Frustrated Lewis pair

Last updated

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. [1] Many kinds of FLPs have been devised, and many simple substrates exhibit activation. [2] [3]

Contents

The discovery that some FLPs split H2 [4] triggered a rapid growth of research into FLPs. Because of their "unquenched" reactivity, such systems are reactive toward substrates that can undergo heterolysis. For example, many FLPs split hydrogen molecules. Thus, a mixture of tricyclohexylphosphine (PCy3) and tris(pentafluorophenyl)borane reacts with hydrogen to give the respective phosphonium and borate ions:

This reactivity has been exploited to produce FLPs which catalyse hydrogenation reactions. [5]

Small molecule activation

Frustrated Lewis pairs have been shown to activate many small molecules, either by inducing heterolysis or by coordination.

Hydrogen

The discovery that some FLPs are able to split, and therefore activate, H2 [4] triggered a rapid growth of research into this area. The activation and therefore use of H2 is important for many chemical and biological transformations. Using FLPs to liberate H2 is metal-free, this is beneficial due to the cost and limited supply of some transition metals commonly used to activate H2 (Ni, Pd, Pt). [6] FLP systems are reactive toward substrates that can undergo heterolysis (e.g. hydrogen) due to the "unquenched" reactivity of such systems. For example, it has been previously shown that a mixture of tricyclohexylphosphine (PCy3) and tris(pentafluorophenyl)borane reacts with H2 to give the respective phosphonium and borate ions:

In this reaction, PCy3 (the Lewis base) and B(C6F5)3 (the Lewis acid) cannot form an adduct due to the steric hindrance from the bulky cyclohexyl and pentafluorophenyl groups. The proton on the phosphorus and hydride from the borate are now ‘activated’ and can subsequently be ‘delivered’hydrogenation.

Mechanism of dihydrogen activation by FLP

The mechanism for the activation of H2 by FLPs has been discussed for both the intermolecular and intramolecular cases. Intermolecular FLPs are where the Lewis base is a separate molecule to the Lewis acid, it is thought that these individual molecules interact through secondary London dispersion interactions to bring the Lewis base and acid together (a pre-organisational effect) where small molecules may then interact with the FLPs. The experimental evidence for this type of interaction at the molecular level is unclear. However, there is supporting evidence for this type of interaction based on computational density functional theory studies. Intramolecular FLPs are where the Lewis acid and Lewis base are combined in one molecule by a covalent linker. Despite the improved ‘pre-organisational effects’, rigid intramolecular FLP frameworks are thought to have a reduced reactivity to small molecules due to a reduction in flexibility.

Other small molecule substrates

FLPs are also reactive toward many unsaturated substrates beyond H2. Some FLPs react with CO2, specifically in the deoxygenative reduction of CO2 to methane. [7]

Ethene also reacts with FLPs: [8]

For acid-base pairs to behave both nucleophilically and electrophilically at the same time offers a method for the ring-opening of cyclic ethers such as THF, 2,5-dihydrofuran, coumaran, and dioxane. [9]

Use in catalysis

Imine, nitrile and aziridine hydrogenation

Catalytic cycle for reduction of imine to an amine using an FLP FLPs chemdraw 2.svg
Catalytic cycle for reduction of imine to an amine using an FLP

Reduction of imines, nitriles, and aziridines to primary and secondary amines traditionally is effected by metal hydride reagents, e.g. lithium aluminium hydride and sodium cyanoborohydride. Hydrogenations of these unsaturated substrates can be effected by metal-catalyzed reactions. Metal-free catalytic hydrogenation was carried out using the phosphonium borate catalyst (R2PH)(C6F4)BH(C6F5)2 (R = 2,4,6-Me3C6H2) 1. This type of metal-free hydrogenation has the potential to replace high cost metal catalyst.

The mechanism of imine reduction is proposed to involve protonation at nitrogen giving the iminium salt. The basicity of the nitrogen centre determines the rate of reaction. More electron rich imines reduce at faster rates than electron poor imines. The resulting iminium center undergoes nucleophilic attack by the borohydride anion to form the amine. Small amines bind to the borane, quenching further reactions. This problem can be overcome using various methods: 1) Application of elevated temperatures 2) Using sterically bulky imine substituents 3) Protecting the imine with the B(C6F5)3group, which also serves as a Lewis acid promoter. [10]

Enantioselective imine hydrogenation

A chiral boronate Lewis acid derived from (1R)-(+)-camphor forms a frustrated Lewis pair with tBu3P, which is isolable as a salt. This FLP catalyses the enantioselective hydrogenation of some aryl imines in high yield but modest ee (up to 83%).

Asymmetric imine hydrogenation by an FLP Chiara and Liam Wiki Entry.svg
Asymmetric imine hydrogenation by an FLP

Although conceptually interesting, the protocol suffers from lack of generality. It was found that increasing steric bulk of the imine substituents lead to decreased yield and ee of the amine product. methoxy-substituted imines exhibit superior yield and ee's. [10]

Asymmetric hydrosilylations

Frustrated Lewis pairs of chiral alkenylboranes and phosphines are beneficial for asymmetric Piers-type hydrosilylations of 1,2-dicarbonyl compounds and alpha-keto esters, giving high yield and enantioselectivity. However, in comparison to conventional Piers-type hydrosilyations, asymmetric Piers-type hydrosilylations are not as well developed.

In the following example, the chiral alkenylborane is formed in situ from a chiral diyne and the HB(C6F5)2. Heterolytic cleavage of the Si-H bond from PhMe2SiH by the FLP catalyst forms a silylium and hydridoborate ionic complex. [11]

Asymmetric hydrosilylation of a diketone by an FLP Asymmetric Hydrosilylation of Diketone.png
Asymmetric hydrosilylation of a diketone by an FLP

Alkyne hydrogenation

Metal free hydrogenation of unactivated internal alkynes to cis-alkenes is readily achieved using FLP-based catalysts. [12] The condition for this reaction were relatively mild utilising 2 bar of H2. In terms of mechanism, the alkyne material is first hydroborated and then the resulting vinylborane-based FLP can then activate dihydrogen. A protodeborylation step releases the cis-alkene product, which is obtained due to the syn-hydroborylation process, and regenerating the catalyst. While active for alkyne hydrogenation the FLP-based catalysts do not however facilitate the hydrogenation of alkenes to alkanes.

The reaction is a syn-hydroboration, and as a result a high cis selectivity is observed. At the final stage of the catalytic cycle the C6F5 group is cleaved more easily than an alkyl group, causing catalyst degradation rather than alkane release. The catalytic cycle has three steps:

Binding of terminal alkyne to the FLP catalyst FLP Alkyne Binding.png
Binding of terminal alkyne to the FLP catalyst

With internal alkynes, a competitive reaction occurs where the proton bound to the nitrogen can be added to the fluorobenzenes. Therefore, this addition does not proceed that much, the formation of the alkene seems favoured.

But terminal alkynes do not bind to the boron through hydroboration but rather through C-H activation. Thus, the addition of the proton to the alkyne will result in the initial terminal alkyne. Hence this hydrogenation process is not suitable to terminal alkynes and will only give pentafluorobenzene.

The metal free hydrogenation of terminal alkynes to the respective alkenes was recently achieved using a pyridone borane based system. [13] This system activates hydrogen readily at room temperature yielding a pyridone borane complex. [14] Dissociation of this complex allows hydroboration of an alkyne by the free borane. Upon protodeborylation by the free pyridone the cis alkene is generated. Hydrogenation of terminal alkynes is possible with this system, because the C-H activation is reversible and competes with hydrogen activation.

Borylation

Amine-borane FLPs catalyse the borylation of electron-rich aromatic heterocycles (Scheme 1). [15] The reaction is driven by release of hydrogen via C-H activation by the FLP. Aromatic borylations are often used in pharmaceutical development, particularly due to the abundance, low cost and low toxicity of boron compounds compared to noble metals.,

Scheme 1: Mechanism for borylation catalysed by FLP FLP Mechanism.svg
Scheme 1: Mechanism for borylation catalysed by FLP

The substrate for the reaction has two main requirements, strongly linked to the mechanism of borylation. First, the substrate must be electron rich, exemplified by the absence of a reaction with thiophene, whereas its more electron rich derivatives - methoxythiophene and 3,4-ethylenedioxythiophene - can undergo a reaction with the amino-borane. Furthermore, substitution of 1-methylpyrrole (which can react) with the strongly electron withdrawing tertbutyloxycarbonyl (Boc) group at the 2-position completely inhibits the reaction. The second requirement is for the absence of basic amine groups in the substrate, which would otherwise form an unwanted adduct. This can be illustrated by the lack of a reaction with pyrrole, whereas both 1-methyl and N-benzylpyrrole derivatives are able to react.

Further work by the same authors revealed that simply piperidine as the amine R group (as opposed to tetramethylpiperidine, pictured above) accelerated the rate of reaction. Through kinetic and DFT studies the authors proposed that the C-H activation step was more facile than with larger substituents. [16]

Dearomatisation can also be achieved under similar conditions but using N-tosyl indoles. Syn-hyrdoborylated indolines are obtained. [17]

Deromatisation of N-tosyl indole by HBpin Deromatisation of N tosyl indole by HBpin.png
Deromatisation of N-tosyl indole by HBpin

Borylation of S-H bonds in thiols by a dehydrogenative process has also been observed. Alcohols and amines such as tert-Butanol and tert-Butylamine form stable products that prevent catalysis due to a strong π-bond between the N/O atom's lone pair and boron, whereas the same is not true for thiols, thus allowing for successful catalysis. In addition, successful borylation of Se-H bonds has been achieved. In all cases, the formation of H2 gas is a strong driving force for the reactions. [18]

Carbon capture

FLP chemistry is conceptually relevant to carbon capture. [19] Both an intermolecular (Scheme 1) and intramolecular (Scheme 2) FLP consisting of a phosphine and a borane were used to selectively capture and release carbon dioxide. When a solution of the FLP was covered by an atmosphere of CO2 at room temperature, the FLP-CO2 compound immediately precipitated as a white solid. [19] [20]

Scheme 1: Intermolecular FLP CO2 capture and release FLPCO2Capture1.jpg
Scheme 1: Intermolecular FLP CO2 capture and release

Heating the intermolecular FLP-CO2 compound in bromobenzene at 80 °C under vacuum for 5 hours caused the release of around half of the CO2 and regenerating the two constituent components of the FLP. After several more hours of sitting at room temperature under vacuum, total release of CO2 and FLP regeneration had occurred. [19]

Scheme 2: Intramolecular FLP CO2 capture and release FLPCO2Capture.jpg
Scheme 2: Intramolecular FLP CO2 capture and release

The intramolecular FLP-CO2 compound by contrast was stable as a solid at room temperature but fully decomposed at temperatures above -20 °C as a solution in dichloromethane releasing CO2 and regenerating the FLP molecule. [19]

This method of FLP carbon capture can be adapted to work in flow chemistry systems. [21]

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.

<span class="mw-page-title-main">Wilkinson's catalyst</span> Chemical compound

Wilkinson's catalyst (chlorido­tris(triphenylphosphene)­rhodium(I)) is a coordination complex of rhodium with the formula [RhCl(PPh3)3], where 'Ph' denotes a phenyl group. It is a red-brown colored solid that is soluble in hydrocarbon solvents such as benzene, and more so in tetrahydrofuran or chlorinated solvents such as dichloromethane. The compound is widely used as a catalyst for hydrogenation of alkenes. It is named after chemist and Nobel laureate Sir Geoffrey Wilkinson, who first popularized its use.

In organic chemistry, hydroboration refers to the addition of a hydrogen-boron bond to certain double and triple bonds involving carbon. This chemical reaction is useful in the organic synthesis of organic compounds.

In chemistry, transfer hydrogenation is a chemical reaction involving the addition of hydrogen to a compound from a source other than molecular H2. It is applied in laboratory and industrial organic synthesis to saturate organic compounds and reduce ketones to alcohols, and imines to amines. It avoids the need for high-pressure molecular H2 used in conventional hydrogenation. Transfer hydrogenation usually occurs at mild temperature and pressure conditions using organic or organometallic catalysts, many of which are chiral, allowing efficient asymmetric synthesis. It uses hydrogen donor compounds such as formic acid, isopropanol or dihydroanthracene, dehydrogenating them to CO2, acetone, or anthracene respectively. Often, the donor molecules also function as solvents for the reaction. A large scale application of transfer hydrogenation is coal liquefaction using "donor solvents" such as tetralin.

<span class="mw-page-title-main">Tris(pentafluorophenyl)borane</span> Chemical compound

Tris(pentafluorophenyl)borane, sometimes referred to as "BCF", is the chemical compound (C6F5)3B. It is a white, volatile solid. The molecule consists of three pentafluorophenyl groups attached in a "paddle-wheel" manner to a central boron atom; the BC3 core is planar. It has been described as the “ideal Lewis acid” because of its high thermal stability and the relative inertness of the B-C bonds. Related fluoro-substituted boron compounds, such as those containing B−CF3 groups, decompose with formation of B-F bonds. Tris(pentafluorophenyl)borane is thermally stable at temperatures well over 200 °C, resistant to oxygen, and water-tolerant.

Organophosphines are organophosphorus compounds with the formula PRnH3−n, where R is an organic substituent. These compounds can be classified according to the value of n: primary phosphines (n = 1), secondary phosphines (n = 2), tertiary phosphines (n = 3). All adopt pyramidal structures. Organophosphines are generally colorless, lipophilic liquids or solids. The parent of the organophosphines is phosphine (PH3).

<span class="mw-page-title-main">Hydroamination</span> Addition of an N–H group across a C=C or C≡C bond

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. In the ideal case, hydroamination is atom economical and green. Amines are common in fine-chemical, pharmaceutical, and agricultural industries. 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.

Boroles represent a class of molecules known as metalloles, which are heterocyclic 5-membered rings. As such, they can be viewed as structural analogs of cyclopentadiene, pyrrole or furan, with boron replacing a carbon, nitrogen and oxygen atom respectively. They are isoelectronic with the cyclopentadienyl cation C5H+5(Cp+) and comprise four π electrons. Although Hückel's rule cannot be strictly applied to borole, it is considered to be antiaromatic due to having 4 π electrons. As a result, boroles exhibit unique electronic properties not found in other metalloles.

<span class="mw-page-title-main">Shvo catalyst</span> Chemical compound

The Shvo catalyst is an organoruthenium compound that catalyzes the hydrogenation of polar functional groups including aldehydes, ketones and imines. The compound is of academic interest as an early example of a catalyst for transfer hydrogenation that operates by an "outer sphere mechanism". Related derivatives are known where p-tolyl replaces some of the phenyl groups. Shvo's catalyst represents a subset of homogeneous hydrogenation catalysts that involves both metal and ligand in its mechanism.

<span class="mw-page-title-main">Hydrogenation of carbon–nitrogen double bonds</span>

In chemistry, the hydrogenation of carbon–nitrogen double bonds is the addition of the elements of dihydrogen (H2) across a carbon–nitrogen double bond, forming amines or amine derivatives. Although a variety of general methods have been developed for the enantioselective hydrogenation of ketones, methods for the hydrogenation of carbon–nitrogen double bonds are less general. Hydrogenation of imines is complicated by both syn/anti isomerization and tautomerization to enamines, which may be hydrogenated with low enantioselectivity in the presence of a chiral catalyst. Additionally, the substituent attached to nitrogen affects both the reactivity and spatial properties of the imine, complicating the development of a general catalyst system for imine hydrogenation. Despite these challenges, methods have been developed that address particular substrate classes, such as N-aryl, N-alkyl, and endocyclic imines.

<span class="mw-page-title-main">Boranylium ions</span>

In chemistry, a boranylium ion is an inorganic cation with the chemical formula BR+
2
, where R represents a non-specific substituent. Being electron-deficient, boranylium ions form adducts with Lewis bases. Boranylium ions have historical names that depend on the number of coordinated ligands:

Dehydrogenation of amine-boranes or dehydrocoupling of amine-boranes is a chemical process in main group and organometallic chemistry wherein dihydrogen is released by the coupling of two or more amine-borane adducts. This process is of due to the potential of using amine-boranes for hydrogen storage.

In chemistry, the Gutmann–Beckett method is an experimental procedure used by chemists to assess the Lewis acidity of molecular species. Triethylphosphine oxide is used as a probe molecule and systems are evaluated by 31P-NMR spectroscopy. In 1975, Viktor Gutmann used 31P-NMR spectroscopy to parameterize Lewis acidity of solvents by acceptor numbers (AN). In 1996, Michael A. Beckett recognised its more generally utility and adapted the procedure so that it could be easily applied to molecular species, when dissolved in weakly Lewis acidic solvents. The term Gutmann–Beckett method was first used in chemical literature in 2007.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

<span class="mw-page-title-main">Sulfinylamine</span> Type of organosulfur compound

Sulfinylamines are organosulfur compounds with the formula RNSO where R = an organic substituent. These compounds are, formally speaking, derivatives of HN=S=O, i.e. analogues of sulfur dioxide and of sulfur diimide. A common example is N-sulfinylaniline. Sulfinyl amines are dienophile. They undergo [2+2] cycloaddition to ketenes.

Germanium(II) hydrides, also called germylene hydrides, are a class of Group 14 compounds consisting of low-valent germanium and a terminal hydride. They are also typically stabilized by an electron donor-acceptor interaction between the germanium atom and a large, bulky ligand.

Metal-ligand cooperativity (MLC) is a mode of reactivity in which a metal and ligand of a complex are both involved in the bond breaking or bond formation of a substrate during the course of a reaction. This ligand is an actor ligand rather than a spectator, and the reaction is generally only deemed to contain MLC if the actor ligand is doing more than leaving to provide an open coordination site. MLC is also referred to as "metal-ligand bifunctional catalysis." Note that MLC is not to be confused with cooperative binding.

Heteroatomic multiple bonding between group 13 and group 15 elements are of great interest in synthetic chemistry due to their isoelectronicity with C-C multiple bonds. Nevertheless, the difference of electronegativity between group 13 and 15 leads to different character of bondings comparing to C-C multiple bonds. Because of the ineffective overlap between p𝝅 orbitals and the inherent lewis acidity/basicity of group 13/15 elements, the synthesis of compounds containing such multiple bonds is challenging and subject to oligomerization. The most common example of compounds with 13/15 group multiple bonds are those with B=N units. The boron-nitrogen-hydride compounds are candidates for hydrogen storage. In contrast, multiple bonding between aluminium and nitrogen Al=N, Gallium and nitrogen (Ga=N), boron and phosphorus (B=P), or boron and arsenic (B=As) are less common.

In organic chemistry, carboboration describes an addition of both a carbon and a boron moiety to certain carbon-containing double and triple bonds, such as alkenes, alkynes, and allenes.

<i>N</i>-Heterocyclic olefins Neutral heterocyclic compound

An N-heterocyclic olefin (NHO) is a neutral heterocyclic compound with a highly polarized, electron-rich C=C olefin attached to a heterocycle made up of two nitrogen atoms. A derivative of N-heterocyclic carbenes (NHCs), NHO was first synthesized in 1961 by Horst Böhme and Fritz Soldan, but the term NHO was not used until 2011 by Eric Rivard and coworkers. Since its discovery, NHOs have been applied in organocatalysis, metal ligation, and polymerization.

References

  1. Stephan, Douglas W (2008). "Frustrated Lewis pairs: a concept for new reactivity and catalysis". Org. Biomol. Chem. 6 (9): 1535–1539. doi:10.1039/b802575b. PMID   18421382. Closed Access logo transparent.svg
  2. Stephan, Douglas W.; Erker, Gerhard (2010). "Frustrated Lewis Pairs: Metal-free Hydrogen Activation and More". Angewandte Chemie International Edition. 49 (1): 46–76. doi:10.1002/anie.200903708. ISSN   1433-7851. PMID   20025001. Closed Access logo transparent.svg
  3. Stephan, Douglas W.; Erker, Gerhard (2017). "Frustrated Lewis pair chemistry". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 375 (2101): 20170239. Bibcode:2017RSPTA.37570239S. doi:10.1098/rsta.2017.0239. ISSN   1364-503X. PMC   5540845 . PMID   28739971. Open Access logo PLoS transparent.svg
  4. 1 2 Welch, Gregory C.; Juan, Ronan R. San; Masuda, Jason D.; Stephan, Douglas W. (2006). "Reversible, Metal-Free Hydrogen Activation". Science. 314 (5802): 1124–1126. Bibcode:2006Sci...314.1124W. doi:10.1126/science.1134230. ISSN   0036-8075. PMID   17110572. S2CID   20333088. Closed Access logo transparent.svg
  5. Lam, Jolie; Szkop, Kevin M.; Mosaferi, Eliar; Stephan, Douglas W. (2018). "FLP catalysis: main group hydrogenations of organic unsaturated substrates". Chemical Society Reviews. 41 (13): 3592–3612. doi:10.1039/C8CS00277K. PMID   30178796. S2CID   206130644.
  6. Welch, Gregory C.; Juan, Ronan R. San; Masuda, Jason D.; Stephan, Douglas W. (2006-11-17). "Reversible, Metal-Free Hydrogen Activation". Science. 314 (5802): 1124–1126. Bibcode:2006Sci...314.1124W. doi:10.1126/science.1134230. ISSN   0036-8075. PMID   17110572. S2CID   20333088.
  7. Berkefeld, Andreas; Piers, Warren E.; Parvez, Masood (2010-08-11). "Tandem Frustrated Lewis Pair/Tris(pentafluorophenyl)borane-Catalyzed Deoxygenative Hydrosilylation of Carbon Dioxide". Journal of the American Chemical Society. 132 (31): 10660–10661. doi:10.1021/ja105320c. ISSN   0002-7863. PMID   20681691.
  8. Stephan, D. W. (2009). ""Frustrated Lewis Pairs": A New Strategy to Small Molecule Activation and Hydrogenation Catalysis". Dalton Trans (17): 3129–3136. doi:10.1039/b819621d. PMID   19421613. Closed Access logo transparent.svg
  9. Tochertermam, W (1966). "Structures and Reactions of Organic ate-Complexes". Angew. Chem. Int. Ed. 5 (4): 351–171. doi:10.1002/anie.196603511.
  10. 1 2 Chen, Dianjun; Wang, Yutian; Klankermayer, Jürgen (2010-12-03). "Enantioselective Hydrogenation with Chiral Frustrated Lewis Pairs". Angewandte Chemie International Edition. 49 (49): 9475–9478. doi:10.1002/anie.201004525. ISSN   1521-3773. PMID   21031385.
  11. Ren, Xiaoyu; Du, Haifeng (2016-01-15). "Chiral Frustrated Lewis Pairs Catalyzed Highly Enantioselective Hydrosilylations of 1,2-Dicarbonyl Compounds". Journal of the American Chemical Society. 138 (3): 810–813. doi:10.1021/jacs.5b13104. ISSN   0002-7863. PMID   26750998.
  12. Chernichenko, Konstantin; Madarász, Ádám; Pápai, Imre; Nieger, Martin; Leskelä, Markku; Repo, Timo (2013). "A frustrated-Lewis-pair approach to catalytic reduction of alkynes to cis-alkenes" (PDF). Nature Chemistry . 5 (8): 718–723. Bibcode:2013NatCh...5..718C. doi:10.1038/nchem.1693. PMID   23881505. S2CID   22474399. Closed Access logo transparent.svg
  13. Wech, Felix; Hasenbeck, Max; Gellrich, Urs (2020-09-18). "Semihydrogenation of Alkynes Catalyzed by a Pyridone Borane Complex: Frustrated Lewis Pair Reactivity and Boron–Ligand Cooperation in Concert". Chemistry – A European Journal. 26 (59): 13445–13450. doi: 10.1002/chem.202001276 . ISSN   0947-6539. PMC   7693047 . PMID   32242988.
  14. Gellrich, Urs (2018). "Reversible Hydrogen Activation by a Pyridonate Borane Complex: Combining Frustrated Lewis Pair Reactivity with Boron-Ligand Cooperation". Angewandte Chemie International Edition. 57 (17): 4779–4782. doi:10.1002/anie.201713119. ISSN   1521-3773. PMID   29436754.
  15. Légaré, Marc A.; Courtmanche, Marc A.; Rochette, Étienne; Fontaine, Frédéric G. (2015-07-30). "Metal-free catalytic C-H bond activation and borylation of heteroarenes". Science. 349 (6247): 513–516. Bibcode:2015Sci...349..513L. doi:10.1126/science.aab3591. hdl: 20.500.11794/30087 . ISSN   0036-8075. PMID   26228143. S2CID   206638394. Closed Access logo transparent.svg
  16. Légaré Lavergne, Julien; Jayaraman, Arumugam; Misal Castro, Luis C.; Rochette, Étienne; Fontaine, Frédéric-Georges (2017-10-06). "Metal-Free Borylation of Heteroarenes Using Ambiphilic Aminoboranes: On the Importance of Sterics in Frustrated Lewis Pair C–H Bond Activation". Journal of the American Chemical Society. 139 (41): 14714–14723. doi:10.1021/jacs.7b08143. hdl: 20.500.11794/30144 . ISSN   0002-7863. PMID   28901757.
  17. Jayaraman, Arumugam; Misal Castro, Luis C.; Desrosiers, Vincent; Fontaine, Frédéric-Georges (2018). "Metal-free borylative dearomatization of indoles: exploring the divergent reactivity of aminoborane C–H borylation catalysts". Chemical Science. 9 (22): 5057–5063. doi:10.1039/c8sc01093e. ISSN   2041-6520. PMC   5994747 . PMID   29938036.
  18. Rochette, Étienne; Boutin, Hugo; Fontaine, Frédéric-Georges (2017-06-30). "Frustrated Lewis Pair Catalyzed S–H Bond Borylation". Organometallics. 36 (15): 2870–2876. doi:10.1021/acs.organomet.7b00346. hdl: 20.500.11794/30088 . ISSN   0276-7333.
  19. 1 2 3 4 Mömming, Cornelia M.; Otten, Edwin; Kehr, Gerald; Fröhlich, Roland; Grimme, Stefan; Stephan, Douglas W.; Erker, Gerhard (2009-08-24). "Reversible Metal-Free Carbon Dioxide Binding by Frustrated Lewis Pairs" (PDF). Angewandte Chemie International Edition. 48 (36): 6643–6646. doi:10.1002/anie.200901636. ISSN   1433-7851. PMID   19569151. S2CID   28050646.
  20. Stephan, Douglas W.; Erker, Gerhard (2015-05-14). "Frustrated Lewis Pair Chemistry: Development and Perspectives". Angewandte Chemie International Edition. 54 (22): 6400–6441. doi:10.1002/anie.201409800. ISSN   1433-7851. PMID   25974714.
  21. Abolhasani, Milad; Günther, Axel; Kumacheva, Eugenia (2014-06-24). "Microfluidic Studies of Carbon Dioxide". Angewandte Chemie International Edition. 53 (31): 7992–8002. doi:10.1002/anie.201403719. ISSN   1433-7851. PMID   24961230.