Protodeboronation

Last updated
Simple protodeboronation scheme Simple protodeboronation scheme.png
Simple protodeboronation scheme

Protodeboronation, or protodeborylation is a chemical reaction involving the protonolysis of a boronic acid (or other organoborane compound) in which a carbon-boron bond is broken and replaced with a carbon-hydrogen bond. Protodeboronation is a well-known undesired side reaction, and frequently associated with metal-catalysed coupling reactions that utilise boronic acids (see Suzuki reaction). [1] For a given boronic acid, the propensity to undergo protodeboronation is highly variable and dependent on various factors, such as the reaction conditions employed and the organic substituent of the boronic acid.

Contents

The deliberate protodeboronation of boronic acids (and derivatives) have been applied to some synthetic procedures, such as the installation of a stereospecific proton at chiral centers, [2] and also in purification procedures, such as the removal of unwanted regioisomeric boronic acid by-products. [3]

Recent mechanistic studies have revealed a variety of protodeboronation pathways in aqueous media, and have demonstrated the reaction pH (and subsequently the boronic acid speciation) to be an important factor in understanding the modes of protodeboronation. [4] [5]

A general reaction scheme for the protodeboronation of boronic acids Protodeboronation Scheme.png
A general reaction scheme for the protodeboronation of boronic acids

History

One of the earliest reports of protodeboronation was made by Ainley and Challenger, who were the first researchers to explore the reactivity of boronic acids with common chemical reagents. [6] They reported the reaction of phenylboronic acid in water (140-150 °C) to afford the protodeboronated product, benzene, after 40 hours.

Initial synthetic applications of protodeboronation were found alongside the discovery of the hydroboration reaction, in which sequential hydroboration-protodeboronation reactions were used to convert alkynes or alkenes into the corresponding saturated compounds. [7] Beyond this synthetic application, protodeboronation was rarely noted or valued in other chemical processes throughout the early 20th century. However, in more recent years, protodeboronation has emerged as a problematic side reaction with many chemical processes that utilise boronic acids. In particular, boronic acids have become increasingly important reagents for the facile construction of carbon-carbon and carbon-heteroatom bonds via metal-catalysed cross-coupling reactions. This has resulted in an increased usage of boronic acids, and subsequently followed by an increased number of reports concerning problematic protodeboronation. Many boronic acids are now commercially available and many novel boronic acids and derivatives are constantly in pursuit.

Many efforts have been put towards mitigating undesired protodeboronation in cross-coupling reactions. Catalyst design and optimisation has led the way for very efficient systems that can undergo rapid catalytic turnover. [8] This increases the rate of productive reaction and thus subdues unwanted decomposition pathways such as protodeboronation. Cross-coupling reactions have also been accelerated with metal additives such as silver [9] [10] [11] [12] and copper. [13] [14]

Boronic acid derivatives have also been used to suppress protodeboronation. [15] MIDA boronate esters and organotrifluoroborates have both been utilised in "slow release" strategies, in which the reaction conditions are optimised to provide a slow release of boronic acid. This protocol has proved useful in the cross-coupling of some notoriously unstable boronic acids, such as the 2-pyridine boronic acid. [16] [17] This ensures that the boronic acid concentration is low during the cross-coupling reaction, which in turn minimises the potential for side reactions.

Reaction mechanism

Simple non-basic boronic acids

The mechanism of protodeboronation was initially investigated by Kuivila in the 1960s, long before the discovery of the Suzuki reaction and the popularisation of boronic acids. Their studies focused on the protodeboronation of some substituted aromatic boronic acids in aqueous conditions, and they reported the presence of two distinct mechanisms; a general acid-catalysed and a specific base-catalysed mechanism. [18] [19] The acid-catalysed process is dependent on a reaction between boronic acid and an acid, such as sulfuric acid. On the other hand, the base-catalysed process arises from a pre-equilibrium between boronic acid and hydroxide to form the corresponding boronate, this is usually followed by a rate-limiting reaction between boronate and water (acting as the proton source). Substrates that display only these two modes of protodeboronation (typically simple aromatic and alkyl boronic acids) are generally very stable in neutral pH solution, where both acid- and base-catalysed processes are minimised. For aromatic boronic acids bearing electron-withdrawing substituents, there is a competing dissociative mechanism involving generation of a transient aryl anion. These substrates are stabilized by acidic conditions. [5]

Acid and base catalyzed protodeboronation Kuivila LloydJones mechanisms.png
Acid and base catalyzed protodeboronation

Basic heteroaromatic boronic acids

Basic heteroaromatic boronic acids (boronic acids that contain a basic nitrogen atom, such as 2-pyridine boronic acid) display additional protodeboronation mechanisms. [4] A key finding shows the speciation of basic heteroaromatic boronic acids to be analogous to that of simple amino acids, with zwitterionic species forming under neutral pH conditions. For the 2-pyridine boronic acid, the zwitterionic compound is responsible for its rapid protodeboronation under neutral pH, through a unimolecular fragmentation of the C-B bond. In fact, the addition of acid (H+) or hydroxide (OH-) acts to attenuate protodeboronation by shifting the speciation away from the reactive zwitterion.

It is important to note that not all basic heteroaromatic boronic acids are reactive through a zwitterionic intermediate.

Scheme for the speciation of 2-pyridine boronic acid in aqueous solution 2-pyridyl speciation.png
Scheme for the speciation of 2-pyridine boronic acid in aqueous solution

See also

Related Research Articles

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.

The Suzuki reaction is an organic reaction, classified as a cross-coupling reaction, where the coupling partners are a boronic acid and an organohalide and the catalyst is a palladium(0) complex. It was first published in 1979 by Akira Suzuki, and he shared the 2010 Nobel Prize in Chemistry with Richard F. Heck and Ei-ichi Negishi for their contribution to the discovery and development of palladium-catalyzed cross-couplings in organic synthesis. This reaction is also known as the Suzuki–Miyaura reaction or simply as the Suzuki coupling. It is widely used to synthesize polyolefins, styrenes, and substituted biphenyls. Several reviews have been published describing advancements and the development of the Suzuki reaction. The general scheme for the Suzuki reaction is shown below, where a carbon-carbon single bond is formed by coupling a halide (R1-X) with an organoboron species (R2-BY2) using a palladium catalyst and a base.

<span class="mw-page-title-main">Organoboron chemistry</span> Study of compounds containing a boron-carbon bond

Organoboron chemistry or organoborane chemistry is the chemistry of organoboron compounds or organoboranes, which are chemical compounds of boron and carbon that are organic derivatives of borane (BH3), for example trialkyl boranes..

<span class="mw-page-title-main">Bamford–Stevens reaction</span>

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

The Hiyama coupling is a palladium-catalyzed cross-coupling reaction of organosilanes with organic halides used in organic chemistry to form carbon–carbon bonds. This reaction was discovered in 1988 by Tamejiro Hiyama and Yasuo Hatanaka as a method to form carbon-carbon bonds synthetically with chemo- and regioselectivity. The Hiyama coupling has been applied to the synthesis of various natural products.

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

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

<span class="mw-page-title-main">Directed ortho metalation</span> Chemical reaction

Directed ortho metalation (DoM) is an adaptation of electrophilic aromatic substitution in which electrophiles attach themselves exclusively to the ortho- position of a direct metalation group or DMG through the intermediary of an aryllithium compound. The DMG interacts with lithium through a hetero atom. Examples of DMG's are the methoxy group, a tertiary amine group and an amide group.The compound can be produced by directed lithiation of anisole.

In organic chemistry, the Buchwald–Hartwig amination is a chemical reaction for the synthesis of carbon–nitrogen bonds via the palladium-catalyzed coupling reactions of amines with aryl halides. Although Pd-catalyzed C-N couplings were reported as early as 1983, Stephen L. Buchwald and John F. Hartwig have been credited, whose publications between 1994 and the late 2000s established the scope of the transformation. The reaction's synthetic utility stems primarily from the shortcomings of typical methods for the synthesis of aromatic C−N bonds, with most methods suffering from limited substrate scope and functional group tolerance. The development of the Buchwald–Hartwig reaction allowed for the facile synthesis of aryl amines, replacing to an extent harsher methods while significantly expanding the repertoire of possible C−N bond formation.

<span class="mw-page-title-main">Boronic acid</span> Organic compound of the form R–B(OH)2

A boronic acid is an organic compound related to boric acid in which one of the three hydroxyl groups is replaced by an alkyl or aryl group. As a compound containing a carbon–boron bond, members of this class thus belong to the larger class of organoboranes.

<span class="mw-page-title-main">Phenylboronic acid</span> Chemical compound

Phenylboronic acid or benzeneboronic acid, abbreviated as PhB(OH)2 where Ph is the phenyl group C6H5-, is a boronic acid containing a phenyl substituent and two hydroxyl groups attached to boron. Phenylboronic acid is a white powder and is commonly used in organic synthesis. Boronic acids are mild Lewis acids which are generally stable and easy to handle, making them important to organic synthesis.

<span class="mw-page-title-main">Liebeskind–Srogl coupling</span>

The Liebeskind–Srogl coupling reaction is an organic reaction forming a new carbon–carbon bond from a thioester and a boronic acid using a metal catalyst. It is a cross-coupling reaction. This reaction was invented by and named after Jiri Srogl from the Academy of Sciences, Czech Republic, and Lanny S. Liebeskind from Emory University, Atlanta, Georgia, USA. There are three generations of this reaction, with the first generation shown below. The original transformation used catalytic Pd(0), TFP = tris(2-furyl)phosphine as an additional ligand and stoichiometric CuTC = copper(I) thiophene-2-carboxylate as a co-metal catalyst. The overall reaction scheme is shown below.

In organic chemistry, a cross-coupling reaction is a reaction where two different fragments are joined. Cross-couplings are a subset of the more general coupling reactions. Often cross-coupling reactions require metal catalysts. One important reaction type is this:

<span class="mw-page-title-main">Akira Suzuki (chemist)</span> Japanese chemist (born 1930)

Akira Suzuki is a Japanese chemist and Nobel Prize Laureate (2010), who first published the Suzuki reaction, the organic reaction of an aryl- or vinyl-boronic acid with an aryl- or vinyl-halide catalyzed by a palladium(0) complex, in 1979.

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

Organotrifluoroborates are organoboron compounds that contain an anion with the general formula [RBF3]. They can be thought of as protected boronic acids, or as adducts of carbanions and boron trifluoride. Organotrifluoroborates are tolerant of air and moisture and are easy to handle and purify. They are often used in organic synthesis as alternatives to boronic acids (RB(OH)2), boronate esters (RB(OR′)2), and organoboranes (R3B), particularly for Suzuki-Miyaura coupling.

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

PEPPSI is an abbreviation for pyridine-enhanced precatalyst preparation stabilization and initiation. It refers to a family of commercially available palladium catalysts developed around 2005 by Prof. Michael G. Organ and co-workers at York University, which can accelerate various carbon-carbon and carbon-heteroatom bond forming cross-coupling reactions. In comparison to many alternative palladium catalysts, Pd-PEPPSI-type complexes are stable to air and moisture and are relatively easy to synthesize and handle.

In chemistry, metal-catalysed hydroboration is a reaction used in organic synthesis. It is one of several examples of homogeneous catalysis.

Decarboxylative cross coupling reactions are chemical reactions in which a carboxylic acid is reacted with an organic halide to form a new carbon-carbon bond, concomitant with loss of CO2. Aryl and alkyl halides participate. Metal catalyst, base, and oxidant are required.

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.

Miyaura borylation, also known as the Miyaura borylation reaction, is a named reaction in organic chemistry that allows for the generation of boronates from vinyl or aryl halides with the cross-coupling of bis(pinacolato)diboron in basic conditions with a catalyst such as PdCl2(dppf). The resulting borylated products can be used as coupling partners for the Suzuki reaction.

Norio Miyaura was a Japanese organic chemist. He was a professor of graduate chemical engineering at Hokkaido University. His major accomplishments surrounded his work in cross-coupling reactions / conjugate addition reactions of organoboronic acids and addition / coupling reactions of diborons and boranes. He is also the co-author of Cross-Coupling Reactions: A Practical Guide with M. Nomura E. S.. Miyaura was a world-known and accomplished researcher by the time he retired and so, in 2007, he won the Japan Chemical Society Award.

References

  1. Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials, 2nd ed.; Hall, D., Ed; Wiley VCH: Weinheim, Germany, 2011; Vols. 1 and 2, pp 1-133.
  2. Roesner, S.; Blair, D. J.; Aggarwal, V. K. "Enantioselective installation of adjacent tertiary benzylic stereocentres using lithiation–borylation–protodeboronation methodology. Application to the synthesis of bifluranol and fluorohexestrol". Chem. Sci. 2015, 6, 3718. doi : 10.1039/C4SC03901G
  3. Klingensmith, L. M.; Bio, M. M.; Moniz, G. A., "Selective protodeboronation: synthesis of 4-methyl-2-thiopheneboronic anhydride and demonstration of its utility in Suzuki-Miyaura reactions". Tetrahedron Lett. 2007, 48, 8242. doi : 10.1016/j.tetlet.2007.09.060
  4. 1 2 Cox, P. A.; Leach, A. G.; Campbell, A. D.; Lloyd-Jones, G. C. "Protodeboronation of Heteroaromatic, Vinyl, and Cyclopropyl Boronic Acids: pH–Rate Profiles, Autocatalysis, and Disproportionation" J. Am. Chem. Soc. 2016, 138, 9145. doi : 10.1021/jacs.6b03283
  5. 1 2 Cox, P. A.; Reid, M.; Leach A. G.; Campbell, A. D.; King, E. J.; Lloyd-Jones, G. C. "Base-Catalyzed Aryl-B(OH)2 Protodeboronation Revisited: From Concerted Proton Transfer to Liberation of a Transient Aryl Anion". J. Am. Chem. Soc. 2017, 139, 13156. doi : 10.1021/jacs.7b07444
  6. Ainley, A. D.; Challenger, F. "Studies of the boron-carbon linkage" J. Chem. Soc. 1930, 2171.
  7. Brown, H. C.; Zweifel, G. "Hydroboration. XI. The Hydroboration of Acetylenes—A Convenient Conversion of Internal Acetylenes into cis-Olefins and of Terminal Acetylenes into Aldehydes" J. Am. Chem. Soc. 1961, 83 (18), 3834. doi : 10.1021/ja01479a024
  8. Billingsley, K. L.; Anderson, K. W.; Buchwald, S. L. "A Highly Active Catalyst for Suzuki–Miyaura Cross-Coupling Reactions of Heteroaryl Compounds". Angew. Chemie Int. Ed. 2006, 45, 3484. doi : 10.1002/anie.200600493
  9. Chen, J.; Cammers-Goodwin, A. "2-(Fluorophenyl)pyridines by the Suzuki-Miyaura method: Ag2O accelerates coupling over undesired ipso substitution (SNAr) of fluorine" Tetrahedron Lett. 2003, 44, 1503. doi : 10.1016/S0040-4039(02)02793-4
  10. Nishihara, Y.; Onodera, H.; Osakada, K. "Synthesis and structural characterization of the first unsymmetrical from transmetallation between 2,4,6-trifluorophenylboronic acid and trans-Pd(C6F5)I(PEt3)2" Chem. Commun. 2004, 192. doi : 10.1039/b308741g
  11. Korenaga, T.; Kosaki, T.; Fukumura, R.; Ema, T.; Sakai, T. "Suzuki-Miyaura coupling reaction using pentafluorophenylboronic acid". Org. Lett. 2005, 7, 4915. doi : 10.1021/ol051866i
  12. Imao, D.; Glasspoole, B. W.; Laberge, V. S.; Crudden, C. M. "Cross Coupling Reactions of Chiral Secondary Organoboronic Esters With Retention of Configuration". J. Am. Chem. Soc. 2009, 131 (14), 5024. doi : 10.1021/ja8094075
  13. Leconte, N.; Keromnes-Wuillaume, A.; Suzenet, F.; Guillaumet, G. "Efficient Palladium-Catalyzed Synthesis of Unsymmetrical (Het)aryltetrazines". Synlett 2007, 204. doi : 10.1055/s-2007-967991
  14. Deng, J. Z.; Paone, D. V.; Ginnetti, A. T.; Kurihara, H.; Dreher, S. D.; Weissman, S. A.; Stauffer, S. R.; Burgey, C. S. "Copper-facilitated Suzuki reactions: application to 2-heterocyclic boronates". Org. Lett. 2009, 11, 345. doi : 10.1021/ol802556f
  15. Lennox, A. J. J.; Lloyd-Jones, G. C. "The Slow-Release Strategy in Suzuki–Miyaura Coupling". Isr. J. Chem. 2010, 50, 664. doi : 10.1002/ijch.201000074
  16. Knapp, D. M.; Gillis, E. P.; Burke, M. D. "A general solution for unstable boronic acids: slow-release cross-coupling from air-stable MIDA boronates.". J. Am. Chem. Soc. 2009, 131, 6961. doi : 10.1021/ja901416p
  17. Molander, G. A.; Ellis, N. "Organotrifluoroborates:  Protected Boronic Acids That Expand the Versatility of the Suzuki Coupling Reaction". Acc. Chem. Res. 2007, 40, 275. doi : 10.1021/ar050199q
  18. Kuivila, H. G.; Reuwer, J. F.; Mangravite, J. A. "Electrophilic Displacement Reactions. X. General Acid Catalysis in the Protodeboronation of Areneboronic Acids". Can. J. Chem. 1963, 41, 3081. doi : 10.1021/ja01470a028
  19. Kuivila, H. G.; Nahabedian, K. V. "Electrophilic displacement reactions. XV. Kinetics and Mechanism of the Base-catalyzed Protodeboronation of Areneboronic Acids". J. Am. Chem. Soc. 1961, 83, 2159. doi : 10.1139/v63-451