalysis

Small chiral molecule H-bond donors can be used as a powerful tool for enantioselective synthesis. [25] These low molecular weight entities containing structural frameworks with distinct H-bond donor motifs can catalyse a wide range of carbon-carbon and carbon-heteroatom bond-forming reactions, occurring via H-bond donor activation of the reaction partners as well as through organisation of their spatial arrangement. This area of organic chemistry received limited attention until the seminal work of Jacobsen and Sigman [26] in which they reported a highly enantioselective Strecker reaction using a H-bond donor organocatalyst:

Jacobsen's thiourea organocatalyst.svg

Building on the work of Jacobsen, it was recognised that H-bond donor motifs can be linked via a chiral scaffold to Brønsted basic moieties, creating a new class of bifunctional organocatalysts (see concept figure below). The incorporation of these two functionalities allows the simultaneous activation of the nucleophile (via deprotonation by the Brønsted base) and electrophile (via H-bond donation), thus allowing the development of novel enantioselective reactions through new activation modes.

Concept Bronsted organocatalysis.svg

Based on this concept, Takemoto and co-workers reported the first bifunctional Brønsted base/H-bond donor thiourea organocatalyst 62 (see below) in 2003. [27] This organocatalyst, based on the 1,2-trans-cyclohexanediamine scaffold, imparts high levels of enantiocontrol in the Michael addition of dimethylmalonate to a variety of nitrostyrenes. After this seminal report, numerous other bifunctional organocatalysts were developed derived from the readily available cinchona alkaloid scaffold. The quinidine-derived bifunctional organocatalyst 63 (first reported by Deng and co-workers) acts as a proficient catalyst for Michael addition reactions. [28] In this organocatalytic system, the H-bonding interaction arising from the quinoline alcohol is thought to be crucial for achieving high enantioselectivities.

Selection of pioneering organocatalysts.svg

Also the bifunctional thioureas 64 and 65, again derived from the cinchona alkaloids, are very effective catalysts in Michael addition reactions. [29] [30] [31] [32] The bifunctional thiourea 66 is able to impart high levels of enantiocontrol in the nitro-aldol (Henry) reaction. [33] Bifunctional thiourea 66 differs structurally from bifunctional thioureas 64 and 65, as the thiourea moiety is attached to the quinoline ring of the cinchona scaffold instead of the central stereocentre. Also numerous other bifunctional organocatalyst systems are described, which further expand the range of reactions that can be conducted using bifunctional (thio)urea organocatalysis. [25]

Related Research Articles

The Michael reaction or Michael addition is the nucleophilic addition of a carbanion or another nucleophile to an α,β-unsaturated carbonyl compound containing an electron withdrawing group. It belongs to the larger class of conjugate additions. This is one of the most useful methods for the mild formation of C–C bonds. Many asymmetric variants exist.

In chemistry hydrocyanation is a process for conversion of alkenes to nitriles. The reaction involves the addition of hydrogen cyanide and requires a catalyst. This conversion is conducted on an industrial scale for the production of precursors to nylon.

The Mannich reaction is an organic reaction which consists of an amino alkylation of an acidic proton placed next to a carbonyl functional group by formaldehyde and a primary or secondary amine or ammonia. The final product is a β-amino-carbonyl compound also known as a Mannich base. Reactions between aldimines and α-methylene carbonyls are also considered Mannich reactions because these imines form between amines and aldehydes. The reaction is named after chemist Carl Mannich.

Henry reaction

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

The aza-Baylis–Hillman reaction or aza-BH reaction in organic chemistry is a variation of the Baylis–Hillman reaction and describes the reaction of an electron deficient alkene, usually an α,β-unsaturated carbonyl compound, with an imine in the presence of a nucleophile. The reaction product is an allylic amine. The reaction can be carried out in enantiomeric excess of up to 90% with the aid of bifunctional chiral BINOL and phosphinyl BINOL compounds, for example in the reaction of n-(4-chloro-benzylidene)-benzenesulfonamide with methyl vinyl ketone (MVK) in cyclopentyl methyl ether and toluene at -15°C.

Petasis reaction

The Petasis reaction is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

Transfer hydrogenation is the addition of hydrogen (H2; dihydrogen in inorganic and organometallic chemistry) to a molecule from a source other than gaseous H2. It is applied in industry and in organic synthesis, in part because of the inconvenience and expense of using gaseous H2. One large scale application of transfer hydrogenation is coal liquefaction using "donor solvents" such as tetralin.

Organocatalysis

In organic chemistry, the term organocatalysis refers to a form of catalysis, whereby the rate of a chemical reaction is increased by an organic catalyst referred to as an "organocatalyst" consisting of carbon, hydrogen, sulfur and other nonmetal elements found in organic compounds. Because of their similarity in composition and description, they are often mistaken as a misnomer for enzymes due to their comparable effects on reaction rates and forms of catalysis involved.

A carbometalation is any reaction where a carbon-metal bond reacts with a carbon-carbon π-bond to produce a new carbon-carbon σ-bond and a carbon-metal σ-bond. The resulting carbon-metal bond can undergo further carbometallation reactions or it can be reacted with a variety of electrophiles including halogenating reagents, carbonyls, oxygen, and inorganic salts to produce different organometallic reagents. Carbometalations can be performed on alkynes and alkenes to form products with high geometric purity or enantioselectivity, respectively. Some metals prefer to give the anti-addition product with high selectivity and some yield the syn-addition product. The outcome of syn and anti- addition products is determined by the mechanism of the carbometalation.

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 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 awarded one half of the 2001 Nobel Prize in Chemistry.

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

Within the area of organocatalysis, (thio)urea organocatalysis describes the use of ureas and thioureas to accelerate and stereochemically alter organic transformations. The effects arise through hydrogen-bonding interactions between the substrate and the (thio)urea. Unlike classical catalysts, these organocatalysts interact by non-covalent interactions, especially hydrogen bonding. The scope of these small-molecule H-bond donors termed (thio)urea organocatalysis covers both non-stereoselective and stereoselective applications.

Lectka enantioselective beta-lactam synthesis

Lectka and co-workers developed a catalytic, asymmetric method to synthesize β-lactams.

The Baylis–Hillman reaction is a carbon-carbon bond forming reaction between the α-position of an activated alkene and a carbon electrophile such as an aldehyde. Employing a nucleophilic catalyst, such as a tertiary amine and phosphine, this reaction provides a densely functionalized product. It is named for Anthony B. Baylis and Melville E. D. Hillman, two of the chemists who developed this reaction while working at Celanese. This reaction is also known as the Morita–Baylis–Hillman reaction or MBH reaction, as Morita had published earlier work on it.

Hydrogen-bond catalysis

Hydrogen-bond catalysis is a type of organocatalysis that relies on use of hydrogen bonding interactions to accelerate and control organic reactions. In biological systems, hydrogen bonding plays a key role in many enzymatic reactions, both in orienting the substrate molecules and lowering barriers to reaction. However, chemists have only recently attempted to harness the power of using hydrogen bonds to perform catalysis, and the field is relatively undeveloped compared to research in Lewis acid catalysis.

Synergistic catalysis

Synergistic catalysis is a specialized approach to catalysis whereby at least two different catalysts act on two different substrates simultaneously to allow reaction between the two activated materials. While a catalyst works to lower the energy of reaction overall, a reaction using synergistic catalysts work together to increase the energy level of HOMO of one of the molecules and lower the LUMO of another. While this concept has come to be important in developing synthetic pathways, this strategy is commonly found in biological systems as well.

Proline organocatalysis is the use of proline as an organocatalyst in organic chemistry. This theme is often considered the starting point for the area of organocatalysis, even though early discoveries went unappreciated. Modifications, such as MacMillan’s catalyst and Jorgensen's catalysts, proceed with excellent stereocontrol.

The Roskamp reaction was first discovered by Eric J. Roskamp and co-workers in 1989. This reaction is very useful in synthesizing β-keto esters from aldehydes and diazoacetate, using various Lewis acids as catalysts (such as BF3, SnCl2, GeCl2).

The Crabbé reaction is an organic reaction that converts a terminal alkyne and aldehyde into an allene in the presence of a soft Lewis acid catalyst and secondary amine. Given continued developments in scope and generality, it is a convenient and increasingly important method for the preparation of allenes, a class of compounds often viewed as exotic and synthetically challenging to access.

Mizoroki-Heck vs. Reductive Heck

The Mizoroki−Heck coupling of aryl halides and alkenes to form C(sp2)–C(sp2) bonds has become a staple transformation in organic synthesis, owing to its broad functional group compatibility and varied scope. In stark contrast, the palladium-catalyzed reductive Heck reaction has received considerably less attention, despite the fact that early reports of this reaction date back almost half a century. From the perspective of retrosynthetic logic, this transformation is highly enabling because it can forge alkyl–aryl linkages from widely available alkenes, rather than from the less accessible and/or more expensive alkyl halide or organometallic C(sp3) synthons that are needed in a classical aryl/alkyl cross-coupling.

References

CC-BY-icon-80x15.png  This article incorporates text by David Michael Barber available under the CC BY 2.5 license.

  1. 1 2 Noble, Adam; Anderson, James C. (2013-05-08). "Nitro-Mannich Reaction". Chemical Reviews. 113 (5): 2887–2939. doi:10.1021/cr300272t. ISSN   0009-2665.
  2. R. Mahrwald, Modern Aldol Reactions; Wiley-VCH, Weinheim, 2004.
  3. Luzzio, Frederick A. (February 2001). "The Henry reaction: recent examples". Tetrahedron. 57 (6): 915–945. doi:10.1016/S0040-4020(00)00965-0.
  4. Kobayashi, Shu̅; Mori, Yuichiro; Fossey, John S.; Salter, Matthew M. (2011-04-13). "Catalytic Enantioselective Formation of C−C Bonds by Addition to Imines and Hydrazones: A Ten-Year Update". Chemical Reviews. 111 (4): 2626–2704. doi:10.1021/cr100204f. ISSN   0009-2665.
  5. 1 2 L. Henry, Bull. Acad. Roy. Belg. 1896, 32, 33.
  6. 1 2 Adams, Harry; Anderson, James C.; Peace, Simon; Pennell, Andrew M. K. (December 1998). "The Nitro-Mannich Reaction and Its Application to the Stereoselective Synthesis of 1,2-Diamines †". The Journal of Organic Chemistry. 63 (26): 9932–9934. doi:10.1021/jo981700d. ISSN   0022-3263.
  7. T. Mousset, Bull. Acad. Roy. Belg. 1901, 37, 622.
  8. P. Duden, K. Bock, H. J. Reid, Chem. Ber. 1905, 33, 2036.
  9. (a) H. Cerf de Mauny, Bull. Soc. Chim. 1931, 4, 1451; (b) H. Cerf de Mauny, Bull. Soc. Chim. 1931, 4, 1460.
  10. Senkus, Murray (January 1946). "Reaction of Primary Aliphatic Amines with Formaldehyde and Nitroparaffins *". Journal of the American Chemical Society. 68 (1): 10–12. doi:10.1021/ja01205a004. ISSN   0002-7863.
  11. (a) H. G. Johnson, J. Am. Chem. Soc. 1946, 68, 12; (b) H. G. Johnson, J. Am. Chem. Soc. 1946, 68, 14.
  12. Hurd, Charles D.; Strong, James S. (October 1950). "Condensation of Nitroalkanes with Anils". Journal of the American Chemical Society. 72 (10): 4813–4814. doi:10.1021/ja01166a510. ISSN   0002-7863.
  13. Senkus, Murray (August 1946). "The Preparation of Some Hexahydropyrimidines from Nitroparaffins 1". Journal of the American Chemical Society. 68 (8): 1611–1613. doi:10.1021/ja01212a073. ISSN   0002-7863.
  14. Hirst, E. L.; Jones, J. K. N.; Minahan, S.; Ochynski, F. W.; Thomas, A. T.; Urbanski, T. (1947). "172. The reaction of 1-nitropropane with formaldehyde and ammonia". Journal of the Chemical Society (Resumed): 924. doi:10.1039/jr9470000924. ISSN   0368-1769.
  15. Smiley, Robert A. (August 1958). "Dinitro Amines and Their Reduction Products from the Mannich Reaction of Nitrocyclohexane with Aliphatic Diamines or Ammonia". The Journal of Organic Chemistry. 23 (8): 1115–1117. doi:10.1021/jo01102a008. ISSN   0022-3263.
  16. Anderson, James C.; Blake, Alexander J.; Howell, Gareth P.; Wilson, Claire (January 2005). "Scope and Limitations of the Nitro-Mannich Reaction for the Stereoselective Synthesis of 1,2-Diamines". The Journal of Organic Chemistry. 70 (2): 549–555. doi:10.1021/jo048304h. ISSN   0022-3263.
  17. Anderson, James; Chapman, Helen (2006-09-05). "Aluminum Amalgam for the Reduction of Sensitive β -Nitroamines to 1,2-Diamines". Synthesis. 2006 (19): 3309–3315. doi:10.1055/s-2006-950229. ISSN   0039-7881.
  18. Anderson, James C.; Peace, Simon; Pih, Steven (2000-10-10). "ChemInform Abstract: The Lewis Acid Catalyzed Addition of 1-Trimethylsilyl Nitropropanate to Imines". ChemInform. 31 (41): no–no. doi:10.1002/chin.200041018.
  19. Qian, Changtao; Gao, Feifeng; Chen, Ruifang (July 2001). "Yb(OiPr)3, a highly efficient catalyst for the nitro-Mannich reaction". Tetrahedron Letters. 42 (28): 4673–4675. doi:10.1016/S0040-4039(01)00786-9.
  20. Yamada, Ken-ichi; Harwood, Simon J.; Gröger, Harald; Shibasaki, Masakatsu (1999). "The First Catalytic Asymmetric Nitro-Mannich-Type Reaction Promoted by a New Heterobimetallic Complex". Angewandte Chemie International Edition. 38 (23): 3504–3506. doi:10.1002/(SICI)1521-3773(19991203)38:23<3504::AID-ANIE3504>3.0.CO;2-E. ISSN   1521-3773.
  21. Nishiwaki, Nagatoshi; Knudsen, Kristian Rahbek; Gothelf, Kurt V.; Jørgensen, Karl Anker (2001). "Catalytic Enantioselective Addition of Nitro Compounds to Imines—A Simple Approach for the Synthesis of Optically Active β-Nitro-α-Amino Esters". Angewandte Chemie International Edition. 40 (16): 2992–2995. doi:10.1002/1521-3773(20010817)40:16<2992::AID-ANIE2992>3.0.CO;2-3. ISSN   1521-3773.
  22. Zhou, Hui; Peng, Dan; Qin, Bo; Hou, Zongrui; Liu, Xiaohua; Feng, Xiaoming (2007-12-01). "Highly Enantioselective Aza-Henry Reaction of N -Tosyl Imines Catalyzed by N,N '-Dioxide−Cu(I) Complexes". The Journal of Organic Chemistry. 72 (26): 10302–10304. doi:10.1021/jo702200u. ISSN   0022-3263.
  23. Handa, Shinya; Gnanadesikan, Vijay; Matsunaga, Shigeki; Shibasaki, Masakatsu (April 2007). "syn -Selective Catalytic Asymmetric Nitro-Mannich Reactions Using a Heterobimetallic Cu−Sm−Schiff Base Complex". Journal of the American Chemical Society. 129 (16): 4900–4901. doi:10.1021/ja0701560. ISSN   0002-7863.
  24. Handa, Shinya; Gnanadesikan, Vijay; Matsunaga, Shigeki; Shibasaki, Masakatsu (2010-04-07). "Heterobimetallic Transition Metal/Rare Earth Metal Bifunctional Catalysis: A Cu/Sm/Schiff Base Complex for Syn -Selective Catalytic Asymmetric Nitro-Mannich Reaction". Journal of the American Chemical Society. 132 (13): 4925–4934. doi:10.1021/ja100514y. ISSN   0002-7863.
  25. 1 2 Doyle, Abigail G.; Jacobsen, Eric N. (2007-12-01). "Small-Molecule H-Bond Donors in Asymmetric Catalysis". Chemical Reviews. 107 (12): 5713–5743. doi:10.1021/cr068373r. ISSN   0009-2665.
  26. Sigman, Matthew S.; Jacobsen, Eric N. (May 1998). "Schiff Base Catalysts for the Asymmetric Strecker Reaction Identified and Optimized from Parallel Synthetic Libraries". Journal of the American Chemical Society. 120 (19): 4901–4902. doi:10.1021/ja980139y. ISSN   0002-7863.
  27. (a) T. Okino, Y. Hoashi, Y. Takemoto, J. Am. Chem. Soc.2003, 125, 12672; (b) T. Okino, Y. Hoashi, T. Furukawa, X. Xu, Y. Takemoto, J. Am. Chem. Soc.2005, 127, 119.
  28. Li, Hongming; Wang, Yi; Tang, Liang; Deng, Li (2004-08-01). "Highly Enantioselective Conjugate Addition of Malonate and β-Ketoester to Nitroalkenes: Asymmetric C−C Bond Formation with New Bifunctional Organic Catalysts Based on Cinchona Alkaloids". Journal of the American Chemical Society. 126 (32): 9906–9907. doi:10.1021/ja047281l. ISSN   0002-7863.
  29. Ye, Jinxing; Dixon, Darren J.; Hynes, Peter S. (2005). "Enantioselective organocatalytic Michael addition of malonate esters to nitro olefins using bifunctional cinchonine derivatives". Chemical Communications (35): 4481. doi:10.1039/b508833j. ISSN   1359-7345.
  30. Li, Bang-Jing; Jiang, Lin; Liu, Min; Chen, Ying-Chun; Ding, Li-Sheng; Wu, Yong (2005). "Asymmetric Michael Addition of Arylthiols to α,β-Unsaturated Carbonyl Compounds Catalyzed by Bifunctional Organocatalysts". Synlett (4): 603–606. doi:10.1055/s-2005-863710. ISSN   0936-5214.
  31. McCooey, Séamus H.; Connon, Stephen J. (2005-10-07). "Urea- and Thiourea-Substituted Cinchona Alkaloid Derivatives as Highly Efficient Bifunctional Organocatalysts for the Asymmetric Addition of Malonate to Nitroalkenes: Inversion of Configuration at C9 Dramatically Improves Catalyst Performance". Angewandte Chemie International Edition. 44 (39): 6367–6370. doi:10.1002/anie.200501721. ISSN   1433-7851.
  32. Vakulya, Benedek; Varga, Szilárd; Csámpai, Antal; Soós, Tibor (May 2005). "Highly Enantioselective Conjugate Addition of Nitromethane to Chalcones Using Bifunctional Cinchona Organocatalysts". Organic Letters. 7 (10): 1967–1969. doi:10.1021/ol050431s. ISSN   1523-7060.
  33. Marcelli, Tommaso; van der Haas, Richard N. S.; van Maarseveen, Jan H.; Hiemstra, Henk (2006-01-30). "Asymmetric Organocatalytic Henry Reaction". Angewandte Chemie International Edition. 45 (6): 929–931. doi:10.1002/anie.200503724. ISSN   1433-7851.