CC BY 4.0 · Synlett 2023; 34(18): 2117-2128
DOI: 10.1055/a-2111-9629
account
Modern Boron Chemistry: 60 Years of the Matteson Reaction

Recent Advances in Catalysis Using Organoborane-Mediated Hydride Abstraction

Joseph P. Gillions
a   School of Chemistry, University of Leicester, Leicester, LE1 7RH, UK
,
Salma A. Elsherbeni
b   Cardiff Catalysis Institute, School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff, CF10 3AT, UK
,
Laura Winfrey
a   School of Chemistry, University of Leicester, Leicester, LE1 7RH, UK
,
Lei Yun
a   School of Chemistry, University of Leicester, Leicester, LE1 7RH, UK
,
Rebecca L. Melen
c   Cardiff Catalysis Institute, School of Chemistry, Cardiff University, Translational Research Hub, Maindy Road, Cathays, Cardiff, CF24 4HQ, UK
,
Louis C. Morrill
b   Cardiff Catalysis Institute, School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff, CF10 3AT, UK
,
Alexander P. Pulis
a   School of Chemistry, University of Leicester, Leicester, LE1 7RH, UK
› Author Affiliations
A.P.P. thanks the Engineering and Physical Sciences Research Council (EPSRC) for funding (EP/W02151X/1 and EP/Y00146X/1). R.L.M. would like to thank the EPSRC for a research fellowship (EP/R026912/1). J.P.G., L.W., L.Y. and A.P.P. thank the University of Leicester for generous support. L.Y. thanks the China Scholarship Council for funding. S.A.E., R.L.M. and L.C.M. gratefully acknowledge the School of Chemistry, Cardiff University for generous support. S.A.E. thanks the British Council and the Egyptian Cultural Affairs and Missions Sector for a Ph.D. studentship through the Newton-Mosharafa Fund.
 


Dedicated to 60 years of Donald Matteson’s boron homologation

Abstract

C–H functionalization is widely regarded as an important area in the development of synthetic methodology, enabling the design of more time- and atom-efficient syntheses. The ability of electron-deficient organoboranes to mediate hydride abstraction from α-amino C–H bonds is therefore of great interest, as the reactive iminium and hydridoborate moieties generated are able to participate in a range of synthetically useful transformations. In this review, we cover the recent advances made in organoborane-mediated hydride abstraction, and focus on the catalytic applications of electron-deficient boranes in α- or β-functionalization, α,β-difunctionalization, and the dehydrogenation of amines.

1 Introduction

2 α-Functionalization of Amines

3 β-Functionalization of Amines

4 α,β-Difunctionalization of Amines

5 Dehydrogenation of Amines

6 Summary and Future Prospects


#

Biographical Sketches

Zoom Image

from left to right


Joseph P. Gillions completed his MChem degree with a Year in Industry in 2020 at the University of York and GlaxoSmithKline. Joe is currently studying for his PhD at the University of Leicester with Dr Alex Pulis where he is working on new catalytic methodology for the functionalisation of amines.


Salma A. Elsherbeni received a Masters in Pharmaceutical Sciences at Tanta University (Egypt) and is now pursuing a PhD in Chemistry at Cardiff University.


Laura Winfrey completed her MChem degree with a year abroad in 2019 at the University of Leicester (UK) and Kent State University (USA). Laura is currently studying for her PhD at the University of Leicester with Dr Alex Pulis focused on new catalytic methods for the synthesis of amines.


Lei Yun complete his Masters degree at the Dalian­ University of Technology (China) in 2021. Lei is now studying for his PhD in the Pulis Group at the University of Leicester funded by the China Scholarship Council. He is working on boron-catalyzed functionalisation of bioactive amines.


Prof. Rebecca Melen studied for her undergraduate and Ph.D. degrees at the University of Cambridge, completing her Ph.D. in 2012 with Prof. Wright. Following postdoctoral studies with Prof. Stephan in Toronto and with Prof. Gade in Heidelberg, she took up a position at Cardiff University in 2014, where she is now a Professor in inorganic chemistry. In 2018, she was awarded an EPSRC early career fellowship, and she is the recipient of the 2019 RSC Harrison Meldola Memorial Prize and a 2022 Philip Leverhulme Prize in Chemistry. Her research interests lie in main group chemistry and the applications of main group Lewis acids in synthesis and catalysis.


Louis Morrill received his PhD from the University of St Andrews in 2014 under the direction of Prof. Andrew Smith and undertook postdoctoral research at UC Berkeley with Prof. Richmond Sarpong. In June 2015, he initiated his independent research career at Cardiff University. Research in the group is focused on inventing new reactions in organic chemistry and developing sustainable catalytic methodologies for synthesis.


Dr Alex Pulis obtained his PhD from the University of Bristol (UK) under the guidance of Prof. Varinder K. Aggarwal. In 2014, he joined Prof. Douglas Stephan at the University of Toronto (Canada) for postdoctoral studies. He then moved to the University of Manchester (UK) as a fixed term Lecturer within the group of Prof. David J. Procter. Alex began his independent career at the University of Leicester (UK) in 2018 where he explores the reactivity of main group elements and applies these finding to challenges in chemical synthesis.

1

Introduction

Organoboranes are widely used as reagents and building blocks in synthetic chemistry. For example, hydroboranes and borohydrides are often used in reduction chemistry, whereas boronic acids are commonly employed in cross-coupling processes.[1] [2] The use of organoboranes as catalysts has also received significant attention whereby the interaction of the Lewis acidic boron atom with a pair of non-bonding electrons or π-electron pairs is employed to activate the substrate.[3–5] Recent attention has focused on more electron-deficient organoboranes, such as B(C6F5)3, in catalysis. These more Lewis acidic species can also interact with σ(C–H) bonds, resulting in hydride abstraction and the formation of a formal carbocation and a borohydride counterion.[6] [7] [8] When the C–H-bearing substrate is a cyclohexadiene, Wheland-type intermediates are formed,[9] [10] [11] [12] and when dihydropyridines are used, pyridinium salts form (Scheme [1a]).[13] [14] The most explored reactions of this type are those that involve organoborane-mediated hydride abstraction from α-amino C–H bonds, during which iminium borohydride salts 4 are generated (Scheme [1b]).

Zoom Image
Scheme 1 Borane-mediated hydride abstraction forms Wheland intermediates, pyridinium borohydrides, or iminium borohydrides

Organoborane-mediated hydride abstraction has been exploited in a variety of reactions that generate iminium salts in situ directly from the corresponding alkyl amines and include organoborane-catalyzed amine-based transfer hydrogenation and dehydrogenation, racemization and isomerization, α-functionalization, β-functionalization, dual α,β-functionalization, and C–N bond cleavage. In this review, we include new developments in organoborane-catalyzed processes involving organoborane-mediated hydride abstraction in amines that have been disclosed since our previous review.[6] , [15] [16] [17] New reactions herein include examples that generate complex amine products via cooperative organoborane-metal catalysis, incorporate hydride shuttles, lead to multifunctionalizations, and allow dehydrogenation of liquid organic hydrogen carriers. We will highlight key features of the reactions and discuss the mechanisms in the context of the fate of the iminium ion and how the borohydride reacts to allow catalyst turnover.


# 2

α-Functionalization of Amines

In our previous review[6] we reported studies whereby organoboranes catalyze the α-functionalization of amines. B(C6F5)3-mediated α-amino C(sp3)–hydride abstraction was shown to result in the formation of hydridoborate and iminium ions, which can be intercepted by various nucleophiles to result in formal α-N C–H functionalization processes with amines. In this review, we cover the reports since our prior review, which include α-alkynylation, α-furylation, and cyclization reactions.

Zoom Image
Scheme 2 Borane-mediated hydride abstraction for the direct conversion of N-alkylamines into N-propargylamines; DCE = 1,2-dichloro­ethane

In 2020, Wasa and co-workers reported the conversion of N-alkylamines 5 and alkynyl trimethylsilanes 6 into propargyl amines 7 via dual Lewis acid/organocopper catalysis (Scheme [2]).[18] The catalyst system employed was composed of B(C6F5)3 and Cu(MeCN)4PF6 in combination with various ligands such as (S)-Ph-PyBOX (9a), (S)-(3,5-Me2-C6H3)-PyBOX (9b), or 1,2-bis(diphenylphosphino)ethane (9c). The proposed reaction mechanism is initiated by conversion of 10 into borate ion pair 11 in the presence of trityl alcohol or water. Ligand exchange in 11 with alkynyl silane 6 generates trimethylsilanol (12) and alkynyl borate 13. Subsequent transmetalation generates an alkynyl copper intermediate 16 in addition to B(C6F5)3, which abstracts a hydride from the α-N position within amine 5 to generate the iminium hydridoborate ion pair 17. Addition of the alkynyl copper intermediate 16 to the iminium ion produces N-propargylamine 7, while the hydridoborate ion reacts with trityl alcohol 19 to regenerate 11 and complete the catalytic cycle. Impressively, this approach enabled the derivatization of a wide range of N-protected bioactive molecules, such as the antidepressants fluoxetine (cf. 7a) and duloxetine (cf. 7c). The protocol tolerated a range of functional groups, such as protected alcohols, amides, esters, and halogens. For amine substrates with multiple sites of potential hydride abstraction, the regioselectivity of propargylation was attributed to the rapid consumption of short-lived CH2 iminium ions (derived from B(C6F5)3-mediated hydride abstraction at N-Me sites) before isomerization to lower energy iminium ions can occur. Furthermore, through employing chiral PyBOX ligand 9b, a variety of N-propargylamines (e.g., 7b) could be accessed with high levels of enantiocontrol (up to 94% ee).

Zoom Image
Scheme 3 Borane-mediated hydride abstraction utilized in the α-furanylation of amines

In 2022, Wang and co-workers reported the α-furylation of N-methyl-substituted tertiary amines using a borane/gold(I) co-catalytic system (Scheme [3]).[19] It was found that a range of tertiary N-methylamines 20 and α-alkynylenones 21 could be converted into substituted furans 22 in the presence of B(C6F5)3 and AuCl(PPh3) catalysts. α-Furylation occurred regioselectively at N-methyl groups in the presence of N-benzyl groups. The mechanism was proposed to proceed via B(C6F5)3-mediated α-amino hydride abstraction to form an iminium hydridoborate ion 23 with concurrent gold-promoted cycloisomerization of α-alkynylenone 21 to produce an [Au]-associated furyl 1,3-dipole 24. Borohydride reduction of the furyl cation forms furyl species 25, which then adds to the iminium ion 26 to produce the observed α-amino furylation product 22, whilst regenerating the borane and gold catalysts. The procedure could be applied to a range of compounds, including bioactive compounds such as butenafine to yield derivative 22a. The reaction showed good functional group tolerance, including protected alcohols 22b, ethers 22c, trifluoromethyl groups, and halogens. When the iminium ions contained β-protons, enamine intermediates were formed and engaged the aurated furans in [3+2] cycloadditions resulting in α- and β-functionalization (see Section 4).

In 2021, Wang and co-workers reported the synthesis of N-heterocycles 28 and 29 via the B(C6F5)3-catalyzed dehydrogenative cyclization of 2-cyclopropyl-N,N-dimethylanilines 27 (Scheme [4]).[20]

Zoom Image
Scheme 4 Borane-mediated hydride abstraction utilized in the cyclization of cyclopropyl anilines

Their protocol combined B(C6F5)3-mediated hydride abstraction with the B(C6F5)3-catalyzed cyclopropane-ring opening previously reported by Wang’s group in 2017.[21] Initially, B(C6F5)3 opens the cyclopropane ring to form zwitterion 30, before deprotonation by Barton’s base (2-tert-butyl-1,1,3,3-tetramethylguanidine) (BTMG) yields alkene 31. B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride of the aniline then yields zwitterionic intermediate 33. Finally, intramolecular attack of the alkene moiety on the iminium yields the cyclization products 28 or 29. Electron-donating groups on the non-aniline aryl group promoted attack through the terminal alkenyl carbon (C2 of 33) to yield 1,2,3,4-tetrahydroquinolines 28. B(C6F5)3 is regenerated via an acid–base reaction within guanidinium borohydride salt 32, generating H2. Electron-withdrawing groups gave mixtures of 1,2,3,4-tetrahydroquinolines 28 and indolines 29 (often the major product), formed by attack of the alkene through the benzylic carbon (C1) of the allylic borate. Functional group compatibility was demonstrated with, for example, protected phenol (28a), halide (28b), and trifluoromethyl ether (29a) substituents.

In 2022, Maulide and co-workers reported the stereoselective synthesis of a variety of azabicyclic structures 36 from enamines 34 (derived from N-heterocycles) and Michael acceptors 35 (Scheme [5]).[22] Initially, the two substrates, enamine 34 and Michael acceptor 35, react to form iminium 37. A mixture of B(2,6-F2C6H3)3 and the corresponding tetrabutylammonium borohydride salt effectively isomerizes iminium ion 38 to form iminium 41 via a sequence of hydride donation to form 39 followed by hydride abstraction. A similar isomerization was also proposed by Oestreich in the β,β′-H silylation of tertiary amines described below (see Scheme [9]). Interestingly, the Lewis acid B(2,6-F2C6H3)3 is significantly less electron-deficient than B(C6F5)3, which is almost exclusively used in organoborane-mediated hydride abstraction with amines. Other organoboranes, including B(C6F5)3, BMes(2,6-F2C6H3)2 and BPh3, failed to yield the desired products, whilst B(2,4,6-F3C6H2)3 gave yields of <10%. With the iminium in the correct position, cyclization occurs to give the desired aza-bicycles 36 as single diastereomers in most cases. The reaction tolerated a range of substituents on the Michael acceptor 35 and enamine 34, including different heterocycles (e.g., 36a), ethers, halides and protected alcohols. Nitro or trifluoromethyl ketones could also act as the electron-withdrawing group in 35. Acyclic enamines were amenable to the method, forming highly substituted monocyclic piperidines. Additionally, it was shown that the process could be telescoped from the aldehyde and amine corresponding to the enamine. Impressively, an enantioselective variant was also reported, whereby an enantioselective organocatalyzed Michael addition yielded enantioenriched aldehyde 44 prior to formation of iminium 45. Lewis acid/borohydride catalysis enabled the formation of azabicyclic products (e.g., 36c) with very high enantioselectivity.

Zoom Image
Scheme 5 Borane-mediated hydride abstraction utilized in the diastereoselective construction of azabicycles from enamines and Michael acceptors. Asymmetric Michael addition conditions: (i) l-phenylalanine lithium salt (10 mol%), CH2Cl2, 48 h; or (ii) O-(tert-butyl)-l-threonine (5 mol%), 4-dimethylaminopyridine (5 mol%), sulfamide (5 mol%), toluene, 6–36 h.

# 3

β-Functionalization of Amines

Our previous review covered organoborane-catalyzed β-functionalization of amines, describing reactions such as β-silylation, β-alkylation, and β-deuteration.[6] These reactions occur when the iminium species formed upon hydride abstraction undergoes deprotonation at the β-position, yielding reactive enamines. These enamines are capable of acting as nucleophiles, which, if an appropriate electrophile is available, can lead to β-functionalization. In this review, we cover new β-functionalizations with isatins and Michael acceptors, as well as β-silylation.

Yang, Ma and co-workers reported the β-functionalization of pyrrolidines 46 with isatins 47 to give substituted pyrrolidines 48 (Scheme [6]).[23] The catalytic cycle proceeds via B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride on pyrrolidine 46. The β-proton of the resultant iminium 49 is then removed to form enamine 50 and ammonium borohydride salt 51. Enamine 50 attacks the isatin 47 to form a new C–C bond in 52. Elimination of water then forms unsaturated species 53. Reduction of the iminium moiety in 53 with the hydride from the borohydride counterion forms β-functionalized product 48 and regenerates the B(C6F5)3 catalyst. The reaction demonstrated good stereoselectivity and tolerated various functional groups such as halides, nitro groups (e.g., 48a), N-branched alkenes and alkynes (e.g., 48b). The use of diethyl ketomalonate rather than isatin 47 furnished product 48c, where water was not eliminated. It was also possible to dehydrogenate the pyrrolidine products 48 to form pyrroles in situ (see Section 5).

Zoom Image
Scheme 6 Borane-mediated hydride abstraction for consecutive β-functionalization and acceptorless dehydrogenation
Zoom Image
Scheme 7 Borane-mediated hydride abstraction for β-functionalization via a Michael-type reaction

In 2021, Wasa and co-workers studied the application of B(C6F5)3 in the β-amino C–H functionalization of amines 54 with Michael acceptors 55 to synthesize N-alkylamines 56 or 57 (Scheme [7]).[24] The reaction proceeds via B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride yielding iminium borohydride 59, before deprotonation by a base (e.g., 54 or 56) to give enamine 61. This then undergoes nucleophilic attack on the B(C6F5)3-activated Michael acceptor 62 forming enolate 63, before reduction by the protonated base and hydridoborate yields the product 56 (path i). Alternatively, enolate 63 can tautomerize to form alternative product 57, with the protonated base/hydridoborate instead reducing the Michael acceptor 55 to yield alkane 58 (path ii). The method showed exceptional functional group tolerance, with examples including ethers, secondary amines, and protected alcohols. Excitingly, the method could also be used to derivatize bioactive compounds, including silyl-protected raloxifene and risperidone to give products 56a and 57a, respectively.

In the same paper,[24] the Wasa group also reported an enantioselective variation of the reaction, using a B(C6F5)3/Sc(OTf)3/PyBOX catalytic system to enable enantioselective C–C bond formation between N-alkylamines and Michael acceptors (Scheme [8]). As with the racemic version, the reaction proceeds through B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride to give iminium borohydride 59. Mechanistic studies suggest that in the next step the borohydride reduces one equivalent of the Michael acceptor, yielding boron enolate 68. Interestingly, kinetic and NMR spectroscopic studies suggest that the rate-limiting step is the proton transfer from the β-position to the N-position of the enamine, yielding intermediate 69. This intermediate then protonates the boron enolate, giving 67 as a by-product, and freeing up the enamine to undergo nucleophilic attack on the Sc-bound Michael acceptor 70, forming enolate 71. Due to the chiral scandium complex, attack selectively occurs on the Michael acceptor. Product 66 can then form by intramolecular proton transfer, whilst product 65 forms upon reduction by a protonated base and the borohydride. As with the racemic version, the reaction showed exceptional functional group tolerance, with examples including trifluoromethyl groups (e.g., 65a), ketones, and protected alcohols. Additionally, bioactive molecules were also derivatized (cf. 66a and 66b).

Zoom Image
Scheme 8 Borane-mediated hydride abstraction for the enantioselective β-functionalization of amines through a Michael-type addition

In 2021, Oestreich and co-workers reported a B(C6F5)3-catalyzed β,β′-H silylation of tertiary amines 72, yielding sila analogues of piperidines 74 (Scheme [9]).[25]

Zoom Image
Scheme 9 Borane-mediated hydride abstraction leads to the β,β′-H silylation of tertiary amines for the synthesis of sila-piperidines

The mechanism proceeds through B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride in 72, to give iminium borohydride 75, before deprotonation by another equivalent of the amine 72 yields enamine 76 and ammonium borohydride 77. The ammonium borohydride then eliminates H2 to reform the active B(C6F5)3 catalyst and amine 72. The B(C6F5)3 catalysis can also activate the Si–H bond (cf. 78), creating an Si electrophile that can be attacked by enamine 76 to form a C–Si bond. The resulting iminium 79 can then undergo borohydride-induced tautomerization to give iminium 80, which can be deprotonated by another equivalent of 72 to yield enamine 81 and ammonium borohydride 77. The enamine 81 then proceeds through further B(C6F5)3-mediated C–Si bond formation, as before, to yield the desired product 74, with the ammonium borohydride 77 evolving H2 to reform the active catalyst B(C6F5)3. The reaction tolerated functional groups on the amine starting material such as unprotected phenols (e.g., 74a), halogens, and trifluoromethyl groups, and the silane could also be varied to give unsymmetric silanes (e.g., 74b) and substituted diarylsilanes (e.g., 74c).

In 2023, He, Zhao and co-workers reported the B(C6F5)3-catalyzed β-alkylation of tertiary amines 82 with 3H-indol-3-ones 83 to yield 2-alkylindolin-3-ones 84 (Scheme [10]).[26]

Zoom Image
Scheme 10 Borane-mediated hydride abstraction facilitates the β-functionalization of amines with 3H-indol-3-ones

The reaction is proposed to proceed through B(C6F5)3-mediated abstraction of the α-N C(sp3)–H hydride to yield iminium borohydride 85. This can then be deprotonated by a base (e.g., amines 82 or 88) to yield enamine 87, which undergoes nucleophilic addition to the 3H-indol-3-one 83 to give zwitterionic intermediate 88. This can then accept a proton from 86, before borohydride reduction to yield the desired product 84, as well as reforming B(C6F5)3. Functional group tolerance was demonstrated with ethers (e.g., 84a) and halogens (e.g., 84b), as was various substitution patterns around the aromatic rings. Additionally, the reaction was demonstrated to work with N-butyl-N,N-diethylamine to give amine 84c.


# 4

α,β-Difunctionalization of Amines

α,β-Difunctionalization of amines facilitated by organoborane-mediated hydride abstraction had not been reported at the time of our previous review,[6] and represents a new reaction class. Here the enamine can participate in cycloaddition reactions resulting in α,β-difunctionalization. In this section we will look at how this is leveraged to form cyclobutenes, furan-fused cyclopentenes, and alkyl-amino-functionalized quinolines.

In 2021, Wang and Zhang reported the enantioselective organoborane-catalyzed coupling of 1,2-dihydroquinolines 89 with alkynes 90 to deliver cyclobutene-fused 1,2,3,4-tetrahydroquinolines 92 (Scheme [11]).[27] The catalytic cycle is proposed to occur via hydride abstraction mediated by organoborane 91 that converts 1,2-dihydroquinoline 89 into quinolinium 93. Transfer of the hydride in 93 to the quinolinium fragment forms 1,4-dihydroquinoline 94. Effectively, organoborane 91 mediates isomerization of the dihydroquinoline via hydride abstraction. Organoborane 91 then plays the role of a conventional Lewis acid, activating alkyne 90 to allow the [2+2] cycloaddition with 94 to occur.

Zoom Image
Scheme 11 Borane-mediated hydride abstraction leads to a [2+2] cycloaddition reaction between the α- and β-positions of amines and alkynes
Zoom Image
Scheme 12 Borane-mediated hydride abstraction in conjunction with gold(I)-catalysis leads to the formal [3+2] cycloaddition of amines and α-alkynylenones

As part of their study into the borane/gold(I)-co-catalyzed α-furylation of tertiary N-methylamines 20 (c.f. Scheme [3]), Wang and co-workers discovered that formal [3+2] cycloaddition products were observed when utilizing N-alkyl-substituted tertiary amines 96 containing β-hydrogens (Scheme [12]).[19] It was found that a broad range of cycloadducts 97 could be accessed from tertiary N-alkylamines 96 and α-alkynylenones 21. In this case, B(C6F5)3-mediated hydride abstraction generates an iminium borohydride 99, which is deprotonated by a base (e.g., amines 96 or 97) to generate enamine intermediate 100. A subsequent [3+2] cycloaddition involving the enamine 100 and the aurated furan 1,3-dipole 24 gives the cycloaddition adduct 97 and regenerates the gold catalyst, whilst B(C6F5)3 can be reformed by transfer hydrogenation of another molecule of 24. The reaction was shown to tolerate substituents at the β-position of the amine (e.g., 97b), as well as ethers (e.g., 97a), halogens, and protected phenols. The reaction conditions could also be applied to the derivatization of bioactive compounds, such as in the formation of paroxetine derivative 97d.

In 2022, He, Fan and co-workers reported an α,β-functionalization strategy involving anilines 102 and benzo[c]isoxazoles 103 followed by a C–N bond cleavage to furnish functionalized quinoline derivatives 104 (Scheme [13]).[28] The catalytic cycle proceeds via B(C6F5)3-mediated abstraction of the α-N C(sp3)–H on N-aryl N,N-dialkyl amine 102. Deprotonation of the resultant iminium 105 (e.g., with 102 or 104) forms enamine 106 and ammonium borohydride salt 108. A [4+2] cycloaddition of enamine 106 with isoxazole 103 forms α,β-functionalized piperidine 107, which is then protonated by the ammonium salt 108 to form intermediate borohydride salt 109. Hydride from the borohydride counterion cleaves the N–O bond in 109 to form alcohol 110 and regenerate the catalyst. Elimination of water gives 111 before a tautomerization and subsequent C–N bond cleavage furnishes functionalized quinoline 104. The reaction tolerated a broad range of functional groups on the aniline 102 ring including cyano (e.g., 104a), trifluoromethyl, nitro, ether and thioether groups. The saturated heterocycle in 102 was varied to allow for pyrrolidines, piperidines, azepanes (e.g., 104b), and even azocanes and azonanes to all be used in the protocol. Acyclic examples such as diethylaniline and dipropylaniline were used to furnish quinoline 112 and 3-methylquinoline respectively. When ethyl 1-phenylpiperidine-4-carboxylate was used, a further intermolecular condensation of the secondary amine and ester group delivered product 113 with a lactam moiety.

Zoom Image
Scheme 13 Borane-mediated hydride abstraction for the conversion of benzo[c]isoxazoles into functionalized quinolines

# 5

Dehydrogenation of Amines

In our previous review,[6] we observed how hydride abstraction had been utilized in the dehydrogenation of a variety of benzofused N-heterocycles,[29] as well as the dehydrogenative coupling of indoles with silanes and boranes. These reactions are believed to proceed via an acceptorless dehydrogenation pathway, where an ammonium borohydride intermediate (generated after hydride abstraction and iminium tautomerization) undergoes an acid–base reaction to evolve H2. We have also seen a few examples of H2 evolution in several of the examples described above. Here, we look at the dehydrogenation of β-functionalized pyrrolidines, and how the dehydrogenation of tetrahydroquinolines has been utilized for hydrogen storage and purification.

Zoom Image
Scheme 14 Borane-mediated hydride abstraction for the acceptorless dehydrogenation of isatin-functionalized pyrrolidines

In the report from Yang, Ma and co-workers described in Section 3 on the β-functionalization of pyrrolidines 46 with isatins 47, a pyrrolidine acceptorless dehydrogenation process was also reported.[23] The products of the aforementioned β-functionalization of pyrrolidines 48 were converted into pyrroles 114 (Scheme [14]). Simply increasing the temperature of the standard β-functionalization conditions led to the formation of pyrroles. The authors proposed that the mechanism proceeds through B(C6F5)3-mediated hydride abstraction on product 48 (as formed above in Scheme [6]) to form iminium species 115. Isomerization of 115 into 117 (or 116) creates an acidic proton that reacts with the borohydride counterion, causing aromatization, loss of H2 and reformation of B(C6F5)3. Functional group tolerance was demonstrated for the two-step procedure with nitro groups (e.g., 114a), halogens, alkenes (e.g., 114b), and ethers (e.g., 114c).

Ogoshi, Hoshimoto and co-workers have reported an organoborane-catalyzed dehydrogenation applied to the purification of molecular hydrogen (Scheme [15]).[30]

Zoom Image
Scheme 15 Borane-mediated hydride abstraction for the interconversion of liquid organic hydrogen carriers.

A selection of organoboranes was shown to mediate the dehydrogenation of tetrahydroquinoline 118 to form quinoline 119 in high yield. Organoborane derivatives based on (2,6-dichlorophenyl)bis(2,6-difluorophenyl)borane (e.g., 120122), were shown to be optimal for the dehydrogenation of 118, with the brominated derivative 120 performing across both dehydrogenation and hydrogenation steps (see below). Whilst no mechanism for the dehydrogenation was proposed, it is likely that the reaction proceeds in a manner analogous to the acceptorless dehydrogenations encountered above (e.g., Scheme [13], conversion of 102 into 106) and in our previous review.[6] The ability of organoboranes to mediate dehydrogenation via hydride abstraction was applied to the purification of hydrogen from gaseous mixtures via a hydrogenation–dehydrogenation sequence. Hydrogen was removed from a mixture of H2, CO and CO2 and incorporated into tetrahydroquinoline 118 via frustrated Lewis pair hydrogenation of quinoline 119.[31] Tetrahydroquinoline 118 was easily removed from the gas mixture, and separately dehydrogenated with borane 120, concurrently forming high purity H2 gas. Interestingly, the ubiquitous B(C6F5)3 borane performed significantly worse in the initial hydrogenation step due to competing reactions with the H2/CO/CO2 mixture, as well as in the dehydrogenation step (88% vs 18% of 119).


# 6

Summary and Future Prospects

In this review update, we have covered the recent advances made in organoborane-mediated hydride abstraction. We have shown that various Lewis acidic borane catalysts are able to mediate hydride abstraction from α-amino C–H bonds to yield a reactive iminium cation and a hydridoborate anion. These have been demonstrated to participate in a range of synthetically useful transformations, including α- or β-functionalization, α,β-difunctionalization, and the dehydrogenation of amines. It has become clear that further investigations are needed around the structure and electronic properties of the organoborane catalyst, as subtle changes play an important role in catalyst performance, as underscored by the work of Maulide,[22] Wang,[27] and Ogoshi and Hoshimoto[30] described above. The repurposing of common amine starting materials for novel transformations is an important area in synthetic chemistry and we envision that further development in the area of organoborane-mediated hydride abstraction will advance synthetic methodology, and lead to more time- and atom-efficient syntheses in the future.


#
#

Conflict of Interest

The authors declare no conflict of interest.

  • References and Notes

  • 1 Seyden-Penne J. In Reductions by the Alumino- and Borohydrides in Organic Synthesis. Curran D. Wiley-VCH; New York: 1997: 5
  • 2 Sun H.-Y, Hall DG. In Synthesis and Application of Organoboron Compounds . Fernández E, Whiting A. Springer International Publishing; Switzerland: 2015: 221
  • 3 Graham BJ, Raines RT. J. Org. Chem. 2022; in press; DOI DOI: 10.1021/acs.joc.2c01695.
  • 4 Ji X, Liu L. Eur. J. Org. Chem. 2022; e202200853
  • 5 Tan X, Wang H. Chem. Soc. Rev. 2022; 51: 2583
  • 6 Basak S, Winfrey L, Kustiana BA, Melen RL, Morrill LC, Pulis AP. Chem. Soc. Rev. 2021; 50: 3720
  • 7 Ma Y, Loub S.-J, Hou Z. Chem. Soc. Rev. 2021; 50: 1945
  • 8 Kumar G, Roy S, Chatterjee I. Org. Biomol. Chem. 2021; 19: 1230
  • 9 Walker JC. L, Oestreich M. Synlett 2019; 30: 2216
  • 10 Keess S, Oestreich M. Chem. Sci. 2017; 8: 4688
  • 11 Oestreich M. Angew. Chem. Int. Ed. 2016; 55: 494
  • 12 Li L, Yang X. Chin. J. Org. Chem. 2023; 43: 1036
  • 13 Webb JD, Laberge VS, Geier SJ, Stephan DW, Crudden CM. Chem. Eur. J. 2010; 16: 4895
  • 14 Wilkins LC, Santi N, Luk LY. P, Melen RL. Phil. Trans. R. Soc. A 2017; 375: 20170009

    • Organoborane-catalyzed processes that do not include direct hydride abstraction from C–H to B–H, such as pericyclic 1,n-hydride shifts, are beyond the scope of this review. For examples, see:
    • 15a Zhou R, Paradies J. Eur. J. Org. Chem. 2021; 6334
    • 15b Zhang B.-B, Peng S, Wang F, Lu C, Nie J, Chen Z, Yang G, Ma C. Chem. Sci. 2022; 13: 775
    • 15c Wang J, Zhou H, Wei J, Liu F, Wang T. Chem. Commun. 2022; 58: 6910
    • 15d Wicker G, Schoch R, Paradies J. Org. Lett. 2021; 23: 3626
  • 16 Borane-mediated α-O C(sp3)–H hydride abstraction has been proposed, see: Liang X.-S, Li R.-D, Sun W, Zhong L, Wang X.-C. ACS Catal. 2022; 12: 9153
  • 17 Borane-mediated NH hydride abstraction has been proposed, see: Dipika, Sharma YB, Pant S, Dhaked DK, Guru MM. Org. Chem. Front. 2022; 9: 3428
  • 18 Chan JZ, Yesilcimen A, Cao M, Zhang Y, Zhang B, Wasa M. J. Am. Chem. Soc. 2020; 142: 16493
  • 19 Tian J.-J, Sun W, Li R.-R, Tian G.-X, Wang X.-C. Angew. Chem. Int. Ed. 2022; 61: e202208427
  • 20 Zhang Z.-Y, Ren J, Zhang M, Xu X.-F, Wang X.-C. Chin. J. Chem. 2021; 39: 1641
  • 21 Zhang Z.-Y, Liu Z.-Y, Guo R.-T, Zhao Y.-Q, Li X, Wang X.-C. Angew. Chem. Int. Ed. 2017; 56: 4028
  • 22 Klose I, Di Mauro G, Kaldre D, Maulide N. Nat. Chem. 2022; 14: 1306
  • 23 Chen Y, Wan H.-L, Huang Y, Liu S, Wang F, Lu C, Nie J, Chen Z, Yang G, Ma C. Org. Lett. 2020; 22: 7797
  • 24 Chang Y, Cao M, Chan JZ, Zhao C, Wang Y, Yang R, Wasa M. J. Am. Chem. Soc. 2021; 143: 2441
  • 25 Fang H, Xie K, Kemper S, Oestreich M. Angew. Chem. Int. Ed. 2021; 60: 8542
  • 26 Zou C.-P, Ma T, Qiao X.-X, Wu X.-X, Li G, He Y, Zhao X.-J. Org. Biomol. Chem. 2023; 21: 4393
  • 27 Zhang M, Wang X.-C. Angew. Chem. Int. Ed. 2021; 60: 17185
  • 28 He Y, Liu Q, Du Z, Xu Y, Cao L, Zhang X, Fan X. J. Org. Chem. 2022; 87: 14840
  • 29 Organoborane-mediated α-N hydride abstraction may also be invoked at one point in the mechanism proposed in: Zhang S, Xu H, He J, Zhang Y. Adv. Synth. Cat. 2021; 363: 5319
  • 30 Hashimoto T, Asada T, Ogoshi S, Hoshimoto Y. Sci. Adv. 2022; 8: eade0189
  • 31 Stephan D. J. Am. Chem. Soc. 2021; 143: 20002

Corresponding Authors

Rebecca L. Melen
Cardiff Catalysis Institute, School of Chemistry, Cardiff University, Translational Research Hub
Maindy Road, Cathays, Cardiff, CF24 4HQ
UK   
Louis C. Morrill
Cardiff Catalysis Institute, School of Chemistry, Cardiff University
Main Building, Park Place, Cardiff, CF10 3AT
UK   
Alexander P. Pulis
School of Chemistry, University of Leicester
Leicester, LE1 7RH
UK   

Publication History

Received: 31 May 2023

Accepted after revision: 14 June 2023

Accepted Manuscript online:
16 June 2023

Article published online:
04 September 2023

© 2023. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution License, permitting copying and reproduction so long as the original work is given appropriate credit. Contents may not be used for commercial purposes or adapted, remixed, transformed or built upon. (https://creativecommons.org/licenses/by/4.0/)

Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany

  • References and Notes

  • 1 Seyden-Penne J. In Reductions by the Alumino- and Borohydrides in Organic Synthesis. Curran D. Wiley-VCH; New York: 1997: 5
  • 2 Sun H.-Y, Hall DG. In Synthesis and Application of Organoboron Compounds . Fernández E, Whiting A. Springer International Publishing; Switzerland: 2015: 221
  • 3 Graham BJ, Raines RT. J. Org. Chem. 2022; in press; DOI DOI: 10.1021/acs.joc.2c01695.
  • 4 Ji X, Liu L. Eur. J. Org. Chem. 2022; e202200853
  • 5 Tan X, Wang H. Chem. Soc. Rev. 2022; 51: 2583
  • 6 Basak S, Winfrey L, Kustiana BA, Melen RL, Morrill LC, Pulis AP. Chem. Soc. Rev. 2021; 50: 3720
  • 7 Ma Y, Loub S.-J, Hou Z. Chem. Soc. Rev. 2021; 50: 1945
  • 8 Kumar G, Roy S, Chatterjee I. Org. Biomol. Chem. 2021; 19: 1230
  • 9 Walker JC. L, Oestreich M. Synlett 2019; 30: 2216
  • 10 Keess S, Oestreich M. Chem. Sci. 2017; 8: 4688
  • 11 Oestreich M. Angew. Chem. Int. Ed. 2016; 55: 494
  • 12 Li L, Yang X. Chin. J. Org. Chem. 2023; 43: 1036
  • 13 Webb JD, Laberge VS, Geier SJ, Stephan DW, Crudden CM. Chem. Eur. J. 2010; 16: 4895
  • 14 Wilkins LC, Santi N, Luk LY. P, Melen RL. Phil. Trans. R. Soc. A 2017; 375: 20170009

    • Organoborane-catalyzed processes that do not include direct hydride abstraction from C–H to B–H, such as pericyclic 1,n-hydride shifts, are beyond the scope of this review. For examples, see:
    • 15a Zhou R, Paradies J. Eur. J. Org. Chem. 2021; 6334
    • 15b Zhang B.-B, Peng S, Wang F, Lu C, Nie J, Chen Z, Yang G, Ma C. Chem. Sci. 2022; 13: 775
    • 15c Wang J, Zhou H, Wei J, Liu F, Wang T. Chem. Commun. 2022; 58: 6910
    • 15d Wicker G, Schoch R, Paradies J. Org. Lett. 2021; 23: 3626
  • 16 Borane-mediated α-O C(sp3)–H hydride abstraction has been proposed, see: Liang X.-S, Li R.-D, Sun W, Zhong L, Wang X.-C. ACS Catal. 2022; 12: 9153
  • 17 Borane-mediated NH hydride abstraction has been proposed, see: Dipika, Sharma YB, Pant S, Dhaked DK, Guru MM. Org. Chem. Front. 2022; 9: 3428
  • 18 Chan JZ, Yesilcimen A, Cao M, Zhang Y, Zhang B, Wasa M. J. Am. Chem. Soc. 2020; 142: 16493
  • 19 Tian J.-J, Sun W, Li R.-R, Tian G.-X, Wang X.-C. Angew. Chem. Int. Ed. 2022; 61: e202208427
  • 20 Zhang Z.-Y, Ren J, Zhang M, Xu X.-F, Wang X.-C. Chin. J. Chem. 2021; 39: 1641
  • 21 Zhang Z.-Y, Liu Z.-Y, Guo R.-T, Zhao Y.-Q, Li X, Wang X.-C. Angew. Chem. Int. Ed. 2017; 56: 4028
  • 22 Klose I, Di Mauro G, Kaldre D, Maulide N. Nat. Chem. 2022; 14: 1306
  • 23 Chen Y, Wan H.-L, Huang Y, Liu S, Wang F, Lu C, Nie J, Chen Z, Yang G, Ma C. Org. Lett. 2020; 22: 7797
  • 24 Chang Y, Cao M, Chan JZ, Zhao C, Wang Y, Yang R, Wasa M. J. Am. Chem. Soc. 2021; 143: 2441
  • 25 Fang H, Xie K, Kemper S, Oestreich M. Angew. Chem. Int. Ed. 2021; 60: 8542
  • 26 Zou C.-P, Ma T, Qiao X.-X, Wu X.-X, Li G, He Y, Zhao X.-J. Org. Biomol. Chem. 2023; 21: 4393
  • 27 Zhang M, Wang X.-C. Angew. Chem. Int. Ed. 2021; 60: 17185
  • 28 He Y, Liu Q, Du Z, Xu Y, Cao L, Zhang X, Fan X. J. Org. Chem. 2022; 87: 14840
  • 29 Organoborane-mediated α-N hydride abstraction may also be invoked at one point in the mechanism proposed in: Zhang S, Xu H, He J, Zhang Y. Adv. Synth. Cat. 2021; 363: 5319
  • 30 Hashimoto T, Asada T, Ogoshi S, Hoshimoto Y. Sci. Adv. 2022; 8: eade0189
  • 31 Stephan D. J. Am. Chem. Soc. 2021; 143: 20002

Zoom Image
Zoom Image
Scheme 1 Borane-mediated hydride abstraction forms Wheland intermediates, pyridinium borohydrides, or iminium borohydrides
Zoom Image
Scheme 2 Borane-mediated hydride abstraction for the direct conversion of N-alkylamines into N-propargylamines; DCE = 1,2-dichloro­ethane
Zoom Image
Scheme 3 Borane-mediated hydride abstraction utilized in the α-furanylation of amines
Zoom Image
Scheme 4 Borane-mediated hydride abstraction utilized in the cyclization of cyclopropyl anilines
Zoom Image
Scheme 5 Borane-mediated hydride abstraction utilized in the diastereoselective construction of azabicycles from enamines and Michael acceptors. Asymmetric Michael addition conditions: (i) l-phenylalanine lithium salt (10 mol%), CH2Cl2, 48 h; or (ii) O-(tert-butyl)-l-threonine (5 mol%), 4-dimethylaminopyridine (5 mol%), sulfamide (5 mol%), toluene, 6–36 h.
Zoom Image
Scheme 6 Borane-mediated hydride abstraction for consecutive β-functionalization and acceptorless dehydrogenation
Zoom Image
Scheme 7 Borane-mediated hydride abstraction for β-functionalization via a Michael-type reaction
Zoom Image
Scheme 8 Borane-mediated hydride abstraction for the enantioselective β-functionalization of amines through a Michael-type addition
Zoom Image
Scheme 9 Borane-mediated hydride abstraction leads to the β,β′-H silylation of tertiary amines for the synthesis of sila-piperidines
Zoom Image
Scheme 10 Borane-mediated hydride abstraction facilitates the β-functionalization of amines with 3H-indol-3-ones
Zoom Image
Scheme 11 Borane-mediated hydride abstraction leads to a [2+2] cycloaddition reaction between the α- and β-positions of amines and alkynes
Zoom Image
Scheme 12 Borane-mediated hydride abstraction in conjunction with gold(I)-catalysis leads to the formal [3+2] cycloaddition of amines and α-alkynylenones
Zoom Image
Scheme 13 Borane-mediated hydride abstraction for the conversion of benzo[c]isoxazoles into functionalized quinolines
Zoom Image
Scheme 14 Borane-mediated hydride abstraction for the acceptorless dehydrogenation of isatin-functionalized pyrrolidines
Zoom Image
Scheme 15 Borane-mediated hydride abstraction for the interconversion of liquid organic hydrogen carriers.