Fact-checked by Grok 2 weeks ago

Hydroboration

Hydroboration is an involving the addition of (BH₃) to an or to form an organoborane intermediate, which can then be oxidized to yield alcohols (from alkenes) or carbonyl compounds (from alkynes). The addition proceeds with anti-Markovnikov regioselectivity, where the atom attaches to the less substituted carbon of the multiple bond, and syn stereoselectivity, with both boron and hydrogen adding to the same face of the . This process, discovered in 1956 by and B. C. Subba Rao during investigations into reductions, provides a mild, selective method for functionalizing unsaturated hydrocarbons, contrasting with traditional electrophilic additions that follow . In the full hydroboration-oxidation sequence, the intermediate organoborane is treated with alkaline (H₂O₂/NaOH), replacing the boron with a hydroxyl group while retaining the anti-Markovnikov orientation and syn addition, thus enabling the direct synthesis of primary alcohols from terminal alkenes under neutral conditions. Brown's pioneering work on hydroboration, which made organoboranes readily accessible for the first time, revolutionized synthetic by facilitating the preparation of alcohols, halides, amines, and carbon-carbon bonds with exceptional control over regiochemistry and . This contribution earned Brown the in 1979, shared with Georg Wittig for related phosphorus chemistry. The reaction's versatility extends to asymmetric variants using chiral boranes, enhancing its utility in complex molecule synthesis.

Borane Reagents

Borane Complexes and Adducts

(BH₃) is a highly reactive acid that does not exist as a stable in the gas or in , instead tending to dimerize to (B₂H₆) or form oligomeric structures unless coordinated to a . To facilitate safe handling and storage, BH₃ forms stable adducts with electron-donating ligands such as ethers, amines, and sulfides, which donate a to the vacant orbital on , resulting in tetrahedral coordination at the center. These adducts serve as convenient sources of BH₃ for synthetic applications, mitigating the hazards associated with gaseous . Key examples include the borane-tetrahydrofuran (BH₃·THF), which is a colorless solution typically supplied at 1 M concentration in THF ; it exhibits good in solvents but is thermally unstable, requiring below 0°C to prevent , with a of several weeks at . The borane-dimethyl (BH₃·SMe₂) is a neat, colorless with a BH₃ concentration of approximately 10 M, offering superior thermal (stable for months at ) and broader in non-polar solvents compared to the THF , making it preferable for large-scale reactions. The ammonia-borane (BH₃·NH₃) is a white crystalline solid, stable under ambient conditions with high thermal up to 60–70°C before , and it shows moderate in (up to 11.4 M) and polar solvents, though less so in hydrocarbons. All three adducts are commercially available from chemical suppliers such as and . In solution, these adducts undergo partial to generate the active, uncoordinated BH₃ species essential for reactivity: \ce{BH3 \cdot L <=> BH3 + L} where L represents the coordinating (e.g., THF, SMe₂, or NH₃); the extent of depends on the and ligand strength, with weaker donors like sulfides providing more free BH₃. The development of these borane adducts traces back to the pioneering work of in the early 1950s, who synthesized the first BH₃·THF complex as part of efforts to make reagents more accessible and less hazardous than for , enabling widespread adoption in hydroboration processes. Borane adducts are highly reactive and require careful handling under inert atmospheres (e.g., nitrogen or argon) to avoid ignition or explosion, as they are pyrophoric upon exposure to air and react violently with water or protic solvents to liberate hydrogen gas; they are also toxic, causing severe irritation to eyes, skin, and respiratory tract, and are classified as flammable liquids or solids with low flash points (e.g., 18°C for BH₃·SMe₂). Proper use involves glove box manipulation or Schlenk techniques, with spills neutralized using aqueous sodium hypochlorite.

Preparation Methods

Borane reagents for hydroboration are typically prepared as stable adducts of BH₃, such as BH₃·THF and BH₃·SMe₂, since free BH₃ is unstable and tends to dimerize to diborane (B₂H₆). The most common laboratory-scale routes involve the reduction of boron trifluoride (BF₃) derivatives with sodium borohydride (NaBH₄), which generates diborane as an intermediate that is then trapped by a Lewis base to form the desired adduct. A key step in these preparations is the formation of diborane via the reaction of NaBH₄ with BF₃ etherate in a suitable solvent, following the stoichiometry: $3 \mathrm{NaBH_4} + 4 \mathrm{BF_3 \cdot OEt_2} \rightarrow 2 \mathrm{B_2H_6} + 4 \mathrm{Et_2O} + 3 \mathrm{NaBF_4} This reaction is typically conducted at 0–5°C in ether or glyme solvents to control the exothermic process and minimize side reactions. For BH₃·THF, the diborane is generated directly in tetrahydrofuran (THF), where it dissociates and coordinates to the solvent, yielding a 1 M solution of the adduct suitable for immediate use in hydroboration. Similarly, BH₃·SMe₂ is prepared by generating diborane in diglyme (or directly in methyl sulfide) from NaBH₄ and BF₃·OEt₂, followed by addition of dimethyl sulfide (Me₂S) to form the stable complex, which offers advantages in volatility and ease of handling over ether-based adducts. Less commonly, borane adducts can be derived from boric acid precursors, such as through thermal decomposition or reduction of borate esters, though these methods are more suited to specialized applications rather than routine hydroboration. To circumvent the hazards of isolating gaseous or unstable BH₃, generation is widely employed during hydroboration reactions. This involves adding NaBH₄ to a of the (or ) substrate and BF₃·OEt₂ (or other activators like I₂ or H₂SO₄) in the reaction solvent, allowing controlled release of BH₃ for immediate addition to the unsaturated compound. Such approaches enhance safety and efficiency, particularly for sensitive substrates. For scalability, optimized procedures using higher glymes like triglyme or tetraglyme as solvents enable quantitative generation at larger scales (up to several moles) by improving and reaction rates, with yields exceeding 95% under controlled conditions. Continuous flow has also been adapted for generation, integrating NaBH₄/BF₃ reactions in microreactors to produce BH₃ solutions on demand, minimizing handling risks and enabling industrial quantities for hydroboration processes. Post-2000 advancements emphasize eco-friendly methods, including solvent-free preparations where NaBH₄ reacts with BF₃ in the absence of ether solvents to directly afford , reducing emissions. Additionally, media have been explored for dissolving BF₃ and facilitating NaBH₄ reductions, offering recyclable, non-volatile alternatives that maintain high yields while aligning with principles.

Mechanism and Selectivity

Concerted Addition Mechanism

The hydroboration reaction proceeds as a stepwise of (BH₃) across the carbon-carbon of an , ultimately forming a trialkylborane product under mild conditions. The process begins with the insertion of one into a B-H bond of BH₃, yielding a monoalkylborane (RCH₂CH₂BH₂), which then reacts with two additional equivalents of to form the dialkyl- and finally the trialkylborane ((RCH₂CH₂)₃B). This multi-step sequence occurs without the formation of charged intermediates, distinguishing it from ionic mechanisms like oxymercuration. The addition step is concerted and stereospecific, involving a four-center cyclic where the and hydrogen atoms from the B-H bond simultaneously bond to adjacent carbons of the , resulting in syn addition. In this , the electrophilic atom, bearing a partial positive charge due to its empty p-orbital, is approached by the π-electron density of the , with preferentially attaching to the less substituted carbon to minimize steric hindrance and stabilize the developing partial negative charge on the more substituted carbon. This electronic and steric preference, further influenced by nonstatistical dynamic effects in the reaction pathway, leads to the characteristic anti-Markovnikov regiochemistry. The key initial addition can be represented as: \ce{R-CH=CH2 + BH3 ->[concerted][four-center TS] R-CH2-CH2-BH2} Subsequent additions follow analogous pathways to complete the trialkylborane. (DFT) computational studies have elucidated the energy profile of this , revealing low barriers (typically 5-15 kcal/ depending on the computational ) for the BH₃-alkene addition, which aligns with the reaction's rapidity at and its insensitivity to typical rearrangements. These calculations confirm the concerted nature by showing no discrete intermediates along the , with the featuring partial B-C and H-C bond formation and B-H bond cleavage. Kinetic isotope effect experiments using BD₃ instead of BH₃ yield a primary KIE (k_H/k_D) of about 2-3, further evidencing that B-H bond breaking occurs in the rate-determining . The reaction rate is modulated by solvent and temperature, with coordinating solvents like (THF) accelerating the process by forming stable BH₃·THF adducts that enhance Lewis acidity without altering the mechanistic pathway. In THF, hydroboration of terminal alkenes proceeds efficiently at 0-25°C, consistent with the ordered four-center . Non-coordinating solvents, such as , slow the rate due to weaker stabilization of the , while elevated temperatures (up to 50°C) are sometimes used for less reactive alkenes but are generally unnecessary for the classic reaction.

Regio- and Stereoselectivity Principles

Hydroboration reactions exhibit pronounced , with the electrophilic atom preferentially attaching to the less substituted, more electron-rich carbon of the , resulting in anti-Markovnikov orientation. This behavior stems from a of steric factors, which disfavor to more hindered positions, and electronic factors, whereby the partial positive charge on in the is better accommodated at the terminal carbon, augmented by dynamic effects along the reaction trajectory. For simple terminal alkenes such as propene, hydroboration with BH₃ affords approximately 90% anti-Markovnikov product. The degree of regioselectivity is modulated by the steric bulk of the borane reagent and the electronic nature of alkene substituents. Less hindered reagents like BH₃ provide good but not absolute selectivity, whereas bulkier dialkylboranes, such as 9-borabicyclo[3.3.1]nonane (9-BBN), enhance anti-Markovnikov preference to >99% for terminal alkenes by amplifying steric repulsion at substituted carbons. Electron-withdrawing groups on the alkene can partially reverse this regioselectivity by stabilizing the transition state for boron addition to the more substituted carbon through inductive effects. In conjugated systems, regioselectivity is often diminished due to delocalization effects that alter electron density. For instance, styrene yields about 80% anti-Markovnikov product (primary alcohol after oxidation) with BH₃, compared to higher selectivity in aliphatic terminal alkenes. The following table summarizes regioselectivity data for representative alkenes using BH₃:
Alkene% Anti-Markovnikov Product
Propene90
Styrene80
Stereoselectivity in hydroboration is exclusively , with and adding to the same face of the π-bond in a concerted fashion, preserving the geometry and avoiding intermediates. This leads to cis-vinylboranes from alkynes and, upon oxidation, anti-Markovnikov alcohols with no . For alkenes bearing chiral centers, the syn addition produces diastereomerically pure products, such as erythro adducts from (Z)- or threo from (E)-, enabling stereocontrolled synthesis.

Classic Hydroboration Reactions

Of Alkenes

Hydroboration of proceeds via the of (BH₃), typically generated from or as a complex such as BH₃·THF, to the carbon-carbon , yielding trialkylboranes as the primary products. The reaction is conducted under mild conditions at in ether-based solvents like (THF) or , with 1 equivalent of BH₃ reacting with up to 3 equivalents of alkene to afford the trialkylborane quantitatively. This stepwise occurs rapidly, often completing within minutes to hours depending on the substrate, and is highly efficient for simple alkenes. The scope of the reaction encompasses a broad range of alkenes, with alkenes providing high yields exceeding 95% for the formation of primary trialkylboranes, while internal alkenes deliver moderate yields of 70–90%, influenced by steric hindrance at the . The process tolerates many functional groups, including halides, ethers, and esters, without interference; notably, isolated carbonyl groups remain unreactive under these mild conditions, allowing selective hydroboration of alkenes in their presence. A representative example is the hydroboration of , where three molecules add to BH₃ to form tri-n-hexylborane in near-quantitative yield: \ce{3 CH3(CH2)4CH=CH2 + BH3 -> (CH3(CH2)5)3B} This product can be isolated if desired or advanced directly. For dienes and polyenes, controlled mono-hydroboration is achievable by employing 1 equivalent of BH₃ relative to the substrate, limiting addition to a single double bond and enabling the synthesis of monoalkylboranes for subsequent selective transformations. Practical workup of the trialkylboranes involves either distillation under reduced pressure for isolation or immediate subjection to oxidation protocols to generate alcohols, streamlining synthetic sequences.

Of Alkynes

Hydroboration of alkynes involves the syn of reagents to the , typically under mild conditions similar to those used for alkenes but proceeding at a slower rate. With BH₃, the reaction often proceeds to double addition, particularly for alkynes, affording gem-diborylalkanes. To achieve selective monoaddition and yield cis-vinylboranes with greater than 95% cis , sterically hindered dialkylboranes such as ((sia)₂BH) are employed due to the concerted . For terminal alkynes, the reaction displays excellent , with preferentially attaching to the less substituted carbon in an anti-Markovnikov fashion, producing clean cis-vinyl of the general form R-CH=CH-BR'₂ when using hindered . A representative example is the hydroboration of 1-hexyne with ()₂BH, which yields the corresponding (Z)-1-hexenylborane intermediate in high yield. In contrast, internal s exhibit reduced and are more susceptible to double hydroboration, leading to gem-diborylalkanes upon treatment with excess BH₃. This over-addition arises from the lower steric differentiation between the alkyne carbons, making monoaddition challenging without specialized reagents. For example, undergoes hydroboration with BH₃ to form a of vinylborane and the gem-diboryl product, with the latter predominating under non-controlled conditions. Vinylboranes derived from alkyne hydroboration serve as versatile synthetic intermediates, functioning as precursors to cis-alkenes via protonolysis or to enolates for further carbon-carbon bond formation. These transformations highlight the utility of hydroboration in constructing stereodefined unsaturated systems.

Transformations of Organoboranes

Oxidation to Alcohols

The oxidation of organoboranes represents a pivotal step in hydroboration, transforming the C-B bonds into C-O bonds to yield alcohols with high fidelity to the regiochemistry and established during the hydroboration phase. This process, first demonstrated by and coworkers in the late 1950s, provides a mild, acid-free route to anti-Markovnikov alcohols, circumventing the rearrangements and side products common in traditional acid-catalyzed methods. The standard procedure involves treating the organoborane intermediate, typically generated from an and (BH₃), with aqueous (H₂O₂) in the presence of (NaOH) at temperatures between 0°C and 25°C. This two-step sequence—hydroboration followed by oxidation—converts terminal alkenes of the form R-CH=CH₂ into primary alcohols R-CH₂-CH₂-OH, with the hydroxyl group attaching to the less substituted carbon. For example, yields as the major product under these conditions. The reaction is typically carried out in a protic solvent like (THF) or diglyme, with the oxidation performed or after workup of the organoborane. Mechanistically, the oxidation proceeds via nucleophilic attack by hydroperoxide anion (HOO⁻) on the atom of the trialkylborane, displacing the s and forming a trialkylborate . This is followed by migration of an from to the adjacent oxygen, accompanied by oxidation of the to facilitate the bond rearrangement, ultimately cleaving the C-B bond while retaining the C-O linkage. The process occurs with complete retention of at carbon, preserving the syn from the initial hydroboration. The overall transformation can be represented by the equation: (RCH_2CH_2)_3B + 3H_2O_2 + 3NaOH \rightarrow 3RCH_2CH_2OH + Na_3BO_3 + 3H_2O This stoichiometry reflects the trivalent nature of , where one equivalent of accommodates three units, each yielding one molecule upon oxidation. Yields for the oxidation step are generally excellent, exceeding 90% for primary alcohols derived from alkenes, with overall hydroboration-oxidation efficiencies often approaching quantitative levels under optimized conditions. The method's scope is broad for unhindered alkenes, delivering stereospecific products without skeletal rearrangement, though selectivity may diminish with highly branched or internal alkenes. This transformation's historical impact lies in Brown's 1957 report, which showcased its utility for precise anti-Markovnikov hydration, earning recognition in his 1979 Nobel Prize in Chemistry.

Carbon-Carbon Bond Formations

Organoboranes derived from hydroboration serve as versatile intermediates for carbon-carbon bond formation through several established methods. One classic approach involves the of trialkylboranes with under mild conditions, typically at and , to generate acylboranes. This reaction proceeds via insertion of into the C-B bond, yielding intermediates such as R-C(O)-B(R')₂, which can be further transformed into ketones, aldehydes, or carboxylic acids upon treatment with appropriate reagents like lithium aluminum hydride or alkaline . Developed by in the 1970s, this method provides a direct route to one-carbon homologated products with high efficiency, often achieving yields exceeding 90% for simple alkyl groups, and demonstrates the utility of organoboranes in avoiding harsh conditions required by traditional organometallic reagents. The Suzuki-Miyaura cross-coupling represents a cornerstone for sp²-sp² and sp³-sp² C-C bond formations using organoboranes. In this palladium-catalyzed process, alkyl-, alkenyl-, or arylboranes react with aryl or vinyl halides in the presence of a , forming biaryl or alkyl-aryl products via and steps. For instance, alkylboronic esters or acids from hydroboration couple with aryl bromides under aqueous conditions at 80–100°C, delivering yields of 80–95% while tolerating a wide range of s such as esters and ketones. This method surpasses Grignard reagents in functional group compatibility and milder reaction conditions, enabling selective couplings without side reactions from β-hydride elimination in alkyl cases. Recent advancements post-2010 have incorporated earth-abundant metals like , where Co(II) precatalysts with ligands facilitate sp²-sp³ couplings of alkylboranes with aryl chlorides at , achieving up to 92% yield and expanding accessibility beyond precious metals. Additional strategies include the Matteson homologation for precise chain extension and conjugate additions to α,β-unsaturated carbonyls. The Matteson reaction employs chiral boronic esters treated with dichloromethyllithium or similar carbenoids, inserting a methylene unit into the C-B bond with near-perfect stereocontrol (>99% ), followed by nucleophilic to afford homologated products. This iterative , pioneered in the 1980s, is particularly valuable for synthesizing complex chains, with overall yields of 70–85% per step for multi-carbon extensions. Complementarily, organoboranes undergo 1,4-addition to enones in the presence of oxygen or catalysts, delivering β-substituted carbonyls regioselectively; for example, B-alkyl-9-borabicyclo[3.3.1]nonanes add to derivatives at 25°C to give 1,4-adducts in 85–95% yield, highlighting the anti-Markovnikov selectivity inherited from hydroboration. These transformations underscore the versatility of organoboranes in constructing diverse carbon frameworks under controlled, stereospecific conditions.

Other C-Heteroatom Bond Formations

Organoboranes derived from hydroboration can be transformed into primary amines through reactions using electrophilic nitrogen sources such as chloramine (NH₂Cl) or hydroxylamine-O-sulfonic acid (H₂NOSO₃H). In a typical procedure, trialkylborane (R₃B) reacts with chloramine to afford the corresponding primary amine (R-NH₂) after aqueous workup, retaining the anti-Markovnikov regiochemistry and syn stereochemistry of the initial hydroboration step. This method provides a convenient route to amines from alkenes, with yields typically ranging from 70-90% for simple alkyl systems. A related transformation involves azidation, leading to alkyl azides that can be reduced to primary amines. (HN₃) serves as an for direct , generating primary amines with similar stereochemical fidelity. Thioether formation from organoboranes occurs via reaction with electrophiles like dimethyl (MeSSMe), often under to promote selective B-C cleavage and generate alkyl radicals that couple with the sulfur source, yielding R-SMe products. This approach achieves good for the transfer, with yields around 70-85% for dialkylboranes. Halogenation of organoboranes provides access to alkyl halides, employing iodine (I₂) in the presence of base for iodides or N-chlorosuccinimide (NCS) for chlorides, with retention of at the carbon . These reactions deliver primary alkyl halides in 75-90% yields, enabling anti-Markovnikov halide synthesis from alkenes. Recent advances have introduced metal-catalyzed variants for forming C-O and C-P bonds from organoboranes, such as copper-catalyzed couplings that expand the scope beyond traditional electrophilic substitutions. For instance, Cu(I) systems facilitate the formation of alkyl ethers or phosphines from alkylboranes with oxygen or electrophiles, achieving high efficiency in the with enantioselective control in chiral settings.

Selective and Specialty Reagents

Dialkylboranes for

Dialkylboranes, such as (also known as bis(3-methyl-2-butyl)borane or (sia)2BH), were developed by in the early to enhance in hydroboration reactions through steric hindrance. These reagents are prepared by treating (BH3, typically as its THF or generated from ) with two equivalents of a sterically demanding , such as 2-methyl-2-butene, at low temperatures (0–5°C) to form the dialkylborane. The resulting (sia)2BH features bulky alkyl groups that limit reactivity to less hindered substrates, enabling precise control over addition direction. The primary advantage of dialkylboranes like (sia)2BH lies in their exceptional for terminal alkenes in the presence of internal or more substituted double bonds. Unlike , which reacts indiscriminately, (sia)2BH achieves monohydroboration of terminal alkenes with greater than 99% anti-Markovnikov orientation in competitive mixtures, directing exclusively to the less substituted carbon. This selectivity arises from the steric bulk of the siamyl groups, which impedes approach to hindered sites while maintaining the concerted, syn characteristic of hydroboration. The general is represented as: (\ce{sia})_2\ce{BH} + \ce{R-CH=CH2} \rightarrow \ce{R-CH2-CH2-B(sia)2} Subsequent oxidation yields the corresponding with high fidelity. In applications, dialkylboranes excel in the selective hydroboration of polyenes, such as non-conjugated dienes. For instance, treatment of 1,5-hexadiene with (sia)2BH followed by oxidation provides 5-hexen-1-ol in 93% yield, preserving the internal and avoiding over-hydroboration or cyclization observed with less selective like . This approach has been instrumental in synthesizing unsaturated alcohols from complex dienes, offering synthetic control in and precursor preparation. Despite their utility, dialkylboranes have limitations stemming from their design. The steric bulk slows reaction rates compared to 3, often requiring longer times or excess reagent, and they exhibit thermal instability, decomposing above 25°C and necessitating low-temperature handling. These factors, while enabling selectivity, restrict their use to unhindered alkenes and preclude applications with highly substituted or electron-deficient olefins. Brown's innovations with these reagents in the marked a pivotal advance in achieving steric control for regioselective hydroboration.

Cyclic and Arylboranes

Cyclic and arylboranes are specialized hydroborating agents distinguished by their enhanced thermal and air , improved in organic solvents, and superior relative to acyclic dialkylboranes. These properties arise from their constrained structures, which limit redistribution and multiple additions while facilitating precise placement. Developed in the late 1960s and 1970s through advancements by and collaborators, such reagents like 9-borabicyclo[3.3.1]nonane (9-BBN, discovered in 1968), catecholborane (HBcat, 1970), and (HBpin, 1992) have become commercially available, broadening their adoption in synthetic applications. 9-BBN, a bicyclic dialkylborane, is prepared via the hydroboration of with (typically BH₃·THF or BH₃·SMe₂) at , yielding the air-stable dimeric form (9-BBN)₂ after . This steric bulk confers exceptional for terminal alkenes, with boron adding almost exclusively to the less substituted carbon and minimal or redistribution observed, even with di- or polyenes where mono-addition reaches 95% efficiency. The resulting alkyl-9-BBN species serve as precursors to boronic acids and esters for palladium-catalyzed cross-couplings, such as the Suzuki-Miyaura reaction, and the reagent's commercial availability as a solid dimer simplifies handling. Catecholborane (HBcat), an arylborane featuring a five-membered chelate ring from , is synthesized by reacting with BH₃ or via redistribution from B₂H₆ and , producing a distillable liquid with moderate air stability. It excels in the hydroboration of alkynes, delivering syn addition to form cis-vinylboronates with high and without over-reduction. The key is: \ce{RC#CH + HB(O2C6H4) ->[HBcat] R-CH=CH-B(O2C6H4)} These vinylboronates are stable intermediates convertible to boronic acids for use in stereocontrolled syntheses via cross-coupling, and HBcat's solubility in ethers and hydrocarbons supports its versatility. Commercially available since the 1980s, it has been pivotal in post-1970 developments for functionalization. (HBpin), a six-membered cyclic boronate , is obtained by esterification of pinacol with BH₃ or through of HBBr₂ with pinacol, yielding a colorless with excellent air and moisture stability due to the rigid pinacol framework. In classic non-catalytic hydroboration-oxidation, it adds to alkenes with anti-Markovnikov , followed by H₂O₂/NaOH treatment to afford primary alcohols in yields often exceeding 90%, particularly for unhindered terminal olefins. The alkylboronate products can be hydrolyzed to boronic acids for applications in C-C bond formations, and HBpin's commercial availability as a shelf-stable has facilitated its integration into routine protocols since the early , following its first synthesis in 1992.

Modern Catalytic Hydroboration

Transition Metal Catalysts

Transition metal catalysts have significantly expanded the scope and utility of hydroboration reactions by enabling milder conditions, broader substrate compatibility, and tunable compared to the classic uncatalyzed processes using dialkylboranes. Earth-abundant metals such as , , iron, and are increasingly employed for their cost-effectiveness and , while noble metals like and provide high efficiency and stereocontrol. For instance, (I) complexes, often supported by N-heterocyclic (NHC) ligands, catalyze the hydroboration of terminal alkenes with (HBpin), achieving Markovnikov under room temperature conditions with low catalyst loadings of 1-5 mol%. Similarly, (I) catalysts with chiral phosphine ligands such as facilitate anti-Markovnikov addition to styrene derivatives, yielding products with up to 98:2 enantiomeric ratios. The general mechanism for these catalytic cycles involves of the B-H bond to the metal center, followed by migratory insertion of the unsaturated substrate into the resulting metal-hydride or metal-boryl bond, and culminating in to deliver the organoborane product. This pathway allows for inversion relative to the uncatalyzed reaction in some cases; for example, copper-catalyzed hydroboration of a proceeds via alkene insertion into a copper-hydride , favoring the branched (Markovnikov) product. \mathrm{R-CH=CH_2 + HBpin \xrightarrow{\mathrm{Cu(I)\ NHC\ (1-5\ mol\%)}\ R-CH(Bpin)-CH_3} This regioselectivity is particularly useful for accessing branched alkylboronates from aliphatic terminal alkenes, with yields often exceeding 90% even in the presence of functional groups like esters or ketones. The scope extends to functionalized alkenes, alkynes, and imines, where iron and cobalt catalysts enable selective addition to challenging substrates such as internal alkynes or electron-deficient alkenes, often at ambient temperatures. Asymmetric variants, prominent since the 2010s, leverage chiral ligands on rhodium or copper to achieve high enantioselectivities (ee >95%) for chiral boronate synthesis. Recent advances (2023-2025) highlight the role of manganese catalysts and nanoparticle systems in further enhancing efficiency for alkyne hydroboration. Ligand-free MnBr₂ catalyzes stereoselective hydroboration of terminal alkynes with HBpin, delivering (E)-vinylboronates in yields over 90% without additives. Similarly, rhodium-ruthenium nanoparticles supported on carbon nanotubes enable solvent-free hydroboration of alkynes and alkenes at room temperature, achieving near-quantitative yields (up to 99%) and recyclability over multiple runs. These developments underscore the advantages of transition metal catalysis, including low loadings, operational simplicity, and compatibility with diverse functional groups, making it a cornerstone of modern synthetic methodology.

Metal-Free and Organocatalytic Methods

Metal-free and organocatalytic hydroboration represents a sustainable approach to C-B bond formation, avoiding transition metals while enabling mild conditions and broad substrate compatibility, particularly for challenging unsaturated systems like imines, nitriles, and heterocycles. These methods leverage organic Lewis bases or acids to activate B-H bonds, aligning with principles by minimizing waste and enabling recyclable catalysts. Recent advancements post-2020 have expanded their scope to include and (FLP) mechanisms, enhancing selectivity for novel transformations. Organocatalysts such as N-heterocyclic (NHC)-boranes and phosphines facilitate hydroboration of alkenes and imines through B-H activation, often proceeding via pathways for trans-selective . For instance, NHC-boranes enable hydroboration of internal alkynes and diynes, yielding (E)-vinylboranes with yields of 70-95% and high under visible-light initiation. Phosphines similarly promote anti-Markovnikov hydroboration of terminal alkenes, achieving 80-92% yields with excellent . Choline-based ionic liquids, like choline acetate ([Chol][OAc]), serve as efficient, bio-derived organocatalysts for hydroboration of imines and nitriles at low loadings (4-6 mol%) and in THF, delivering up to 99% yields for imines while preserving sensitive functional groups like and alkenes. The reaction for imines follows the general scheme: \ce{R2C=NR' + HBpin ->[[\Chol][OAc]] R2CH-N(R')Bpin} This process exhibits high chemoselectivity and supports one-pot amine synthesis from nitriles with 90-98% efficiency. Frustrated Lewis pairs (FLPs) provide a metal-free mechanism for hydroboration-related reductions, particularly of CO₂ and heterocycles, by cooperatively activating B-H bonds through ambiphilic interactions. Intramolecular FLPs, such as those combining phosphines and boranes, reduce CO₂ to methanol or formates using HBcat or HBpin, attaining 95-99% yields with turnover numbers exceeding 2000 under ambient conditions. For heterocycles like indoles, radical-mediated metal-free hydroboration installs C-B bonds at C3 with 85-95% yields and anti-Markovnikov regioselectivity, proceeding via boryl radical addition. Dearomative hydroboration of pyridines and quinolines has advanced with organocatalytic variants, such as amine-borane systems yielding 1,4-dihydropyridines in 80-95% yields and high 1,4-selectivity at low temperatures. Asymmetric variants employ chiral organocatalysts, including FLPs derived from precursors, to achieve enantioselective hydroboration of alkenes and imines with 85-95% ee and 80-92% yields, often under solvent-free conditions using ionic liquids as media. These systems activate B-H bonds without metals, promoting diastereoselective to prochiral substrates. Post-2020 literature emphasizes B-H innovations, such as photoinduced chains with NHC-boranes or FLP-mediated transfers, opening opportunities for sustainable C-B formation in .

References

  1. [1]
    [PDF] Herbert C. Brown - Nobel Lecture
    The hydroboration reaction allows the chemist to unite to boron under excep- tionally mild conditions either three different olefins (38), or to cyclize ...
  2. [2]
    Organoboranes as a Source of Radicals | Chemical Reviews
    Brown's Nobel Prize in 1979 for the development of boron-containing compounds into important reagents in organic synthesis. 3 Many efforts have been made in ...
  3. [3]
    Reduction of CO2 to Trimethoxyboroxine with BH3 in THF
    Oct 21, 2014 · Commercially available BH3·THF contains <0.5 mol % of NaBH4 as a stabilizing reagent. It is known that BH3·THF in THF gradually decomposes ...
  4. [4]
    [PDF] Ammonia-Borane and Related N-B-H Compounds and Materials
    Apr 17, 2006 · The stability, reactivity and utility of amine-borane complexes have lead to the commercial availability of many of these compounds. The ...
  5. [5]
  6. [6]
    Borane-tetrahydrofuran complex 14044-65-6 wiki - Guidechem
    The borane-tetrahydrofuran complex was first synthesized by H. C. Brown and his colleagues in the early 1950s during their pioneering work on organoboranes.
  7. [7]
  8. [8]
    [PDF] Borane dimethyl sulfide complex - SAFETY DATA SHEET
    Sep 22, 2009 · Stability. Reacts violently with water, liberating extremely flammable gases. Moisture sensitive. Conditions to Avoid. Keep away from open ...
  9. [9]
    The Journal of Organic Chemistry - ACS Publications
    Hydroboration. 45. New, convenient preparations of representative borane reagents utilizing borane-methyl sulfide | The Journal of Organic Chemistry.
  10. [10]
    [PDF] Boron trifluoride etherate in organic synthesis - MedCrave online
    Jan 21, 2019 · Diborane can be generated by the addition of sodium borohydride to boron trifluoride etherate in tetrahydrofuran or ether at 0o-5oC. Diborane is ...
  11. [11]
    Generation of diborane from boric acid and its chromatographic ...
    Aug 8, 2025 · The method is based on the hydride generation technique. A standard sample of boric acid was well dried and mixed with about 50 mg of powdered ...Missing: derivatives hydroboration
  12. [12]
    Borane Evolution & Organic Synthesis: Phase-Vanishing Method
    Apr 13, 2021 · For reviews of hydroboration, see: (a) H. C. Brown, Tetrahedron 12 (1961) 117-138; (b) M. B. Smith, In March's Advanced... · For a review of BH3- ...
  13. [13]
    In situ generation of radical initiators using amine-borane complexes ...
    Atom transfer radical addition of alkyl halides to alkenes was developed using a low amount of a stable initiator, amine borane complexes.
  14. [14]
    CA2327658A1 - Preparation and purification of diborane
    This process involves reaction of lithium or sodium borohydride with boron trifluoride (BF3) in the absence of a solvent. ... SYNTHESIS OF BORON TRIFLUORIDE.
  15. [15]
    Organocatalytic Hydroboration of Carbonyl Compounds Promoted ...
    The organocatalytic activity of choline-based ionic liquids in the hydroboration of ketones, aldehydes, and carboxylic acids with pinacolborane was ...Missing: eco- borane
  16. [16]
    Unusual kinetics for the reaction of 9-borabicyclo[3.3.1]nonane with ...
    Rate data and rate constants for the hydroboration of cyclopentene (0.400 M) ... Again, this is observed in the competitive studies (Table 3). Solvent Effects.
  17. [17]
    Substituent effects in hydroboration: reaction pathways for the ...
    Hydroboration Reaction and Mechanism of Carboxylic Acids using NaNH2(BH3)2, a Hydroboration Reagent with Reducing Capability between NaBH4 and LiAlH4. The ...Missing: percentage | Show results with:percentage
  18. [18]
    Organic Syntheses Procedure
    The present preparation illustrates the hydroboration of a terminal olefin and the oxidation of the resultant trialkylborane. The hydroboration of an olefin ...
  19. [19]
    8.5 Hydration of Alkenes: Addition of H2O by Hydroboration
    Sep 20, 2023 · When an alkene reacts with BH3 in THF solution, rapid addition to the double bond occurs three times and a trialkylborane, R3B, is formed. For ...
  20. [20]
    Facile reaction of B-alkylboracyclanes with .alpha.,.beta.
    Facile reaction of B-alkylboracyclanes with .alpha.,.beta.-unsaturated carbonyl derivatives. Extension of the 1,4-addition reaction via organoboranes to highly ...
  21. [21]
    Cross‐Coupling Reactions Of Organoboranes: An Easy Way To ...
    May 25, 2011 · Many chemists applied such a Suzuki coupling reaction by using B-saturated alkylboron compounds. For instance, Danishefsky et al. reported a ...
  22. [22]
    Carbon-Carbon Bond Formation via Boron Mediated Transfer
    Aug 6, 2025 · With respect to organoboron chemistry, the most widely used methods (reduction and hydroboration) do not result in carbon-carbon bond formation.
  23. [23]
    Cobalt-Catalyzed C(sp2)–C(sp3) Suzuki-Miyaura Cross Coupling
    New methods that employ Earth-abundant transition metals as catalysts with organoboron coupling partners and common, readily used bases are attractive to ...
  24. [24]
    Boronic ester homologation with 99% chiral selectivity and its use in ...
    Boronic ester homologation with 99% chiral selectivity and its use in syntheses of the insect pheromones (3S,4S)-4-methyl-3-heptanol and exo-brevicomin. Click ...
  25. [25]
    Journal of the American Chemical Society
    The Reaction of Organoboranes with Chloramine and with Hydroxylamine-O-sulfonic Acid. A Convenient Synthesis of Amines from Olefins via Hydroboration. Click to ...
  26. [26]
    Reaction of organoboranes with hydrazoic acid | Organometallics
    Reaction of organoboranes with hydrazoic acid. Click to copy article ... Brown. Electrophilic Amination Routes from Alkenes. 2000, 37-63. https://doi ...
  27. [27]
    A fast reaction of organoboranes with iodine under the influence of ...
    A fast reaction of organoboranes with iodine under the influence of base. A convenient procedure for the conversion of terminal olefins into primary iodides ...Missing: NCS | Show results with:NCS
  28. [28]
    The transition metal-catalysed hydroboration reaction
    Oct 7, 2022 · The hydroboration reaction is the addition of a boron–hydrogen bond to an unsaturated organic group and was first reported by HC Brown in 1956.
  29. [29]
    Hydroboration. VIII. Bis-3-methyl-2-butylborane as a Selective ...
    Brown, K.P. Singh, Brian J. Garner. Hydroboration of terpenes I. The selective hydroboration of myrcene with disiamylborane. Journal of Organometallic ...
  30. [30]
    Journal of the American Chemical Society
    XIII. The Hydroboration of Dienes with Disiamylborane. A Convenient Procedure for the Conversion of Selected Dienes into Unsaturated Alcohols.Missing: original | Show results with:original
  31. [31]
    Hydroboration. 57. Hydroboration with 9-borabicyclo[3.3.1]nonane ...
    Facile Preparation of Anion Trapping Polymer Electrolytes by Reaction between 9-Borabicyclo[3.3.1]nonane (9-BBN) and Poly(propylene oxide). Chemistry ...Missing: original | Show results with:original
  32. [32]
  33. [33]
    Pinacolborane - Organic Chemistry Portal
    The use of 9-BBN dimer as a catalyst and pinacolborane as a turnover reagent enables an efficient hydroboration of nitriles to provide N,N-diborylamines, which ...
  34. [34]
    9-Borabicyclo[3.3.1]nonane (9-BBN) - ResearchGate
    Effect of structure on the reactivity of representative olefins toward hydroboration by 9-borabicyclo[3.3.1]nonane. Article. Aug 1976. Herbert C. Brown · Ronald ...
  35. [35]
  36. [36]
  37. [37]
    Metal‐Free Catalytic Hydroboration of Unsaturated Compounds: A ...
    Nov 24, 2022 · This review discusses the immense progress made over recent years in metal-free catalytic hydroboration.
  38. [38]
    Organocatalytic hydroboration of imines, nitriles, and amides using ...
    This study reports the application of choline-based ionic liquids as organocatalysts in the hydroboration of imines, nitriles, and amides.
  39. [39]
    Recent advances in C–B bond formation by borylation with NHC ...
    Sep 25, 2025 · This review summarises recent advances in C–B bond formation utilising B–H activation strategy of N-heterocyclic carbene borane with various ...
  40. [40]
    Advances in CO 2 activation by frustrated Lewis pairs
    Nov 9, 2023 · Frustrated Lewis pairs (FLPs) can activate small molecules, including CO 2 and convert it into value added products.
  41. [41]
    Metal-free C–H Borylation and Hydroboration of Indoles | ACS Omega
    Oct 2, 2023 · This mini-review discusses the recent progress in the area of C–H borylation and hydroboration reactions of indoles under metal-free conditions.
  42. [42]
    Recent Strategies in the Nucleophilic Dearomatization of Pyridines ...
    Jan 2, 2024 · Additionally, some metal-free and organocatalytic variants have been devised lately. Cycloaddition reactions involving pyridines, quinolines, ...
  43. [43]
    Article Readily accessible chiral frustrated Lewis pair catalysts
    Apr 18, 2024 · The chiral nature of these internal alkene precursors leads to highly selective diastereoselective hydroboration affording the efficient ...
  44. [44]
    Asymmetric catalysis by chiral FLPs: A computational mini‐review
    Apr 25, 2024 · In the realm of metal-free enantioselective organic synthesis, chiral frustrated Lewis pairs (FLPs) shine as exceptional asymmetric catalysts.