Boronic acid

Last updated

The general structure of a boronic acid, where R is a substituent. Boronic-acid-2D.png
The general structure of a boronic acid, where R is a substituent.

A boronic acid is an organic compound related to boric acid (B(OH)3) in which one of the three hydroxyl groups (−OH) is replaced by an alkyl or aryl group (represented by R in the general formula R−B(OH)2). [1] As a compound containing a carbon–boron bond, members of this class thus belong to the larger class of organoboranes.

Contents

Boronic acids act as Lewis acids. Their unique feature is that they are capable of forming reversible covalent complexes with sugars, amino acids, hydroxamic acids, etc. (molecules with vicinal, (1,2) or occasionally (1,3) substituted Lewis base donors (alcohol, amine, carboxylate)). The pKa of a boronic acid is ~9, but they can form tetrahedral boronate complexes with pKa ~7. They are occasionally used in the area of molecular recognition to bind to saccharides for fluorescent detection or selective transport of saccharides across membranes.

Boronic acids are used extensively in organic chemistry as chemical building blocks and intermediates predominantly in the Suzuki coupling. A key concept in its chemistry is transmetallation of its organic residue to a transition metal.

The compound bortezomib with a boronic acid group is a drug used in chemotherapy. The boron atom in this molecule is a key substructure because through it certain proteasomes are blocked that would otherwise degrade proteins. Boronic acids are known to bind to active site serines and are part of inhibitors for porcine pancreatic lipase, [2] subtilisin [3] and the protease Kex2. [4] Furthermore, boronic acid derivatives constitute a class of inhibitors for human acyl-protein thioesterase 1 and 2, which are cancer drug targets within the Ras cycle. [5]

The boronic acid functional group is reputed to have low inherent toxicity. This is one of the reasons for the popularity of the Suzuki coupling in the development and synthesis of pharmaceutical agents. However, a significant fraction of commonly used boronic acids and their derivatives were recently found to gives a positive Ames test and act as chemical mutagens. The mechanism of mutagenicity is thought to involve the generation of organic radicals via oxidation of the boronic acid by atmospheric oxygen. [6]

Structure and synthesis

In 1860, Edward Frankland was the first to report the preparation and isolation of a boronic acid. Ethylboronic acid was synthesized by a two-stage process. First, diethylzinc and triethyl borate reacted to produce triethylborane. This compound then oxidized in air to form ethylboronic acid. [7] [8] [9] Several synthetic routes are now in common use, and many air-stable boronic acids are commercially available.

Boronic acids typically have high melting points. They are prone to forming anhydrides by loss of water molecules, typically to give cyclic trimers.

Examples of boronic acids
Boronic acidRStructure Molar mass CAS number Melting point °C
Phenylboronic acid Phenyl Phenylboronic acid.png 121.9398-80-6216–219
2-Thienylboronic acid Thiophen 2-Thienylboronic acid.svg 127.966165-68-0138–140
Methylboronic acid Methyl Methylboronic acid.svg 59.8613061-96-691–94
cis-Propenylboronic acid propene Cis-Propenylboronic acid.svg 85.907547-96-865–70
trans-Propenylboronic acid propene Trans-Propenylboronic acid.svg 85.907547-97-9123–127

Synthesis

Boronic acids can be obtained via several methods. The most common way is reaction of organometallic compounds based on lithium or magnesium (Grignards) with borate esters. [10] [11] [12] [13] For example, phenylboronic acid is produced from phenylmagnesium bromide and trimethyl borate followed by hydrolysis [14]

PhMgBr + B(OMe)3 → PhB(OMe)2 + MeOMgBr
PhB(OMe)2 + 2 H2O → PhB(OH)2 + 2 MeOH

Another method is reaction of an arylsilane (RSiR3) with boron tribromide (BBr3) in a transmetallation to RBBr2 followed by acidic hydrolysis.

A third method is by palladium catalysed reaction of aryl halides and triflates with diboronyl esters in a coupling reaction known as the Miyaura borylation reaction. An alternative to esters in this method is the use of diboronic acid or tetrahydroxydiboron ([B(OH2)]2). [15] [16] [17]

Boronic esters (also named boronate esters)

Boronic esters are esters formed between a boronic acid and an alcohol.

Comparison between boronic acids and boronic esters
CompoundGeneral formulaGeneral structure
Boronic acidRB(OH)2
Boronic-acid-2D.png
Boronic esterRB(OR)2
Boronate-ester-2D.png

The compounds can be obtained from borate esters [18] by condensation with alcohols and diols. Phenylboronic acid can be selfcondensed to the cyclic trimer called triphenyl anhydride or triphenylboroxin. [19]

Examples of boronic esters
Boronic esterDiolStructural formula Molar mass CAS number Boiling point (°C)
Allylboronic acid pinacol ester pinacol Allylboronic acid pinacol ester.svg 168.0472824-04-550–53 (5 mmHg)
Phenyl boronic acid trimethylene glycol ester trimethylene glycol Phenylboronic acid trimethylene glycol ester.svg 161.994406-77-3106 (2 mm Hg)
Diisopropoxymethylborane isopropanol Diisopropoxymethylborane.svg 144.0286595-27-9105 -107

Compounds with 5-membered cyclic structures containing the C–O–B–O–C linkage are called dioxaborolanes and those with 6-membered rings dioxaborinanes.

Organic chemistry applications

Suzuki coupling reaction

Boronic acids are used in organic chemistry in the Suzuki reaction. In this reaction the boron atom exchanges its aryl group with an alkoxy group from palladium.

 

 

 

 

(1)

Chan–Lam coupling

In the Chan–Lam coupling the alkyl, alkenyl or aryl boronic acid reacts with a N–H or O–H containing compound with Cu(II) such as copper(II) acetate and oxygen and a base such as pyridine [20] [21] forming a new carbon–nitrogen bond or carbon–oxygen bond for example in this reaction of 2-pyridone with trans-1-hexenylboronic acid:

ChanLamCoupling.png

The reaction mechanism sequence is deprotonation of the amine, coordination of the amine to the copper(II), transmetallation (transferring the alkyl boron group to copper and the copper acetate group to boron), oxidation of Cu(II) to Cu(III) by oxygen and finally reductive elimination of Cu(III) to Cu(I) with formation of the product. Direct reductive elimination of Cu(II) to Cu(0) also takes place but is very slow. In catalytic systems oxygen also regenerates the Cu(II) catalyst.

Liebeskind–Srogl coupling

In the Liebeskind–Srogl coupling a thiol ester is coupled with a boronic acid to produce a ketone.

Conjugate addition

The boronic acid organic residue is a nucleophile in conjugate addition also in conjunction with a metal. In one study the pinacol ester of allylboronic acid is reacted with dibenzylidene acetone in such a conjugate addition: [22]

BoronicAcidConjugateAddition.png
The catalyst system in this reaction is tris(dibenzylideneacetone)dipalladium(0) / tricyclohexylphosphine.

Another conjugate addition is that of gramine with phenylboronic acid catalyzed by cyclooctadiene rhodium chloride dimer: [23]

BoronicAcidGramineReaction.png

Oxidation

Boronic esters are oxidized to the corresponding alcohols with base and hydrogen peroxide (for an example see: carbenoid)

Homologation

In this reaction dichloromethyllithium converts the boronic ester into a boronate. A Lewis acid then induces a rearrangement of the alkyl group with displacement of the chlorine group. Finally an organometallic reagent such as a Grignard reagent displaces the second chlorine atom effectively leading to insertion of an RCH2 group into the C-B bond. Another reaction featuring a boronate alkyl migration is the Petasis reaction.

Electrophilic allyl shifts

Allyl boronic esters engage in electrophilic allyl shifts very much like silicon pendant in the Sakurai reaction. In one study a diallylation reagent combines both [25] [note 1] :

DoubleAllylationByBoronicAcid.png

Hydrolysis

Hydrolysis of boronic esters back to the boronic acid and the alcohol can be accomplished in certain systems with thionyl chloride and pyridine. [26] Aryl boronic acids or esters may be hydrolyzed to the corresponding phenols by reaction with hydroxylamine at room temperature. [27]

C–H coupling reactions

The diboron compound bis(pinacolato)diboron [28] reacts with aromatic heterocycles [29] or simple arenes [30] to an arylboronate ester with iridium catalyst [IrCl(COD)]2 (a modification of Crabtree's catalyst) and base 4,4′-di-tert-butyl-2,2′-bipyridine in a C-H coupling reaction for example with benzene:

IridiumCHactivationMiyauraHartwig.svg

In one modification the arene reacts using only a stoichiometric equivalent rather than a large excess using the cheaper pinacolborane: [31]

IridiumAreneBorylation.svg

Unlike in ordinary electrophilic aromatic substitution (EAS) where electronic effects dominate, the regioselectivity in this reaction type is solely determined by the steric bulk of the iridium complex. This is exploited in a meta-bromination of m-xylene which by standard AES would give the ortho product: [32] [note 2]

MetahalogenationArylborylationMurphy2007.svg

Protonolysis

Protodeboronation 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). 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:

A simple protodeboronation in acidic medium Protodeboronation Scheme.png
A simple protodeboronation in acidic medium

Supramolecular chemistry

Saccharide recognition

An example of a fluorescent complex of a diboronic acid and tartaric acid Tony D James enantioselective diboronic acid fluorescent sensor.png
An example of a fluorescent complex of a diboronic acid and tartaric acid

The covalent pair-wise interaction between boronic acids and hydroxy groups as found in alcohols and acids is rapid and reversible in aqueous solutions. The equilibrium established between boronic acids and the hydroxyl groups present on saccharides has been successfully employed to develop a range of sensors for saccharides. [34] One of the key advantages with this dynamic covalent strategy [35] lies in the ability of boronic acids to overcome the challenge of binding neutral species in aqueous media. If arranged correctly, the introduction of a tertiary amine within these supramolecular systems will permit binding to occur at physiological pH and allow signalling mechanisms such as photoinduced electron transfer mediated fluorescence emission to report the binding event.

Potential applications for this research include blood glucose monitoring systems to help manage diabetes mellitus. As the sensors employ an optical response, monitoring could be achieved using minimally invasive methods, one such example is the investigation of a contact lens that contains a boronic acid based sensor molecule to detect glucose levels within ocular fluids. [36]

Notes

  1. In this sequence the boronic ester allyl shift is catalyzed by boron trifluoride. In the second step the hydroxyl group is activated as a leaving group by conversion to a triflate by triflic anhydride aided by 2,6-lutidine. The final product is a vinyl cyclopropane. Note: ee stands for enantiomeric excess
  2. In situ second step reaction of boronate ester with copper(II) bromide

Related Research Articles

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

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.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

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

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

<span class="mw-page-title-main">Organocopper chemistry</span> Compound with carbon to copper bonds

Organocopper chemistry is the study of the physical properties, reactions, and synthesis of organocopper compounds, which are organometallic compounds containing a carbon to copper chemical bond. They are reagents in organic chemistry.

In organic chemistry, the Kumada coupling is a type of cross coupling reaction, useful for generating carbon–carbon bonds by the reaction of a Grignard reagent and an organic halide. The procedure uses transition metal catalysts, typically nickel or palladium, to couple a combination of two alkyl, aryl or vinyl groups. The groups of Robert Corriu and Makoto Kumada reported the reaction independently in 1972.

<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.

<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.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

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.

The Chan–Lam coupling reaction – also known as the Chan–Evans–Lam coupling is a cross-coupling reaction between an aryl boronic acid and an alcohol or an amine to form the corresponding secondary aryl amines or aryl ethers, respectively. The Chan–Lam coupling is catalyzed by copper complexes. It can be conducted in air at room temperature. The more popular Buchwald–Hartwig coupling relies on the use of palladium.

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

The Catellani reaction was discovered by Marta Catellani and co-workers in 1997. The reaction uses aryl iodides to perform bi- or tri-functionalization, including C-H functionalization of the unsubstituted ortho position(s), followed a terminating cross-coupling reaction at the ipso position. This cross-coupling cascade reaction depends on the ortho-directing transient mediator, norbornene.

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

Protodeboronation, or protodeborylation is a chemical reaction involving the protonolysis of a boronic acid 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. 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.

In cross-coupling reactions, the component reagents are called cross-coupling partners or simply coupling partners. These reagents can be further classified according to their nucleophilic vs electrophilic character:

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. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Boronic Acids ". doi : 10.1351/goldbook.B00714
  2. Garner, C. W. (10 June 1980). "Boronic acid inhibitors of porcine pancreatic lipase". The Journal of Biological Chemistry. 255 (11): 5064–5068. doi: 10.1016/S0021-9258(19)70749-2 . ISSN   0021-9258. PMID   7372625.
  3. Lindquist, R. N.; Terry, C. (January 1974). "Inhibition of subtilisin by boronic acids, potential analogs of tetrahedral reaction intermediates". Archives of Biochemistry and Biophysics. 160 (1): 135–144. doi:10.1016/s0003-9861(74)80018-4. ISSN   0003-9861. PMID   4364061.
  4. Holyoak, Todd; Wilson, Mark A.; Fenn, Timothy D.; Kettner, Charles A.; Petsko, Gregory A.; Fuller, Robert S.; Ringe, Dagmar (10 June 2003). "2.4 A resolution crystal structure of the prototypical hormone-processing protease Kex2 in complex with an Ala-Lys-Arg boronic acid inhibitor". Biochemistry. 42 (22): 6709–6718. doi:10.1021/bi034434t. ISSN   0006-2960. PMID   12779325.
  5. Zimmermann, Tobias J.; Bürger, Marco; Tashiro, Etsu; Kondoh, Yasumitsu; Martinez, Nancy E.; Görmer, Kristina; Rosin-Steiner, Sigrid; Shimizu, Takeshi; Ozaki, Shoichiro (2 January 2013). "Boron-based inhibitors of acyl protein thioesterases 1 and 2". ChemBioChem. 14 (1): 115–122. doi:10.1002/cbic.201200571. ISSN   1439-7633. PMID   23239555. S2CID   205557212.
  6. Hansen, Marvin M.; Jolly, Robert A.; Linder, Ryan J. (29 July 2015). "Boronic Acids and Derivatives—Probing the Structure–Activity Relationships for Mutagenicity". Organic Process Research & Development. 19 (11): 1507–1516. doi:10.1021/acs.oprd.5b00150. ISSN   1083-6160.
  7. Frankland, E.; Duppa, B. F. (1860). "Vorläufige Notiz über Boräthyl". Justus Liebigs Ann. Chem. 115 (3): 319. doi:10.1002/jlac.18601150324.
  8. Frankland, E.; Duppa, B. (1860). "On Boric Ethide". Proceedings of the Royal Society . 10: 568–570. doi: 10.1098/rspl.1859.0112 .
  9. Frankland, E. (1862). "On a new series of organic compounds containing boron". J. Chem. Soc. 15: 363–381. Bibcode:1862RSPT..152..167F. doi:10.1039/JS8621500363.
  10. Dennis G. Hall, ed. (2005). Boronic Acids. Wiley. ISBN   978-3-527-30991-7.
  11. Example: Kristensen, Jesper Langgaard; Lysén, Morten; Vedsø, Per; Begtrup, Mikael (2005). "Synthesis of Ortho Substituted Arylboronic Esters by in situ Traping of Unstable Lithio Intermediates: 2-(5,5-Dimethyl-1,3,2-dioxaborinan-2-yl)benzoic acid ethyl ester". Organic Syntheses . 81: 134.; Collective Volume, vol. 11, pp. 1015 prep= v81p0134
  12. Example: Li, Wenjie; Nelson, Dorian P.; Jensen, Mark S.; Scott Hoerrner, R.; Cai, Dongwei; Larsen, Robert D. (2005). "Synthesis of 3-Pyridylboronic Acid and its Pinacol Ester. Application of 3-Pyridylboronic acid in Suzuki Coupling to Prepare 3-Pyridin-3-ylquinoline". Organic Syntheses . 81: 89.; Collective Volume, vol. 11, p. 393
  13. Charette, André B.; Lebel, Hélène (1999). "(2S,3S)-(+)-(3-Phenylcyclopropyl)methanol". Organic Syntheses . 76: 86.; Collective Volume, vol. 10, p. 613
  14. Washburn, Robert M.; Levens, Ernest; Albright, Charles F.; Billig, Franklin A. (1959). "Benzeneboronic anhydride". Organic Syntheses . 39: 3.; Collective Volume, vol. 4, p. 68
  15. Pilarski, Lukasz T.; Szabó, Kálmán J. (2011). "Palladium-Catalyzed Direct Synthesis of Organoboronic Acids". Angewandte Chemie International Edition. 50 (36): 8230–8232. doi:10.1002/anie.201102384. PMID   21721088.
  16. Molander, Gary A.; Trice, Sarah L. J.; Dreher, Spencer D. (2010). "Palladium-Catalyzed, Direct Boronic Acid Synthesis from Aryl Chlorides: A Simplified Route to Diverse Boronate Ester Derivatives". Journal of the American Chemical Society. 132 (50): 17701–17703. doi:10.1021/ja1089759. PMC   3075417 . PMID   21105666.
  17. Ishiyama, Tatsuo; Murata, Miki; Miyaura, Norio (1 November 1995). "Palladium(0)-Catalyzed Cross-Coupling Reaction of Alkoxydiboron with Haloarenes: A Direct Procedure for Arylboronic Esters". The Journal of Organic Chemistry. 60 (23): 7508–7510. doi:10.1021/jo00128a024.
  18. Kidwell, R. L.; Murphy, M.; Darling, S. D. (1969). "Phenols: 6-Methoxy-2-Naphthol". Organic Syntheses . 49: 90.; Collective Volume, vol. 5, p. 918
  19. Washburn, Robert M.; Levens, Ernest; Albright, Charles F.; Billig, Franklin A. (1959). "Benzeneboronic anhydride". Organic Syntheses . 39: 3.; Collective Volume, vol. 4, p. 68
  20. Chan, Dominic M.T. (2003). "Copper promoted C–N and C–O bond cross-coupling with phenyl and pyridylboronates". Tetrahedron Letters. 44 (19): 3863–3865. doi:10.1016/S0040-4039(03)00739-1.
  21. Lam, Patrick Y.S. (2003). "Copper-promoted/catalyzed C–N and C–O bond cross-coupling with vinylboronic acid and its utilities". Tetrahedron Letters. 44 (26): 4927–4931. doi:10.1016/S0040-4039(03)01037-2.
  22. Sieber, Joshua D. (2007). "Catalytic Conjugate Addition of Allyl Groups to Styryl-Activated Enones". Journal of the American Chemical Society. 129 (8): 2214–2215. CiteSeerX   10.1.1.624.3153 . doi:10.1021/ja067878w. PMID   17266312.
  23. Gabriela (2007). "Benzylic Substitution of Gramines with Boronic Acids and Rhodium or Iridium Catalysts †". Organic Letters. 9 (6): 961–964. doi:10.1021/ol063042m. PMID   17305348.
  24. Matteson, Donald S. (1986). "99% Chirally selective synthesis via pinanediol boronic esters: insect pheromones, diols, and an amino alcohol". Journal of the American Chemical Society. 108 (4): 810–819. doi:10.1021/ja00264a039.
  25. Peng, Feng (2007). "Simple, Stable, and Versatile Double-Allylation Reagents for the Stereoselective Preparation of Skeletally Diverse Compounds". Journal of the American Chemical Society. 129 (11): 3070–3071. doi:10.1021/ja068985t. PMID   17315879.
  26. Matteson, Donald S. (2003). "New asymmetric syntheses with boronic esters and fluoroboranes" (PDF). Pure Appl. Chem. 75 (9): 1249–1253. doi:10.1351/pac200375091249. S2CID   15944330.
  27. Kianmehr, Ebrahim; Yahyaee, Maryam; Tabatabai, Katayoun (2007). "A mild conversion of arylboronic acids and their pinacolyl boronate esters into phenols using hydroxylamine". Tetrahedron Letters. 48 (15): 2713–2715. doi:10.1016/j.tetlet.2007.02.069.
  28. Ishiyama, Tatsuo; Murata, Miki; Ahiko, Taka-aki; Miyaura, Norio (2000). "Bis(pinacolato)diboron". Organic Syntheses . 77: 176.; Collective Volume, vol. 10, p. 115
  29. Takagi, Jun (2002). "Iridium-catalyzed C–H coupling reaction of heteroaromatic compounds with bis(pinacolato)diboron: regioselective synthesis of heteroarylboronates". Tetrahedron Letters. 43 (32): 5649–5651. doi:10.1016/S0040-4039(02)01135-8. hdl: 2115/56222 .
  30. Ishiyama, Tatsuo (2002). "Mild Iridium-Catalyzed Borylation of Arenes. High Turnover Numbers, Room Temperature Reactions, and Isolation of a Potential Intermediate". Journal of the American Chemical Society. 124 (3): 390–391. doi:10.1021/ja0173019. PMID   11792205.
  31. Ishiyama, Tatsuo (2003). "Room temperature borylation of arenes and heteroarenes using stoichiometric amounts of pinacolborane catalyzed by iridium complexes in an inert solvent". Chemical Communications (23): 2924–5. doi:10.1039/b311103b. hdl: 2115/56377 . PMID   14680243.
  32. Murphy, Jaclyn M. (2007). "Meta Halogenation of 1,3-Disubstituted Arenes via Iridium-Catalyzed Arene Borylation". Journal of the American Chemical Society. 129 (50): 15434–15435. doi:10.1021/ja076498n. PMID   18027947.
  33. Zhao, Jianzhang; Davidson, Matthew G.; Mahon, Mary F.; Kociok-Köhn, Gabriele; James, Tony D. (2004). "An Enantioselective Fluorescent Sensor for Sugar Acids". J. Am. Chem. Soc. 126 (49): 16179–16186. doi:10.1021/ja046289s. PMID   15584754.
  34. James, Tony D.; Phillips, Marcus D.; Shinkai, Seiji (2006). Boronic Acids in Saccharide Recognition. doi:10.1039/9781847557612. ISBN   978-0-85404-537-2.
  35. Rowan, Stuart J.; Cantrill, Stuart J.; Cousins, Graham R. L.; Sanders, Jeremy K. M.; Stoddart, J. Fraser (2002). "Dynamic Covalent Chemistry". Angewandte Chemie International Edition. 41 (6): 898–952. doi:10.1002/1521-3773(20020315)41:6<898::AID-ANIE898>3.0.CO;2-E. PMID   12491278.
  36. US 6850786,Wayne Front March,"Ocular analyte sensor",issued 2005-02-01