Hydroamination

Last updated

In organic chemistry, hydroamination is the addition of an N−H bond of an amine across a carbon-carbon multiple bond of an alkene, alkyne, diene, or allene. [1] In the ideal case, hydroamination is atom economical and green. [2] Amines are common in fine-chemical, pharmaceutical, and agricultural industries. [3] [4] [5] [6] Hydroamination can be used intramolecularly to create heterocycles or intermolecularly with a separate amine and unsaturated compound. The development of catalysts for hydroamination remains an active area, especially for alkenes. Although practical hydroamination reactions can be effected for dienes and electrophilic alkenes, the term hydroamination often implies reactions metal-catalyzed processes.

Contents

History

Hydroamination is well-established technology for generating fragrances from myrcene. In this conversion, diethylamine adds across the diene substituent, the reaction being catalyzed by lithium diethylamide. [7] Intramolecular hydroaminations were reported by Tobin J. Marks in 1989 using metallocene derived from rare-earth metals such as lanthanum, lutetium, and samarium. Catalytic rates correlated inversely with the ionic radius of the metal, perhaps as a consequence of steric interference from the ligands. [8] In 1992, Marks developed the first chiral hydroamination catalysts by using a chiral auxiliary, which were the first hydroamination catalysts to favor only one specific stereoisomer. Chiral auxiliaries on the metallocene ligands were used to dictate the stereochemistry of the product. [9] The first non-metallocene chiral catalysts were reported in 2003, and used bisarylamido and aminophenolate ligands to give higher enantioselectivity. [10]

Notable hydroamination catalysts by year of publication Notable hydroamination catalysts by year of publication.tif
Notable hydroamination catalysts by year of publication

Reaction scope

Hydroamination has been examined with a variety of amines, unsaturated substrates, and vastly different catalysts. Amines that have been investigated span a wide scope including primary, secondary, cyclic, acyclic, and anilines with diverse steric and electronic substituents. The unsaturated substrates that have been investigated include alkenes, dienes, alkynes, and allenes. For intramolecular hydroamination, various aminoalkenes have been examined. [11]

Products

Addition across the unsaturated carbon-carbon bond can be Markovnikov or anti-Markovnikov depending on the catalyst. [12] When considering the possibly of R/S chirality, four products can be obtained: Markovnikov with R or S and anti-Markovnikov addition with R or S. Although there have been many reports of catalytic hydroamination with a wide range of metals, there are far fewer describing enantioselective catalysis to selectively make one of the four possible products. Recently, there have been reports of selectively making the thermodynamic or kinetic product, which can be related to the racemic Markovnikov or anti-Markovnikov structures (see Thermodynamic and Kinetic Product below).

Possible regioselective and stereroselective products Possible regioselective and stereroselective products.tif
Possible regioselective and stereroselective products

Catalysts and catalytic cycle

Hydroamination reactions are atom-efficient processes that generally use readily available and cheap starting materials, therefore a general catalytic strategy is highly desirable. Also, direct catalytic hydroamination strategies have in principle significant benefits over more classical methods to prepare amine containing compounds, including the reduction in the number of synthetic steps required.

[13]

However, hydroamination reactions pose some tough challenges for catalysis: Strong electron repulsion of the nitrogen atom lone pair and the electron rich carbon-carbon multiple bond, coupled with hydroamination reactions being entropically disfavoured (particularly the intermolecular version), [14] [15] results in a large reaction barrier. Regioselectivity issues also hamper the synthetic utility of the resulting products, with Markovnikov addition of the amine being the most common outcome over the less favoured anti-Markovnikov addition (see figure). As a result, there are now numerous catalysts that can be utilised in the hydroamination of alkene, allene and alkyne substrates, including various metal based heterogeneous catalysts, early-transition metal complexes (e.g. titanium and zirconium), late-transition metal complexes (e.g. ruthenium and palladium), lanthanide and actinide complexes (e.g. samarium and lanthanum), as well as Brønsted acids and bases. [16] [17] [18]

General scheme of a catalysed alkyne hydroamination.svg

Catalysts

Many metal-ligand combinations have been reported to catalyze hydroamination, including main group elements including alkali metals such as lithium, [11] group 2 metals such as calcium, [19] as well as group 3 metals such as aluminum, [20] indium, [21] and bismuth. [22] In addition to these main group examples, extensive research has been conducted on the transition metals with reports of early, mid, and late metals, as well as first, second, and third row elements. Finally the lanthanides have been thoroughly investigated. Zeolites have also shown utility in hydroamination. [11]

Catalytic cycles

The mechanism of metal-catalyzed hydroamination has been well studied. [11] Particularly well studied is the organolanthanide catalyzed intramolecular hydroamination of alkenes. [23] First, the catalyst is activated by amide exchange, generating the active catalyst (i). Next, the alkene inserts into the Ln-N bond (ii). [24] Finally, protonolysis occurs generating the cyclized product while also regenerating the active catalyst (iii). Although this mechanism depicts the use of a lanthanide catalyst, it is the basis for rare-earth, actinide, and alkali metal based catalysts.

Proposed catalytic cycle for intramolecular hydroamination Proposed catalytic cycle for intramolecular hydroamination.tif
Proposed catalytic cycle for intramolecular hydroamination

Late transition metal hydroamination catalysts have multiple models based on the regioselective determining step. The four main categories are (1) nucleophilic attack on an alkene alkyne, or allyl ligand and (2) insertion of the alkene into the metal-amide bond. [11] Generic catalytic cycles appear below. Mechanisms are supported by rate studies, isotopic labeling, and trapping of the proposed intermediates.

Common catalytic cycles for hydroamination Three of the four most common catalytic cycles for hydroamination.tif
Common catalytic cycles for hydroamination

Thermodynamics and kinetics

The hydroamination reaction is approximately thermochemically neutral. The reaction however suffers from a high activation barrier, perhaps owing to the repulsion of the electron-rich substrate and the amine nucleophile. The intermolecular reaction also is accompanied by highly negative changing entropy, making it unfavorable at higher temperatures.

[14] [15] Consequently, catalysts are necessary for this reaction to proceed. [3] [11] As usual in chemistry, intramolecular processes occur at faster rates than intermolecular versions.

Thermodynamic vs kinetic product

In general, most hydroamination catalysts require elevated temperatures to function efficiently, and as such, only the thermodynamic product is observed. The isolation and characterization of the rarer and more synthetically valuable kinetic allyl amine product was reported when allenes was used at the unsaturated substrate. One system utilized temperatures of 80 °C with a rhodium catalyst and aniline derivatives as the amine. [25] The other reported system utilized a palladium catalyst at room temperature with a wide range of primary and secondary cyclic and acyclic amines. [26] Both systems produced the desired allyl amines in high yield, which contain an alkene that can be further functionalized through traditional organic reactions.

Possible thermodynamic and kinetic products when utilizing an allene Possible thermodynamic and kinetic products when utilizing an allene2.tif
Possible thermodynamic and kinetic products when utilizing an allene

Base catalyzed hydroamination

Strong bases catalyze hydroamination, an example being the ethylation of piperidine using ethene: [27]

Hydroamination of ethene with piperidine proceeds with no transition metal catalyst, but requires a strong base. C2H4+piperidine.png
Hydroamination of ethene with piperidine proceeds with no transition metal catalyst, but requires a strong base.

Such base catalyzed reactions proceed well with ethene but higher alkenes are less reactive.

Hydroamination catalyzed by group (IV) complexes

Certain titanium and zirconium complexes catalyze intermolecular hydroamination of alkynes and allenes. [3] Both stoichiometric and catalytic variants were initially examined with zirconocene bis(amido) complexes. Titanocene amido and sulfonamido complexes catalyze the intra-molecular hydroamination of aminoalkenes via a [2+2] cycloaddition that forms the corresponding azametallacyclobutane, as illustrated in the figure below. Subsequent protonolysis by incoming substrate gives the α-vinyl-pyrrolidine (1) or tetrahydropyridine (2) product. Experimental and theoretical evidence support the proposed imido intermediate and mechanism with neutral group IV catalysts.

The catalytic hydroamination of aminoallenes to form chiral a-vinyl-pyrrolidine (1) and tetrahydropyridine (2) products. Hydroamination of aminoallenes.png
The catalytic hydroamination of aminoallenes to form chiral α-vinyl-pyrrolidine (1) and tetrahydropyridine (2) products.

Formal hydroamination

The addition of hydrogen and an amino group (NR2) using reagents other than the amine HNR2 is known as a "formal hydroamination" reaction. Although the advantages of atom economy and/or ready available of the nitrogen source are diminished as a result, the greater thermodynamic driving force, as well as ability to tune the aminating reagent are potentially useful. In place of the amine, hydroxylamine esters [28] and nitroarenes [29] have been reported as nitrogen sources.

Applications

Hydroamination could find applications due to the valuable nature of the resulting amine, as well as the greenness of the process. Functionalized allylamines, which can be produced through hydroamination, have extensive pharmaceutical application, although presently such species are not prepared by hydroamination. Hydroamination has been utilized to synthesize the allylamine Cinnarizine in quantitative yield. Cinnarizine treats both vertigo and motion sickness related nausea. [26]

Synthesis of cinnarizine via hydroamination. Cinnarizine synthesis via hydroamination.png
Synthesis of cinnarizine via hydroamination.

Hydroamination is also promising for the synthesis of alkaloids. An example was the hydroamination step used in the total synthesis of (-)-epimyrtine. [30]

Gold-catalyzed hydroamination used for the total synthesis of (-)-epimyrtine. Hydroamination in total synthesis of epimyrtine.png
Gold-catalyzed hydroamination used for the total synthesis of (-)-epimyrtine.

See also

Related Research Articles

<span class="mw-page-title-main">Hydrogenation</span> Chemical reaction between molecular hydrogen and another compound or element

Hydrogenation is a chemical reaction between molecular hydrogen (H2) and another compound or element, usually in the presence of a catalyst such as nickel, palladium or platinum. The process is commonly employed to reduce or saturate organic compounds. Hydrogenation typically constitutes the addition of pairs of hydrogen atoms to a molecule, often an alkene. Catalysts are required for the reaction to be usable; non-catalytic hydrogenation takes place only at very high temperatures. Hydrogenation reduces double and triple bonds in hydrocarbons.

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

An alkyne trimerisation is a [2+2+2] cycloaddition reaction in which three alkyne units react to form a benzene ring. The reaction requires a metal catalyst. The process is of historic interest as well as being applicable to organic synthesis. Being a cycloaddition reaction, it has high atom economy. Many variations have been developed, including cyclisation of mixtures of alkynes and alkenes as well as alkynes and nitriles.

<span class="mw-page-title-main">Wacker process</span>

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Pauson–Khand reaction</span> Chemical reaction

The Pauson–Khand (PK) reaction is a chemical reaction, described as a [2+2+1] cycloaddition. In it, an alkyne, an alkene and carbon monoxide combine into a α,β-cyclopentenone in the presence of a metal-carbonyl catalyst.

In organic chemistry, carbon–hydrogen bond functionalization is a type of organic reaction in which a carbon–hydrogen bond is cleaved and replaced with a C−X bond. The term usually implies that a transition metal is involved in the C−H cleavage process. Reactions classified by the term typically involve the hydrocarbon first to react with a metal catalyst to create an organometallic complex in which the hydrocarbon is coordinated to the inner-sphere of a metal, either via an intermediate "alkane or arene complex" or as a transition state leading to a "M−C" intermediate. The intermediate of this first step can then undergo subsequent reactions to produce the functionalized product. Important to this definition is the requirement that during the C−H cleavage event, the hydrocarbyl species remains associated in the inner-sphere and under the influence of "M".

Hydrosilylation, also called catalytic hydrosilation, describes the addition of Si-H bonds across unsaturated bonds. Ordinarily the reaction is conducted catalytically and usually the substrates are unsaturated organic compounds. Alkenes and alkynes give alkyl and vinyl silanes; aldehydes and ketones give silyl ethers. Hydrosilylation has been called the "most important application of platinum in homogeneous catalysis."

A carbometallation 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. Carbometallations 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 carbometallation.

A frustrated Lewis pair (FLP) is a compound or mixture containing a Lewis acid and a Lewis base that, because of steric hindrance, cannot combine to form a classical adduct. Many kinds of FLPs have been devised, and many simple substrates exhibit activation.

Hydroacylation is a type of organic reaction in which an alkene is inserted into the a formyl C-H bond. The product is a ketone. The reaction requires a metal catalyst. It is almost invariably practiced as an intramolecular reaction using homogeneous catalysts, often based on rhodium phosphines.

Organogold chemistry is the study of compounds containing gold–carbon bonds. They are studied in academic research, but have not received widespread use otherwise. The dominant oxidation states for organogold compounds are I with coordination number 2 and a linear molecular geometry and III with CN = 4 and a square planar molecular geometry.

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile.

<span class="mw-page-title-main">Photoredox catalysis</span>

Photoredox catalysis is a branch of photochemistry that uses single-electron transfer. Photoredox catalysts are generally drawn from three classes of materials: transition-metal complexes, organic dyes, and semiconductors. While organic photoredox catalysts were dominant throughout the 1990s and early 2000s, soluble transition-metal complexes are more commonly used today.

Hydrophosphination is the insertion of a carbon-carbon multiple bond into a phosphorus-hydrogen bond forming a new phosphorus-carbon bond. Like other hydrofunctionalizations, the rate and regiochemistry of the insertion reaction is influenced by the catalyst. Catalysts take many forms, but most prevalent are bases and free-radical initiators. Most hydrophosphinations involve reactions of phosphine (PH3).

<span class="mw-page-title-main">Activation of cyclopropanes by transition metals</span>

In organometallic chemistry, the activation of cyclopropanes by transition metals is a research theme with implications for organic synthesis and homogeneous catalysis. Being highly strained, cyclopropanes are prone to oxidative addition to transition metal complexes. The resulting metallacycles are susceptible to a variety of reactions. These reactions are rare examples of C-C bond activation. The rarity of C-C activation processes has been attributed to Steric effects that protect C-C bonds. Furthermore, the directionality of C-C bonds as compared to C-H bonds makes orbital interaction with transition metals less favorable. Thermodynamically, C-C bond activation is more favored than C-H bond activation as the strength of a typical C-C bond is around 90 kcal per mole while the strength of a typical unactivated C-H bond is around 104 kcal per mole.

In organic chemistry, hydrovinylation is the formal insertion of an alkene into the C-H bond of ethylene. The more general reaction, hydroalkenylation, is the formal insertion of an alkene into the C-H bond of any terminal alkene. The reaction is catalyzed by metal complexes. A representative reaction is the conversion of styrene and ethylene to 3-phenybutene:

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.

In organic chemistry, the Conia-ene reaction is an intramolecular cyclization reaction between an enolizable carbonyl such as an ester or ketone and an alkyne or alkene, giving a cyclic product with a new carbon-carbon bond. As initially reported by J. M. Conia and P. Le Perchec, the Conia-ene reaction is a heteroatom analog of the ene reaction that uses an enol as the ene component. Like other pericyclic reactions, the original Conia-ene reaction required high temperatures to proceed, limiting its wider application. However, subsequent improvements, particularly in metal catalysis, have led to significant expansion of reaction scope. Consequently, various forms of the Conia-ene reaction have been employed in the synthesis of complex molecules and natural products.

Thomas E. Müller is a German chemist and an academic. He is Professor of Carbon source and Conversion at Ruhr-Universität Bochum.

<span class="mw-page-title-main">Hydrocupration</span> Chemical reaction

A hydrocupration is a chemical reaction whereby a ligated copper hydride species, reacts with a carbon-carbon or carbon-oxygen pi-system; this insertion is typically thought to occur via a four-membered ring transition state, producing a new copper-carbon or copper-oxygen sigma-bond and a stable (generally) carbon-hydrogen sigma-bond. In the latter instance (copper-oxygen), protonation (protodemetalation) is typical – the former (copper-carbon) has broad utility. The generated copper-carbon bond (organocuprate) has been employed in various nucleophilic additions to polar conjugated and non-conjugated systems and has also been used to forge new carbon-heteroatom bonds.

References

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

  1. Togni, Antionio; Grützmacher, Hansjörg (2001). Catalytic heterofunctionalization: from hydroanimation to hydrozirconation (1. ed.). Weinheim: Wiley-VCH. doi:10.1002/3527600159. ISBN   978-3527302345.
  2. Beller, M.; Bolm, C. (2004). Transition metals for organic synthesis : building blocks and fine chemicals (2nd ed.). Weinheim: Wiley-VCH. doi:10.1002/9783527619405. ISBN   9783527306138.
  3. 1 2 3 Reznichenko, A. L.; Hultszch, K. C. (2015). Hydroamination of Alkenes. Org. React. Vol. 88. p. 1. doi:10.1002/0471264180.or088.01. ISBN   978-0471264187.
  4. Hultzsch, Kai C. (2005). "Catalytic asymmetric hydroamination of non-activated olefins". Org. Biomol. Chem. (Review). 3 (10): 1819–1824. doi:10.1039/b418521h. PMID   15889160.
  5. Hartwig, J. F. (2004). "Development of catalysts for the hydroamination of olefins" (PDF). Pure Appl. Chem. 76 (3): 507–516. doi:10.1351/pac200476030507. S2CID   29945266.
  6. Pohlki, F.; Doye, S. (2003). "The catalytic hydroamination of alkynes". Chem. Soc. Rev. 32 (2): 104–114. doi:10.1039/b200386b. PMID   12683107.
  7. Takabe, K.; Katahiri, T.; Tanaka, J.; Fujita, T.; Watanabe, S.; Suga, K. (1989). "Addition Of Dialkylamines To Myrcene: N,N-diethylgeranylamine". Org. Synth. 67: 44. doi:10.15227/orgsyn.067.0044.
  8. Gagné, M.R.; Marks, T.J. (1989). "Organolanthanide-catalyzed hydroamination. Facile, regiospecific cyclization of unprotected amino olefins". J. Am. Chem. Soc. 111 (11): 4108. doi:10.1021/ja00193a056.
  9. Gagné, M.R.; Brard, L.; Conticello, V.P.; Giardello, M.A.; Marks, T.J.; Stern, C.L. (1992). "Stereoselection effects in the catalytic hydroamination/cyclization of amino olefins at chiral organolanthanide centers". Organometallics . 11 (6): 2003. doi:10.1021/om00042a012.
  10. O'Shaughnessy, P.N.; Scott, P. (2003). "Biaryl amine ligands for lanthanide catalysed enantioselective hydroamination/cyclisation of aminoalkenes". Tetrahedron Asymmetry . 14 (14): 1979. doi:10.1016/s0957-4166(03)00429-4.
  11. 1 2 3 4 5 6 Müller, Thomas E.; Beller, Matthias (1998). "Metal-Initiated Amination of Alkenes and Alkynes". Chem. Rev. 98 (2): 675–704. doi:10.1021/cr960433d. PMID   11848912.
  12. Beller, M.; Seayad, J.; Tillack, A.; Jiao, H. (2004). "Catalytic Markovnikov and anti-Markovnikov Functionalization of Alkenes and Alkynes: Recent Developments and Trends". Angew. Chem. Int. Ed. 43 (26): 3368–3398. doi:10.1002/anie.200300616. PMID   15221826.
  13. Salvatore, R.N.; Yoon, C.H.; Jung, K.W. (2001). "Synthesis of secondary amines". Tetrahedron . 57 (37): 7785–7811. doi:10.1016/S0040-4020(01)00722-0.
  14. 1 2 Brunet, J.-J.; Neibecker, D.; Niedercorn, F. (1989). "Functionalisation of alkenes: catalytic amination of monoolefins". J. Mol. Catal. 49 (3): 235–259. doi:10.1016/0304-5102(89)85015-1.
  15. 1 2 Johns, A.M.; Sakai, N.; Ridder, A.; Hartwig, J.F. (2006). "Direct Measurement of the Thermodynamics of Vinylarene Hydroamination". J. Am. Chem. Soc. 128 (29): 9306–9307. doi:10.1021/ja062773e. PMID   16848446.
  16. Müller, T.E.; Hultzsch, K.C.; Yus, M.; Foubelo, F.; Tada, M. (2008). "Hydroamination: Direct Addition of Amines to Alkenes and Alkynes". Chem. Rev. 108 (9): 3795–3892. doi:10.1021/cr0306788. ISSN   0009-2665. PMID   18729420.
  17. Alonso, F.; Beletskaya, I.P.; Yus, M. (2004). "Transition-Metal-Catalyzed Addition of Heteroatom−Hydrogen Bonds to Alkynes". Chem. Rev. 104 (6): 3079–3160. doi:10.1021/cr0201068. ISSN   0009-2665. PMID   15186189.
  18. Aillaud, I.; Collin, J.; Hannedouche, J.; Schulz, E. (2007). "Asymmetric hydroamination of non-activated carbon–carbon multiple bonds". Dalton Trans. (44): 5105–5118. doi:10.1039/b711126f. ISSN   1477-9226. PMID   17985016.
  19. Crimmin, M.R.; Casely, I.J.; Hill, M.S. (2005). "Calcium-Mediated Intramolecular Hydroamination Catalysis". J. Am. Chem. Soc. 127 (7): 2042–2043. doi:10.1021/ja043576n. PMID   15713071.
  20. Koller, J.; Bergman, R.G. (2010). "Highly Efficient Aluminum-Catalyzed Hydro-amination/-hydrazination of Carbodiimides". Organometallics . 29 (22): 5946–5952. doi:10.1021/om100735q.
  21. Sarma, R.; Prajapati, D. (2011). "Indium catalyzed tandem hydroamination/hydroalkylation of terminal alkynes". Chem. Commun. 47 (33): 9525–7. doi:10.1039/c1cc13486h. PMID   21776504.
  22. Komeyama, K.; Kouya, Y.; Ohama, Y.; Takaki, K. (2011). "Tandem ene-reaction/hydroamination of amino-olefin and -allene compounds catalyzed by Bi(OTf)3". Chem. Commun. 47 (17): 5031–5033. doi:10.1039/c0cc05258b. PMID   21423974.
  23. Hong, S.; Marks, T.J. (2004). "Organolanthanide-Catalyzed Hydroamination". Acc. Chem. Res. 37 (9): 673–686. doi:10.1021/ar040051r. PMID   15379583.
  24. Crabtree, Robert H. (2005). The organometallic chemistry of the transition metals (4th ed.). Hoboken, N.J.: John Wiley. ISBN   978-0-471-66256-3.
  25. Cooke, M.L.; Xu, K.; Breit, B. (2012). "Enantioselective Rhodium-Catalyzed Synthesis of Branched Allylic Amines by Intermolecular Hydroamination of Terminal Allenes". Angew. Chem. Int. Ed. 51 (43): 10876–10879. doi:10.1002/anie.201206594. PMID   23011801.
  26. 1 2 Beck, J.F.; Samblanet, D.C.; Schmidt, J.A.R. (2013). "Palladium catalyzed intermolecular hydroamination of 1-substituted allenes: an atom-economical method for the synthesis of N-allylamines". RSC Adv. 3 (43): 20708–20718. Bibcode:2013RSCAd...320708B. doi:10.1039/c3ra43870h.
  27. Wollensak, J.; Closson, R.D. (1963). "N-Ethylpiperidine". Org. Synth. 43: 45. doi:10.15227/orgsyn.043.0045.
  28. Miki, Y.; Hirano, K.; Satoh, T.; Miura, M. (2013). "Copper-Catalyzed Intermolecular Regioselective Hydroamination of Styrenes with Polymethylhydrosiloxane and Hydroxylamines". Angew. Chem. Int. Ed. 52 (41): 10830–10834. doi:10.1002/anie.201304365. ISSN   1521-3773. PMID   24038866.
  29. Gui, J.; Pan, C.-M.; Jin, Y.; Qin, T.; Lo, J.C.; Lee, B.J.; Spergel, S.H.; Mertzman, M.E.; Pitts, W.J. (2015). "Practical olefin hydroamination with nitroarenes". Science . 348 (6237): 886–891. Bibcode:2015Sci...348..886G. doi: 10.1126/science.aab0245 . ISSN   0036-8075. PMID   25999503.
  30. 1 2 Trinh, T.T.H.; Nguyen, K.H.; Amaral, P. de A.; Gouault, N. (2013). "Total synthesis of (−)-epimyrtine by a gold-catalyzed hydroamination approach". Beilstein J. Org. Chem. 9: 2042–2047. doi:10.3762/bjoc.9.242. PMC   3817515 . PMID   24204417.