Fact-checked by Grok 2 weeks ago

Reductive amination

Reductive amination is a fundamental that forms carbon-nitrogen bonds by converting carbonyl compounds, such as aldehydes or ketones, into amines through the intermediate formation of an or , followed by using a source or with a . First reported in 1921 by Georges Mignonac using catalytic of aldehydes or ketones with , this process, also known as reductive , typically proceeds in one or two steps and is prized for its ability to avoid the overalkylation problems common in direct of amines. Reductive amination holds critical importance in pharmaceutical , where it accounts for at least a quarter of carbon-nitrogen bond-forming steps and is employed in the production of over 70 marketed drugs across therapeutic areas including agents, cardiovascular medications, and anticancer compounds. Its operational simplicity, broad substrate scope, and compatibility with chiral auxiliaries or biocatalysts make it indispensable for constructing complex amine architectures in , agrochemicals, and .

Overview

Definition and scope

Reductive amination is an essential organic transformation that converts aldehydes or ketones into amines by forming an intermediate (from primary amines) or / ion (from secondary amines), which is subsequently reduced using a . This method provides a direct and efficient route for C–N bond formation, avoiding the need for isolation of unstable intermediates in many cases. The general reaction scheme is depicted as: \ce{R2C=O + H2N-R' ->[reducing agent] R2CH-NH-R'} where \ce{R2C=O} represents an (\ce{R = H}) or , and (\ce{H2N-R'}\ ) is the , yielding a secondary product; variations allow for primary or formation by adjusting the source. The scope of reductive amination encompasses the synthesis of primary, secondary, and tertiary amines, accommodating diverse substrates such as aliphatic aldehydes and ketones, aromatic carbonyls, and even functionalized variants like α- acids or esters. This versatility stems from the reaction's tolerance for a broad array of functional groups, enabling applications in complex molecule assembly. Reductive amination holds significant importance in synthetic chemistry, particularly in the production of pharmaceuticals, where it ranks among the top ten most frequently employed reactions for active pharmaceutical ingredients and precursors, with documented use in over 70 approved drugs across therapeutic categories like agents and antivirals. Amines feature in approximately 40–60% of small-molecule pharmaceuticals, underscoring the method's role in . It is also applied in agrochemical synthesis for herbicides and pesticides, as well as in for functional polymers and ligands.

Historical development

The , introduced by Rudolf Leuckart in 1885, represented an early precursor to reductive amination by enabling the conversion of aldehydes and ketones to primary amines through reaction with or , which served as both nitrogen source and . This method laid the groundwork for subsequent developments, though it often required harsh heating and produced formamide byproducts. In 1891, Otto Wallach expanded the reaction, demonstrating its application to alicyclic and ketones and aldehydes using . The mid-20th century saw significant progress toward milder and more selective conditions for reductive amination. In 1969, Richard F. Borch and coworkers introduced (NaBH₃CN) as a that selectively targets ions at 6–8 without reducing the parent carbonyl, enabling efficient one-pot reactions of aldehydes or ketones with amines under ambient conditions. This innovation dramatically expanded the scope of reductive amination, making it a staple in for constructing complex amines while minimizing side products like over-reduced alcohols. During the 1970s and 1980s, further refinements focused on hydride reagents with enhanced stability and selectivity. Ahmed F. Abdel-Magid and colleagues developed sodium triacetoxyborohydride (NaBH(OAc)₃) in 1990, optimizing it for direct reductive amination of a broad range of carbonyls and amines, including less reactive ketones and anilines, with high yields in protic solvents. Concurrently, the 1990s brought advances in techniques, with progress in supported metal catalysts such as for selective reductive amination of carbonyls with or amines under moderate pressures. These methods improved and , reducing reliance on stoichiometric metals. The 2000s witnessed a pivotal shift from predominantly stoichiometric hydride-based processes to catalytic protocols, facilitating industrial-scale production. Reviews and studies, such as those by Armin Börner, highlighted the integration of catalysts like and with molecular for enantioselective and high-throughput reductive aminations, addressing environmental and economic concerns by minimizing waste and reagent costs. This evolution underscored reductive amination's transition to a cornerstone of sustainable synthesis in pharmaceuticals and fine chemicals.

Reaction Mechanism

Imine or iminium ion formation

The formation of an or ion constitutes the initial step in reductive amination, where a primary or secondary condenses with an or to generate a reactive C=N . This process begins with the nucleophilic attack of the nitrogen on the electrophilic carbonyl carbon, yielding a tetrahedral carbinolamine after proton transfer. The carbinolamine formation is reversible and typically occurs under mildly acidic conditions to balance amine nucleophilicity and facilitate subsequent steps. Dehydration of the carbinolamine then produces the from primary amines or the ion from secondary amines. For primary amines, protonates the carbinolamine hydroxyl group, enabling water departure and formation of the neutral , also known as a : \ce{R^2C=O + H2NR' -> R^2C=NR' + H2O} This step is equilibrium-controlled, with the position favoring reactants unless water is removed. Techniques such as a Dean-Stark trap, employing with , effectively shift the equilibrium toward the by sequestering water. Optimal conditions involve a of approximately 4–5, as excessive acidity protonates the (reducing its nucleophilicity), while neutrality hinders . With secondary amines, the absence of a second hydrogen on nitrogen prevents neutral imine formation; instead, dehydration yields a protonated iminium ion: \ce{R^2C=O + HNR_2' ->[H+] R^2C=NR_2'^+ + H2O} Under certain conditions, particularly in the presence of alpha-hydrogens on the carbonyl substrate and basic media, tautomerization can lead to enamine formation: \ce{R2CH-C(O)R'' + HNR2' -> R2C=CR''-NR2' + H2O} However, in typical acid-catalyzed reductive amination, the iminium ion predominates as the key electrophilic species. Imines and iminium ions exhibit stereochemical features, with imines capable of E/Z isomerism due to the partial double-bond character of the C=N linkage, influenced by steric hindrance and electronic effects between substituents.

Reduction of intermediates

The reduction step in reductive amination entails the stereoselective transfer of a or molecular to the electrophilic carbon of the C=N bond in the or intermediate, resulting in the formation of the desired product. This process is typically facilitated by reducing agents that deliver the hydride to the carbon while the accepts a proton, yielding a stable tetrahedral . The general reaction can be represented as: \ce{R2C=NR' + [H]- -> R2CH-NHR'} This transformation is highly selective, with appropriate reducing agents preventing over-reduction of the C=N bond or subsequent hydrogenolysis to alkanes (hydrocarbons), which might occur under forcing conditions with non-selective reductants like lithium aluminum hydride. The kinetics of the reduction are influenced by the stability and protonation state of the intermediate; imines are generally less reactive than their protonated iminium counterparts, with acidic conditions (pH around 6-7) promoting protonation to enhance the electrophilicity of the C=N bond and accelerate hydride transfer. Computational studies indicate activation free energies for imine reduction ranging from 6.9 to 11.8 kcal/mol, depending on substrate electronics and coordination effects, making this step thermodynamically and kinetically favorable over competing reductions. Agents such as sodium cyanoborohydride exhibit high selectivity for iminium ions at mildly acidic pH, minimizing interference from unreacted carbonyls. Common side reactions during this phase include imine back to the starting carbonyl and , driven by trace , or imine dimerization via , particularly with reactive aldehyde-derived imines. These issues are mitigated by employing solvents and conditions, such as dry or molecular sieves, to suppress and stabilize the intermediate for efficient .

Direct versus indirect pathways

Reductive amination can proceed via direct or indirect pathways, each offering distinct approaches to synthesizing from carbonyl compounds and . In the direct pathway, the carbonyl compound reacts with the and in a single reaction vessel, facilitating formation of the or intermediate followed by immediate reduction to the product. This one-pot process is represented conceptually as the combination of an or , , and selective yielding the secondary or tertiary . The direct method enhances laboratory efficiency by minimizing handling and purification steps, making it particularly suitable for routine synthetic applications. The indirect pathway, in contrast, involves a stepwise sequence where the or intermediate is first formed and isolated before undergoing separate . This approach is advantageous when the intermediate is unstable under one-pot conditions or requires purification to remove impurities that could interfere with . For instance, indirect methods are often employed for ketones that form imines sluggishly or in low yields during direct attempts, allowing optimization of condensation conditions independently. However, the isolation step can lead to material losses and lower overall yields due to handling and potential decomposition. A key advantage of the direct pathway is its operational simplicity and reduced waste, as demonstrated with selective hydride reagents like (NaBH₃CN), which operates effectively at mildly acidic pH (6–8) to preferentially reduce ions without significantly affecting the carbonyl . For example, the direct reductive amination of aldehydes such as with primary amines using NaBH₃CN affords secondary amines in high yields (often >80%) under mild conditions, avoiding over-reduction to alcohols. Sodium triacetoxyborohydride (NaBH(OAc)₃) further improves selectivity in direct processes, particularly for ketones and acid-sensitive substrates, by exhibiting even lower reactivity toward carbonyls in solvents like . Despite these benefits, direct methods risk side reactions such as overalkylation of primary amines or incomplete if the reducing agent lacks sufficient selectivity. Indirect pathways mitigate some direct method limitations by enabling intermediate characterization and purification, which is crucial for complex molecules or when enamine formation predominates with . A representative example involves preforming the from a like and a primary under dehydrating conditions, followed by reduction with NaBH₄ or catalytic , achieving yields comparable to direct routes but with greater control over or in sensitive cases. Nonetheless, the multi-step nature increases time and resource demands, potentially resulting in lower overall efficiency compared to optimized direct procedures. The choice between pathways depends on compatibility; direct is preferred for aldehydes due to rapid formation, while indirect suits prone to side products in one-pot settings. Certain reducing agents, such as NaBH₄ activated by additives, are tailored for direct use but can also support indirect reductions.

Reducing Agents

Hydride-based reagents

Hydride-based reagents serve as stoichiometric sources of hydride for the step in reductive amination, offering mild conditions and compatibility with various functional groups, though their selectivity varies depending on the specific employed. (NaBH₄) is a widely used general for both indirect and direct reductive amination, particularly effective in protic solvents such as or , where it reduces preformed imines or in situ-generated intermediates to amines. However, its lack of selectivity can lead to over- of the starting carbonyl compounds to alcohols, making it more suitable for aldehydes than ketones unless modified with additives like carboxylic acids or metal salts to enhance . A seminal of its utility involved the of Schiff bases derived from aromatic aldehydes and amines, yielding secondary amines in high yields under mild conditions. Sodium cyanoborohydride (NaBH₃CN) provides greater selectivity for ions over carbonyls, enabling efficient one-pot reductive amination at mildly acidic values of 6-8, typically in or aqueous mixtures. This acid stability allows the reagent to tolerate the protonation of without reducing unreacted aldehydes or ketones, as introduced in the Borch reduction protocol, which has become a standard method for synthesizing secondary and tertiary from diverse carbonyl-amine combinations. For example, the reaction proceeds by adding NaBH₃CN portionwise to a solution of the carbonyl compound and , maintaining control to optimize formation and . Sodium triacetoxyborohydride (NaBH(OAc)₃) offers even higher selectivity for imines in non-aqueous solvents like or , minimizing side reactions with carbonyls and enabling clean reductive amination of both aldehydes and ketones, especially with primary and secondary . Introduced as a versatile , it performs well at , often with acetic acid as a promoter for substrates, and is particularly valuable for acid-sensitive functional groups such as acetals or compounds. A typical procedure involves mixing the carbonyl, , and 1.5-2 equivalents of NaBH(OAc)₃ in dichloroethane, stirring for several hours to afford the product in high yield: \ce{R^2C=O + H2NR' ->[NaBH(OAc)3, DCE] R^2CH-NHR'} This one-pot process exemplifies its efficiency, with yields often exceeding 80% for aliphatic and cyclic substrates. Regarding practical properties, NaBH₄ exhibits high solubility in protic solvents like water and alcohols but decomposes rapidly in strong acids, rendering it inexpensive and relatively non-toxic, though it requires careful handling due to its reactivity with water. In contrast, NaBH₃CN is soluble in water, methanol, and dimethylformamide, stable at pH 3-8, but poses significant toxicity risks from its cyanide content, potentially releasing hydrogen cyanide during workup, and is notably more costly than NaBH₄. NaBH(OAc)₃, soluble in aprotic solvents and acetic acid but insoluble in water, is milder and free of cyanide hazards, making it safer than NaBH₃CN, though it is water-sensitive, flammable, and similarly expensive or higher in cost compared to NaBH₄. These attributes position hydride-based reagents as complementary to catalytic hydrogenation methods for achieving high selectivity in reductive amination.

Catalytic hydrogenation methods

Catalytic hydrogenation methods employ molecular hydrogen (H₂) in the presence of metal catalysts to reduce or ion intermediates formed during reductive amination, enabling the synthesis of primary, secondary, or tertiary amines from carbonyl compounds and amines. These approaches are divided into heterogeneous and homogeneous systems, with often utilizing supported metals like (Pd/C) or (Pt) for broad applicability, while homogeneous systems leverage soluble complexes of (Rh) or (Ru) for enhanced selectivity, particularly in asymmetric variants. Heterogeneous catalysis with Pd/C is a widely adopted method for the hydrogen gas of , typically conducted under mild pressures of 1-5 H₂ in protic solvents such as or , at temperatures ranging from to 80°C. For instance, the reaction of aldehydes with primary amines over 5-10% Pd/C yields secondary amines with high efficiency, as the catalyst facilitates both imine formation and selective while minimizing over-. Pt-based catalysts, such as Pt/C or unsupported nanoporous Pt, operate under similar conditions (10-50 bar H₂, 50-120°C), offering high activity for chemoselective of derived from aromatic aldehydes, though they are more prone to over- of sensitive functional groups compared to Pd systems. These heterogeneous methods are particularly valued in industrial settings for their recyclability and robustness. Homogeneous catalysis using Rh or Ru complexes excels in asymmetric reductive amination, where chiral ligands enable enantioselective formation of amines from prochiral ketones or aldehydes. The general process follows the equation RCHO + RNH₂ + H₂ → RCH₂NHR, catalyzed by complexes such as [Rh(cod)((R,R)-Et-DuPHOS)]BF₄ or Ru arene diphosphine systems under 1-10 atm H₂ at to 80°C in solvents like or , achieving enantiomeric excesses up to 99% for structurally diverse amines. These systems provide precise control over , making them essential for pharmaceutical synthesis, though they require careful ligand design to avoid catalyst decomposition. The advantages of catalytic hydrogenation methods include scalability for large-scale production, atom efficiency due to the use of as a clean reductant, and compatibility with a wide range of substrates, contrasting with hydride-based reagents that may suffer from selectivity issues in complex molecules. However, drawbacks involve the need for specialized equipment to handle pressurized safely, potential by impurities, and higher costs for precious metals like or . Historically, these methods trace back to the early 1930s, with and Adkins demonstrating nickel-catalyzed N-alkylation of under conditions, laying the groundwork for industrial production.

Selective and mild reducers

In reductive amination, selective and mild reducing agents are essential for handling sensitive substrates, such as those with orthogonal functional groups or requiring low-temperature conditions to prevent side reactions like over-reduction or epimerization. These agents preferentially reduce or intermediates while leaving carbonyl groups intact, enabling one-pot processes under ambient or near-ambient conditions. Palladium hydride species, generated from gas and catalysts like /C or Pd(OH)₂ supported on , provide a selective approach for direct reductive amination, particularly for sterically hindered . The mechanism involves the formation of Pd-H intermediates that attack the nitrogen or associated , facilitating hydrogenolysis with high (>97%) toward the product over carbonyl reduction, even at 30°C and 1.5 H₂ in . This method excels with challenging substrates like diisopropyl and , yielding 73% of the hindered with minimal byproduct formation from competing pathways. Silane-based reducers, such as (PMHS) in combination with (IV) isopropoxide [Ti(OiPr)₄], offer metal-catalyzed, mild alternatives that avoid hydrogen gas and operate under neutral, solvent-tolerant conditions. The Ti catalyst activates the to deliver selectively to the , achieving high in one-pot aminations of aldehydes and ketones with primary or secondary amines at , without reducing unreacted carbonyls or sensitive groups like olefins. This system has been applied to diverse substrates, including aromatic aldehydes with anilines, yielding amines in 80-95% isolated yields. Pyridine-borane complexes like 2-picoline-borane (pic-BH₃) enable biocompatible reductive aminations in aqueous or protic media, ideal for biomolecules or water-sensitive syntheses. Pic-BH₃ selectively reduces ions at 6-8, exhibiting low reactivity toward aldehydes and ketones (k_rel < 0.1 relative to imines), thus supporting efficient one-pot reactions with yields up to 99% for aliphatic and aromatic substrates without hydrolysis side products. Its stability in water makes it suitable for enzymatic or carbohydrate conjugations, such as labeling reducing sugars with amines under mild conditions (25°C, MeOH/H₂O).
Reducing AgentSelectivity (Imine vs. )Example Yield/Selectivity (%)ConditionsSource
Pd-H (in situ from H₂/Pd(OH)₂/g-C₃N₄)>97% toward ; negligible 73% yield, 97% selectivity for diisopropylbutylamine30°C, 1.5 MPa H₂, MeOHPMC9320161
PMHS/Ti(OiPr)₄High ; no olefin or reduction80-95% yield for N-benzyl anilinesRT, THF or neat10.1055/s-2000-7922
Pic-BH₃k_imine / k_carbonyl >10; stable in 95-99% yield for cyclohexylmethylamine25°C, MeOH/, 7S0040402004009135

Synthetic Design and Optimization

Substrate selection and compatibility

In reductive amination, the choice of carbonyl substrate is critical, with aldehydes generally preferred over ketones due to their higher reactivity in or formation, stemming from lower steric hindrance and faster . Aldehydes typically react within hours under mild conditions, achieving high yields (e.g., 80-95%), whereas ketones require longer reaction times and often benefit from to enhance selectivity. However, ketones bearing alpha-hydrogens can lead to side reactions via tautomerization to enamines, particularly in direct pathways with primary amines; this issue is more pronounced in aliphatic ketones and can be minimized by using selective reducing agents or indirect isolation strategies. The selection of amine substrates depends on the desired product: primary amines yield secondary amines, secondary amines produce amines, and (often as ) generates primary amines, though the latter requires excess to suppress overalkylation. Primary aliphatic amines are highly compatible, but aromatic amines like anilines exhibit slower reactivity due to lower nucleophilicity, necessitating optimized conditions such as molecular sieves for removal. Secondary amines avoid formation altogether, proceeding via ions, which broadens their utility for amine synthesis without complications. Functional group compatibility is a key advantage of reductive amination, particularly with mild reagents like sodium triacetoxyborohydride (NaBH(OAc)₃), which tolerate halides, esters, ethers, and even groups without significant interference, enabling reactions on complex molecules. For instance, aromatic aldehydes paired with anilines proceed smoothly to diarylmethylamines in 70-90% yields, preserving aryl halides. Aliphatic ketones with alkylamines, such as and , afford tertiary amines efficiently (e.g., 85% yield), though additional carbonyls in the may risk overreduction with non-selective agents like NaBH₄; selective reducers like NaBH₃CN or NaBH(OAc)₃ mitigate this by preferentially targeting intermediates at pH 6-8. groups remain intact under these conditions but can be reduced to amines when using catalytic , requiring careful choice for tolerance.

Reaction conditions and catalysts

Reductive amination reactions are typically performed using protic solvents such as when employing hydride-based reducing agents like , as these conditions promote efficient reduction at ambient pressures. In contrast, aprotic solvents including or are favored for reagents like sodium triacetoxyborohydride to enhance and prevent in the presence of protic . Temperatures are generally maintained between 0 and 25°C to balance reaction rates with selectivity, minimizing over-reduction or side reactions involving sensitive functional groups. The reaction is controlled in the range of 4 to 7 to facilitate or formation while preserving the reducing agent's efficacy, often achieved through the addition of acetic acid. To counteract water formation during condensation and drive equilibrium toward the intermediate, drying agents such as 4 Å molecular sieves are commonly incorporated. Acetic acid serves as a mild acid catalyst to accelerate imine formation in stoichiometric hydride reductions, typically at 0.1 to 1 equivalent relative to the carbonyl . For catalytic variants, heterogeneous metal catalysts like or platinum oxide are employed under gas (1-50 atm), often in or at to 80°C. Yield optimization frequently involves adjusting the of the to 1.1 to 2 equivalents based on the , with higher amounts (up to 2 equivalents) beneficial for less reactive ketones to ensure complete conversion without excess waste.

Scale-up and practical considerations

Scaling up reductive amination from to levels presents significant challenges, particularly in -based processes where and become limiting factors. In batch reactors, exothermic reactions can lead to hotspots and inefficient gas-liquid mixing, exacerbating safety risks and reducing yields at larger volumes. Continuous reactors address these issues by enhancing rates (e.g., kLa ~0.1/s in pipes-in-series designs) and maintaining isothermal conditions, enabling scalable without the need for oversized cooling systems. For instance, an iridium-catalyzed reductive amination was successfully scaled from 48 mL to 360 L using such systems, achieving steady-state operation and minimizing gas accumulation. Safety considerations are paramount during scale-up, especially with gas, which is highly flammable and poses risks in batch setups due to potential accumulation indoors. Flow processes mitigate this by operating outdoors with continuous venting (e.g., ~20 g/h H2) and vapor-liquid separators that keep levels below 0.3%, reducing the need for high-pressure infrastructure. For hydride-based methods, (NaBH3CN) introduces toxicity risks, requiring rigorous handling protocols to prevent free release, which can occur during or quenching; industrial avoidance of NaBH3CN often favors alternatives like picoline borane for its lower toxicity and stability in protic solvents. Purification at scale emphasizes efficient, chromatography-free methods to minimize costs and waste. Inline integrated with reactors allows separation of products from byproducts, followed by via acidification or antisolvent addition, achieving high purity (e.g., 95% with <10 ppm catalyst residue after distillation and ethanol-water ). In the continuous synthesis of safinamide mesylate, and steps yielded 83% overall without intermediate isolation, demonstrating scalability at 22 g/h production rates. Batch workups similarly employ sodium carbonate washes and distillation to prepare for , avoiding column chromatography for economic viability. Economic analysis guides reagent and process selection, with reducing agent costs often dominating at industrial scales. Hydrogenation with catalysts like Ir or Pd offers long-term savings over stoichiometric hydrides, as borohydride production is labor-intensive and waste-generating; for example, switching to catalytic hydrogenation reduced reducing agent costs by at least 50% compared to sodium triacetoxyborohydride (STAB) in a GMP-scale process, yielding a positive net present value over 10 years. In biocatalytic variants, enzyme costs comprise over 96% of expenses, but activity enhancements can lower unit prices to $0.5–0.6/g for chiral amines, making them competitive for high-value pharmaceuticals. Overall, flow-enabled reductive amination balances capital investment with operational efficiency, prioritizing modular designs for flexible production.

Variations and Extensions

Catalytic reductive amination

Catalytic reductive amination represents a powerful subclass of reductive amination reactions, leveraging transition-metal catalysts to facilitate the formation of C–N bonds under mild conditions, often avoiding stoichiometric reducing agents. These methods typically involve the in situ generation and reduction of imines or iminium ions from carbonyl compounds and amines, with catalysts such as ruthenium (Ru), iridium (Ir), and more recently nickel (Ni) enabling high efficiency and selectivity. Transition-metal catalysis enhances atom economy and sustainability by utilizing hydrogen gas, alcohols, or silanes as hydrogen sources, distinguishing these approaches from traditional hydride-based reductions. A prominent strategy within catalytic reductive amination is borrowing hydrogen (BH), where Ru or Ir catalysts employ an alcohol as both the alkylating agent and hydrogen source, allowing for dehydrogenation to form an intermediate carbonyl that reacts with the amine, followed by hydride transfer to reduce the resulting imine. In this process, the catalyst temporarily "borrows" hydrogen from the alcohol, enabling a redox-neutral cycle that regenerates the alcohol-derived carbonyl as a byproduct. For example, the reaction of an aldehyde (RCHO) with an amine (RNH₂) in the presence of a primary alcohol (R'CH₂OH) proceeds as follows: \text{RCHO} + \text{RNH}_2 + \text{R'CH}_2\text{OH} \rightarrow \text{RCH}_2\text{NHR} + \text{R'CHO} Ru-MACHO complexes have been particularly effective for such BH-mediated aminations, achieving high yields with alcohols as hydrogen donors. Iridium catalysts similarly excel in BH processes, offering broad substrate scope including benzylic and allylic alcohols, with turnover numbers often exceeding 1000. These catalysts operate at low loadings of 0.1–1 mol%, minimizing metal use while maintaining high activity under neutral conditions. Asymmetric variants of catalytic reductive amination employ chiral ligands to induce enantioselectivity, enabling the synthesis of enantioenriched amines critical for pharmaceuticals and fine chemicals. Noyori-type Ru catalysts, featuring diamine and diphosphine ligands such as (R,R)-1,2-diphenylethylenediamine (DPEN) and BINAP, achieve exceptional enantiomeric excesses (>95% ee) in the transfer hydrogenation of imines derived from ketones or aldehydes. These bifunctional catalysts facilitate outer-sphere hydride delivery, with the chiral environment controlling stereochemistry through hydrogen bonding interactions. For instance, reductions of cyclic imines or α-branched ketone-derived imines routinely deliver products in 96–99% ee using 0.1–1 mol% catalyst loading. Recent advancements (post-2020) have expanded catalytic reductive amination to more challenging feedstocks and tandem processes. A 2025 report describes a -catalyzed reductive amination of carboxylic acids with amines, using phenylsilane as the reductant and bulky trialkylphosphine ligands to form intermediates directly, bypassing acid activation steps; this method operates at low loadings (1–5 %) and tolerates diverse functional groups. In 2024, a multi-component tandem reductive amination– was developed using a cationic complex ([Ir(COD)₂]BArF) at 0.1 % loading, enabling one-pot conversion of esters to α-branched tertiary amines via hydrosilylation followed by , with scalability to gram quantities and compatibility with triglyceride-derived esters. These innovations highlight the versatility of catalytic approaches, offering low catalyst loadings (0.1–1 %) and enhanced selectivity for complex synthesis.

Reductive amination from carboxylic derivatives

Reductive amination from carboxylic derivatives involves the direct conversion of carboxylic acids, esters, and related compounds into amines, bypassing traditional carbonyl intermediates and offering advantages in availability and step economy. These methods typically proceed through activation of the derivative to form an or species, followed by selective reduction to the product. This approach is particularly valuable for utilizing abundant carboxylic feedstocks in sustainable synthesis. For carboxylic acids, a practical protocol employs zinc acetate [Zn(OAc)₂] as a catalyst with phenylsilane (PhSiH₃) as the reductant to achieve reductive amination with primary or secondary amines, yielding N-alkylated amines in good to excellent yields (up to 99%). The reaction is performed in toluene at 80 °C, tolerating various functional groups such as halides, ethers, and alkenes, and avoids over-reduction or side products common in multi-step sequences. A related nickel(II)-catalyzed variant uses Ni(OTf)₂ with bulky trialkylphosphine ligands and PhSiH₃, enabling efficient C-N bond formation under mild conditions, with broad substrate scope including aromatic and aliphatic acids. Although hydrogen (H₂) can be used in some heterogeneous systems, such as Ru-W catalysts for primary amines, the silane-based methods with Ni or Zn provide versatility for secondary amine synthesis without requiring high-pressure equipment. The overall transformation from carboxylic acids can be represented as: \ce{RCO2H + R'NH2 + 3 PhSiH3 -> RCH2NHR' + (PhSiH2O)3 + H2O} This two-step process first forms the intermediate via silane-mediated , followed by . From esters, a direct reductive N-alkylation method utilizes tert-butylmagnesium chloride (t-BuMgCl) to activate methyl or ethyl , coupled with sodium bis(2-methoxyethoxy)aluminum hydride (Red-Al) as the reductant, affording secondary amines in moderate to high yields (52-92%). The proceeds at in THF, with t-BuMgCl facilitating nucleophilic attack and , and is compatible with aryl and alkyl esters bearing electron-withdrawing groups. Silane-based of esters also enable reductive amination, often via or catalysts like B(C₆F₅)₃, where (PMHS) reduces the intermediate or formed after ester aminolysis, achieving yields up to 95% for aliphatic esters. The for these processes generally involves initial of the carboxylic derivative to an acyl ion (R-C≡N⁺R'), either directly or via an intermediate, followed by transfer from the reductant to reduce the C=N bond while eliminating the carbonyl oxygen as water or . This pathway avoids indirect routes like , which require harsh conditions and generate byproducts. In the Zn/PhSiH₃ system, the is silylated to form an O-silyl , which fragments to the acyl before reduction. These methods find applications in upgrading bio-based feedstocks, such as the reductive amination of (derived from ) with amines to produce N-substituted pyrrolidones, valuable as s and intermediates in pharmaceuticals. For instance, using or catalysts under pressure, levulinic acid reacts with to give N-phenylpyrrolidone in 90% yield, demonstrating scalability for renewable amine production.

Electrochemical and photochemical approaches

Electrochemical reductive amination represents a sustainable alternative to traditional methods, utilizing to drive the reduction of or intermediates formed from carbonyl compounds and amines. In this process, the facilitates the single-electron reduction of the iminium , generating a carbon-centered that is subsequently protonated, often using the or an additive as the source, while hydrogen evolution may occur at the depending on the cell configuration. This metal-free approach operates under mild conditions, such as room temperature in undivided cells, and avoids the need for stoichiometric chemical reductants, thereby enhancing and reducing waste. For instance, aldehydes react with amines in DMSO , where DMSO serves dual roles as solvent and hydrogen donor, yielding secondary amines with high efficiency and enabling deuterium labeling when using DMSO-d6. A key application of electrochemical methods is , where carbonyl compounds undergo reductive amination to form C-N bonds without external reducing agents, as highlighted in recent reviews on electrochemically driven . This strategy is particularly effective for synthesizing complex amines from biomass-derived carbonyls, such as , using as the ultimate source in a process. Advantages include precise control over reaction parameters via continuous flow electrocells, which improve , , and by minimizing the handling of hazardous gases or , making it suitable for applications. Photochemical reductive amination leverages visible to activate photocatalysts, enabling the of imines under ambient conditions without harsh reductants. Typically, a ruthenium-based photocatalyst such as Ru(bpy)32+ absorbs visible to reach an , which undergoes by a hydrogen donor, facilitating single-electron transfer to the imine or iminium intermediate. This generates a that protonates to form the product, often with additives like or Hantzsch esters serving as the H-source to regenerate the catalyst. The general reaction can be represented as: h\nu + \ce{R2C=NR'} + \ce{H-source} \rightarrow \ce{R2CH-NHR'} This method is highly selective for aromatic aldehydes, producing benzylic amines with good functional group tolerance and operational simplicity. Recent advances include the integration of photochemical strategies with nitro compound reduction for amine synthesis, where nitroarenes or nitroalkanes act as amine precursors in one-pot reductive amination processes. Although primarily catalytic, these developments align with photo- and electro-driven sustainability goals by enabling efficient transformation of nitro groups to amines under mild pressures and temperatures, with turnover numbers up to 3,800 using non-noble metal catalysts. Such approaches underscore the potential for light- or electricity-mediated methods to access diverse amine structures relevant to pharmaceuticals and heterocycles.

Applications

In organic synthesis

Reductive amination serves as a cornerstone in organic synthesis for constructing carbon-nitrogen bonds in complex molecules, particularly in the assembly of alkaloid frameworks where precise control over amine installation is essential. In alkaloid synthesis, it enables the transformation of ketones or aldehydes into secondary or tertiary amines under mild conditions, avoiding harsh reagents that might disrupt sensitive functionalities. A prominent strategy involves the reductive amination of cyclic ketones to form tropane alkaloids; for instance, the one-step global reduction/reductive amination of functionalized tropinone derivatives provides access to psychoplastogenic tropanes like 3α-(benzo[1,3]dioxol-5-yl)-8-azabicyclo[3.2.1]octane in high yields, streamlining the synthesis of unfunctionalized C6/C7 positions. This approach has been applied in enantiopure syntheses from chiral terpenes, achieving tropane cores via one-pot reductive amination with ammonia or primary amines. Representative examples highlight its utility in pharmaceutical precursor synthesis. Reductive amination of L-phenylacetylcarbinol (L-PAC) and its analogs with yields and stereoisomers, with modifications like 4-methyl or 4-fluoro substituents producing viable analogs in moderate conversions. In multi-step total syntheses, reductive amination has been pivotal, such as in the 1980 formal synthesis of by , where it facilitated amine installation after dihydrofuran ring construction, contributing to one of the most efficient routes to morphinan alkaloids at the time. As an alternative to traditional peptide coupling, reductive amination of aldehydes with primary amines on solid phase generates diverse tertiary amides, offering stereocontrol and compatibility with complex sequences without . Tandem reductive amination-hydrogenolysis sequences further enhance efficiency by combining C-N bond formation with deprotection. For example, in synthesis, hydrogenolytic deprotection of a terminal N-benzyl under reductive conditions triggers double reductive amination, yielding diastereomeric N-alkylated products in a 4:1 ratio without isolation of intermediates. However, limitations arise with steric hindrance, particularly in forming from bulky ketones or secondary , as / formation is disfavored, often requiring harsher conditions or alternative catalysts to achieve viable yields. These strategies underscore reductive amination's versatility in academic synthesis, with applications extending to pharmaceutical intermediates as detailed in specialized contexts.

Biochemical and biological roles

Reductive amination is a fundamental biochemical process in living organisms, enabling the synthesis and interconversion of through the stereoselective addition of to carbonyl compounds. This occurs primarily via enzymatic mechanisms that ensure high efficiency and specificity under physiological conditions. Key enzymes include pyridoxal 5'-phosphate ()-dependent transaminases, which catalyze the transfer of amino groups between and keto acids, and NAD(P)H-dependent dehydrogenases, which directly incorporate into α-keto acids. These enzymes play essential roles in , metabolic , and the production of biomolecules critical for cellular function. A prominent example is (GDH), which catalyzes the reversible reductive amination of α-ketoglutarate with and NADPH to form L-glutamate, serving as the entry point for into amino acid metabolism. This reaction is vital for , as glutamate acts as a nitrogen donor in subsequent reactions to produce other non-essential amino acids. In the fungal α-aminoadipate pathway for , reductive amination is integral, with PLP-dependent transaminases converting α-ketoadipate (derived from α-ketoglutarate) to α-aminoadipate using glutamate as the amino donor, highlighting the pathway's reliance on these enzymes for chiral amine formation. The general enzymatic reductive amination can be depicted as: \ce{R-C(=O)-R' + NH3 + NADPH -> R-CH(NH2)-R' + NADP+} This equation underscores the hydride transfer from NADPH that reduces the intermediate imine, ensuring stereoselectivity. In neurotransmitter synthesis, reductive amination supports the production of glutamate, the primary excitatory neurotransmitter in the central nervous system, through GDH-mediated incorporation of ammonia into α-ketoglutarate in astrocytes and neurons. This process is tightly regulated within the glutamate-glutamine cycle, where astrocytes synthesize glutamine from glutamate, which is then transported to neurons for reconversion to glutamate via glutaminase, with GDH facilitating direct reductive amination to replenish pools during high demand. Dysregulation of this pathway can lead to excitotoxicity in neurological disorders. The stereoselective nature of these enzymatic reactions inspires drug design strategies that mimic transaminase and dehydrogenase active sites to achieve asymmetric amination in pharmaceutical synthesis, enhancing selectivity for chiral amine therapeutics.

Industrial and pharmaceutical uses

Reductive amination serves as a cornerstone in , accounting for at least 25% of carbon-nitrogen bond-forming reactions in the industry due to its efficiency in constructing chiral amines essential for drug scaffolds. In the of antidepressants, it enables the formation of key amine linkages; for instance, (R)-, the active enantiomer in the Prozac, is prepared through a catalytic asymmetric one-pot process involving formation, borylation, transimination, and , achieving 45% overall yield and 96% enantiomeric excess from simple aldehydes. Similarly, for antihistamines, reductive amination is employed in the final deprotection and steps of loratadine analogues, such as 2-(piperidin-3-yl)-1H-benzimidazole , using sodium triacetoxyborohydride to yield products with 40–90% efficiency and improved H1 receptor selectivity. In industrial applications, reductive amination facilitates the large-scale production of amines used in and related fine chemicals, with biobased variants derived from pulp monosaccharides via selective C–N bond formation to create nonionic exhibiting low critical concentrations and high . For commodity amines like ethanolamines, which exceed 10,000 tons in annual global production and serve as precursors for and dyes, alternative routes involving reductive amination of monoethanolamine under catalytic conditions offer sustainable pathways, achieving 20–80% conversion with nickel-based catalysts. Notable case studies highlight scale-up successes; utilized an engineered reductase for the asymmetric reductive amination in producing a precursor to the abrocitinib, enabling over 3.5 metric tons at >200 kg scale with 77% conversion and >99% enantiomeric excess through a 206-fold activity enhancement. Likewise, Codexis developed an evolved ω-transaminase (ATA-117-Rd11) for the reductive amination step in sitagliptin synthesis, the DPP-4 inhibitor in Januvia for type-2 treatment, achieving high and enabling commercial production with reduced loading and improved specificity. Economically, one-pot reductive amination processes in pharmaceuticals deliver significant cost savings by minimizing intermediate isolations and purification steps, boosting yields while lowering operational expenses and generation compared to multi-step sequences.

Green Chemistry and Sustainability

Eco-friendly reducing agents

Traditional reducing agents in reductive amination, such as and sodium triacetoxyborohydride, often pose environmental concerns due to their and generation of inorganic . Eco-friendly alternatives focus on recyclable catalysts, benign donors, and biological systems to minimize and enhance . Hydrogen gas (H₂) paired with supported catalysts represents a clean reducing system, where heterogeneous supports enable catalyst recovery and reuse. For instance, polymer-bound palladium catalysts, such as triphenylphosphine-palladium acetate immobilized on polystyrene, facilitate indirect reductive amination of aldehydes with high efficiency and recyclability up to five cycles without significant loss in activity. Similarly, palladium nanoparticles embedded in metal-organic framework/polymer composites achieve selective amination of biomass-derived hydroxymethylfurfural (HMF) with yields exceeding 90%, leveraging the support's high surface area for low metal leaching. These systems reduce metal contamination in products, aligning with green chemistry principles. Formic acid serves as a safe, donor in transfer hydrogenation-based reductive amination, decomposing to CO₂ and H₂ . catalysts enable one-pot reductive amination of carbonyls and amines using , yielding secondary amines in up to 99% selectivity under mild conditions (80–120°C), with the byproduct CO₂ posing no disposal issues. Mesoporous graphitic carbon nitride-supported AgPd alloys further enhance this approach for coupling aldehydes and nitroarenes, achieving 95% yields and facile catalyst separation via . This avoids gaseous H₂ handling, improving and . Silanes like (PMHS) and alcohols such as isopropanol enable with low waste profiles when combined with catalysts. PMHS, a industry byproduct, acts as a mild reductant in tin- or titanium-catalyzed reductive amination of carbonyls to amines, with as co-solvent promoting over 90% and generating only polymers as benign waste. For isopropanol-mediated processes, complexes like [Ru(CO)₂(Ph₄C₄CO)]₂ catalyze imine reduction in reductive amination, delivering functionalized amines in 80–95% yields under or conventional heating, with the alcohol solvent doubling as hydrogen donor to minimize auxiliary inputs. These protocols exhibit low environmental factors (E-factors) by avoiding stoichiometric reductants. Biocatalytic reductive amination employs engineered enzymes, such as reductases (IREDs) and amine dehydrogenases (AmDHs), for stereoselective in aqueous media. Protein of IREDs expands scope to bulky ketones and aldehydes, achieving enantiomeric excesses >99% and turnover numbers up to 10,000 with NAD(P)H systems, thus operating under ambient conditions without organic solvents. AmDH variants, evolved from native dehydrogenases, catalyze direct of keto acids to chiral , as demonstrated in the commercial production of intermediates with >200-fold activity improvement over wild-type enzymes. These biological methods yield E-factors below 5 kg waste/kg product, significantly lower than chemical counterparts, by integrating cofactor regeneration and avoiding metal residues.

Solvent-free and continuous processes

Solvent-free reductive amination reactions have gained attention for their environmental benefits, particularly through mechanochemical techniques like ball milling, which enable reactions without organic solvents. In these processes, carbonyl compounds react with amines to form imines , followed by reduction using agents such as NaBH₄ supported on alumina, achieving high to excellent yields of secondary amines at . For instance, grinding aldehydes with primary amines and NaBH₄/alumina under solvent-free conditions proceeds efficiently, with isolated yields often exceeding 90% for various aromatic and aliphatic substrates. Similarly, one-pot procedures employing NaBH₄ with SBA-15-supported trifunctional sulfonated catalysts facilitate reductive amination of ketones and aldehydes without solvents, demonstrating broad substrate scope and reaction times of 10–60 minutes. More advanced mechanochemical approaches, such as Pd-coated ball milling under ambient H₂ pressure, allow ligand-free reductive amination of primary and secondary amines with carbonyls, yielding products in up to 99% efficiency while minimizing waste. Continuous flow processes further enhance the of reductive amination by integrating microreactors that enable precise over H₂ delivery and reaction parameters. These systems typically operate with short times of 5–30 minutes, delivering gaseous H₂ directly into liquid streams via gas-permeable membranes or packed- configurations, resulting in yields often above 90% for reductions. For example, nickel-catalyzed reductive amination in a micropacked reactor converts aldehydes and to primary amines continuously, with space-time yields up to 10 g L⁻¹ h⁻¹ under mild conditions (50–100°C, 10–30 bar H₂). Pd/C-based micro-packed reactors have also been employed for the reductive amination of , achieving >95% conversion in times as low as 10 minutes while maintaining stability over extended operation. The adoption of solvent-free and continuous flow methods in reductive amination offers significant advantages, including substantial reductions in (VOC) emissions—often by over 90% compared to batch processes—and improved through the controlled handling of H₂ in small reactor volumes, mitigating risks. These processes are particularly valuable for active pharmaceutical ingredient () synthesis; a notable example is the 2020 continuous flow reductive amination step in the scalable production of the cationic SST-01, which integrated aerobic oxidation and achieved >85% overall yield while enabling safe handling of hazardous intermediates. Compatibility with eco-friendly reducing agents, such as those discussed in related sustainable methods, further amplifies their utility in industrial settings.

Recent advances in sustainable methods

Recent developments in reductive amination have emphasized through innovative catalytic systems that minimize waste, utilize renewable feedstocks, and enhance . A notable 2020 review highlighted breakthroughs in catalytic reductive amination of aldehydes and ketones with nitro compounds using non-noble metal catalysts like , , and , achieving high turnover numbers (up to 3,800) under milder conditions than traditional methods, with Mo₃S₄ clusters enabling >99% conversion at 70°C and 20–50 H₂, promoting greener alternatives to stoichiometric reductants. In 2022, iridium-catalyzed direct asymmetric reductive amination using primary alkyl amines as nitrogen sources was reported, employing a low-loading (0.05 mol%) Ir-phosphoramidite complex to deliver chiral amines with up to 97% enantioselectivity at 40°C and 40 atm H₂, scalable to gram quantities and tolerant of diverse functional groups, reducing reliance on preformed imines. Bio-based approaches have gained traction for converting renewable furanic oxygenates into valuable amines. A 2023 study detailed for reductive amination of and 5-hydroxymethylfurfural (HMF), with non-noble /Al₂O₃ and /Nb₂O₅ catalysts yielding up to 99% furfurylamine and 96% 5-(hydroxymethyl)-2-furfurylamine under mild pressures (0.1–4 H₂) and temperatures (90–100°C), leveraging biomass-derived substrates to avoid feedstocks and producing as the main byproduct. Electrochemical methods further advance in reductive amination, as outlined in a 2025 review, where cathodic electroreduction facilitates deoxygenative coupling of carbonyls without stoichiometric reagents, enabling sustainable synthesis of amines from oxygenates under ambient conditions and reducing compared to processes. Multi-component strategies have emerged to streamline while enhancing efficiency. In 2024, a cationic complex enabled tandem reductive amination-alkylation of esters with , achieving N-monoalkylation in one pot at 0.1 mol% loading and , with broad substrate tolerance and scalability to 15 g, minimizing steps and use for complex construction. Complementing this, direct reductive N-alkylation of with carboxylic esters using tert-butylmagnesium chloride and Red-Al was developed in 2024, delivering tertiary in up to 90% yield at 0°C without isolating intermediates, applicable to pharmaceutical targets like and favoring common esters for reduced environmental footprint. A 2025 advance introduced a Ni-doped MFM-300(Cr) metal-organic framework for the reductive amination of aldehydes and ketones to primary amines using NH₃ and H₂, achieving >90% across 38 substrates including biomass-derived ones, under mild conditions (160°C, 5 bar H₂ in MeOH). The earth-abundant Ni-based is reusable for at least 5 cycles with >98% retention and no , promoting by replacing metals and enabling efficient conversion of renewables. Broader trends in these advances pursue 100% by prioritizing hydrogen-borrowing and direct routes that generate only water, as seen in conversions yielding >95% selectivity with recyclable s like Ru-PNP complexes. Life-cycle assessments underscore the benefits, showing -derived processes cut versus routes and enable cost-effective scaling (e.g., ~US$1973/ton for pentanediol ), guiding industrial adoption through evaluations of , stability, and minimization.

References

  1. [1]
    Reductive Amination in the Synthesis of Pharmaceuticals
    ### Summary of Reductive Amination from https://pubs.acs.org/doi/10.1021/acs.1021/acs.chemrev.9b00383
  2. [2]
    [PDF] Myers Chem 115
    Overview: • The reductive amination of aldehydes and ketones is an important method for the synthesis of primary, secondary, and tertiary amines.
  3. [3]
    Catalytic Reductive Alkylation of Amines in Batch and Microflow ...
    Oct 9, 2020 · Reductive alkylation of amines (also known as reductive amination of carbonyl compounds) plays an important role in pharmaceutical and ...
  4. [4]
    Imine Reductases and Reductive Aminases in Organic Synthesis
    Sep 12, 2024 · Imine reductases (IREDs) and reductive aminases (RedAms) are NADPH-dependent oxidoreductases catalyzing C–N redox reactions.
  5. [5]
    Reductive Amination - an overview | ScienceDirect Topics
    Reductive amination is the second mechanism biochemistry uses to synthesize amino acids alongside transamination.
  6. [6]
  7. [7]
    Hitchhiker's Guide to Reductive Amination - Thieme Connect
    Most important advantages and drawbacks of the methods, such as selectivity of the target amine formation and toxicity of the reducing agents were compared.
  8. [8]
    Current status and potential value for chiral amine synthesis
    Jun 16, 2022 · Approximately 40% of commonly used drugs in the United States contained chiral amines as the core moieties (Figure 1). Thus, various chemo- and ...
  9. [9]
    Advancement in the Synthesis of Amine through the Leuckart Reaction
    Jan 29, 2023 · The monoterpene amine bornylamine was first synthesized by Leuckart in 1887 through the reaction of camphor and formamide, and then, by the ...<|control11|><|separator|>
  10. [10]
    Cyanohydridoborate anion as a selective reducing agent
    Organocatalytic Reductive Amination of Aldehydes with 2-Propanol: Efficient Access to N-Alkylated Amines. The Journal of Organic Chemistry 2025, Article ASAP.
  11. [11]
    Utilizing the Imine Condensation in Organic Chemistry Teaching ...
    Oct 26, 2023 · Imines contain a carbon–nitrogen double bond and are typically formed via reversible condensation reactions of a carbonyl containing compound ...
  12. [12]
    20.04: Review: Reactions of Aldehydes and Ketones with Amines
    ### Summary of Imine Formation Mechanism from Aldehydes/Ketones and Primary Amines
  13. [13]
    Dynamic stereochemistry of imines and derivatives. 19. Mutarotation ...
    ChemInform Abstract: Dynamic Stereochemistry of Imines and Derivatives. Part 19. Mutarotation and E‐Z Isomerization of Chiral Imines in [2H4]Methanol Solution..
  14. [14]
    Density Functional Theory Study on the Selective Reductive ...
    Aug 19, 2022 · Reductive amination is one of the most important methods to synthesize amines, having a wide application in the pharmaceutical, ...
  15. [15]
  16. [16]
  17. [17]
  18. [18]
  19. [19]
    A Review on the Use of Sodium Triacetoxyborohydride in the ...
    A review on the use of sodium triacetoxyborohydride in the reductive amination of ketones and aldehydes.
  20. [20]
    Reduction of Schiff Bases with Sodium Borohydride
    Reductive Amination of Aldehydes and Ketones with 2-(Tributylamino)-ethoxyborohydride. Monatshefte für Chemie - Chemical Monthly 2007, 138 (11) , 1187-1189 ...
  21. [21]
    Reductive Amination of Aldehydes and Ketones with Sodium ...
    Sodium triacetoxyborohydride is presented as a general reducing agent for the reductive amination of aldehydes and ketones.Introduction · Results and Discussions · Supporting Information Available
  22. [22]
  23. [23]
    Unsupported Nanoporous Platinum Catalyst for the Chemoselective ...
    Sep 3, 2025 · Unsupported nanoporous platinum (PtNPore) enables efficient, chemoselective hydrogenation of imines and reductive amination of benzaldehydes ...
  24. [24]
  25. [25]
    Underlying Mechanisms of Reductive Amination on Pd-Catalysts - NIH
    Review of Methods for the Catalytic Hydrogenation of Carboxamides. Chem ... reductive amination using catalytic hydrogenation with Pd/C. Tetrahedron ...Missing: paper | Show results with:paper
  26. [26]
    A Single Step Reductive Amination of Carbonyl Compounds with ...
    Highly chemoselective `one pot' amination of carbonyl groups is achieved using polymethylhydrosiloxane (PMHS) as reductant and Ti(OiPr)4 as activator.Missing: catalyst | Show results with:catalyst
  27. [27]
    One-pot reductive amination of aldehydes and ketones with α ...
    A one-pot reductive amination of aldehydes and ketones with amines using α-picoline-borane as a reducing agent is described.Missing: Satoh | Show results with:Satoh
  28. [28]
    Direct Reductive Amination of Carbonyl Compounds Catalyzed by a ...
    Solid imines were dried under vacuum and stored under N2, while liquid imines and aldehydes were degassed, dried over 4 Å molecular sieves and stored under N2.
  29. [29]
    [PDF] Development of a Solvent Selection Guide for Aldehyde-based ...
    All reactions were carried out using conventional glassware at room temperature (generally approx. 18 °C) under an air atmosphere and with no special attention ...
  30. [30]
    Development and Manufacturing GMP Scale-Up of a Continuous Ir ...
    The design, development, and scale up of a continuous iridium-catalyzed homogeneous high pressure reductive amination reaction to produce 6, the penultimate ...
  31. [31]
    Reductive Amination - ACS GCI Pharmaceutical Roundtable
    2-Picoline borane is available in bulk and is easy to handle on a pilot scale. The stability of both reagents towards hydrolysis or methanolysis allows ...Missing: aqueous biocompatible<|control11|><|separator|>
  32. [32]
    Method for Screening Sodium Cyanoborohydride for Free Cyanide ...
    Feb 6, 2025 · Since the seminal work by Borch in 1971, (26) the direct reductive amination (Borch reaction) using sodium cyanoborohydride (CBH) has become ...
  33. [33]
    Continuous Synthesis of Safinamide Mesylate using Flow Reactions ...
    Sep 3, 2025 · A continuous synthesis of 5·MsOH via C–O bond formation, reductive amination, extraction, and crystallization was conducted using small reactors ...<|separator|>
  34. [34]
  35. [35]
    Economy Assessment for the Chiral Amine Production with ... - MDPI
    Dec 11, 2020 · The results indicated that enhancing the activity of amine dehydrogenase by 4–5 folds can drop the unit price of reductive amination to $0.5–0.6 ...<|separator|>
  36. [36]
    Borrowing Hydrogen for Organic Synthesis | ACS Central Science
    Mar 25, 2021 · Borrowing hydrogen is a process that is used to diversify the synthetic utility of commodity alcohols. A catalyst first oxidizes an alcohol ...
  37. [37]
    Recent advances in homogeneous borrowing hydrogen catalysis ...
    Sep 17, 2018 · This review will describe the recent advances (2013-present) in homogeneous borrowing hydrogen catalysis using earth-abundant first row transition metal ...
  38. [38]
    Recent Advances in the Enantioselective Synthesis of Chiral Amines ...
    Oct 22, 2021 · We provide an overview of the recent advances in the AH of imines, enamides, enamines, allyl amines, and N-heteroaromatic compounds.
  39. [39]
    [PDF] Ryoji Noyori - Nobel Lecture
    Enantioselectivity of BINAP-Rh (I) catalyzed asymmetric hydrogenation of a-(acylamino) acrylic acids was highly variable and not sa- tisfactory at that time, ee ...
  40. [40]
    Reductive Amination of Carboxylic Acids via Nickel(II) Catalysis
    Oct 1, 2025 · A nickel(II)-catalyzed reductive amination of carboxylic acids with amines is described, employing phenylsilane as the reductant.
  41. [41]
    Catalytic Reductive Amination and Tandem Amination–Alkylation of ...
    Dec 10, 2024 · Reported herein is a convenient and efficient method for one-pot, catalytic reductive amination, as well as the first multi-component tandem ...
  42. [42]
    Review Recent advances in electrochemically driven deoxygenation ...
    Sep 30, 2025 · This review introduces both direct and indirect methods for the activation and transformation of C−O or C=O bonds through reductive ...
  43. [43]
    Electrochemical reductive amination of furfural-based biomass ...
    Electrochemical reductive amination uses water as the hydrogen source without requiring other chemicals as the reducing agent, providing reaction conditions ...<|separator|>
  44. [44]
  45. [45]
    Recent advances of visible-light photocatalysis in the ...
    Visible-light photocatalysis offers unique opportunities to achieve smooth and clean functionalization of organic compounds by unlocking site-specific ...
  46. [46]
    Catalytic Reductive Amination of Aldehydes and Ketones With Nitro ...
    In this short overview, recent advances in the methodology and application of the catalytic reductive amination with nitro compounds are outlined.
  47. [47]
    Rapid Synthesis of Psychoplastogenic Tropane Alkaloids | JACS Au
    Oct 5, 2023 · Moreover, a one-step global reduction/reductive amination sequence seemed to be a viable strategy to rapidly access tropanes that were ...
  48. [48]
    [PDF] EPC SYNTHESIS OF TROPANE ALKALOIDS VIA ...
    from chiral terpenes by 'a one pot" reductive amination reaction (Scheme 69). The reductive amination was most conveniently accomplished in "one pot" by.
  49. [49]
    Investigation of the l-phenylacetylcarbinol process to substituted ...
    Reductive amination of these l-PAC analogues was able to generate viable amounts of the 4-methyl and 4-fluoro analogues of pseudoephedrine/ephedrine but was ...
  50. [50]
    Synthesis of Morphine Alkaloids and Derivatives - ResearchGate
    May 6, 2011 · ... (1980). Rice achieved the shortest and the most efficient formal synthesis of morphine via ... reductive amination before the dihydrofuran. ring was ...<|control11|><|separator|>
  51. [51]
    Solid-Phase Synthesis of Diverse Peptide Tertiary Amides By ...
    An alternative to the use of chiral bromides and amines to create PTAs would be to build them from amino acids and aldehydes using reductive amination (Figure ...
  52. [52]
    [PDF] The double reductive amination approach to the synthesis of ...
    In these hydrogenolytic conditions, deprotection of the terminal amine followed by DRA occurred, providing a 4:1 mixture of diastereoisomeric N-alkylated.
  53. [53]
    [PDF] The Reductive Amination of Aldehydes and Ketones and the ...
    Reductive amination of aldehydes/ketones and hydrogenation of nitriles are methods for amine preparation, using an imine intermediate, and often producing a ...<|separator|>
  54. [54]
    Biocatalytic Imine Reduction and Reductive Amination of Ketones
    May 7, 2015 · This review provides a comprehensive overview of biocatalytic imine reduction and reductive amination of ketones, highlighting the natural roles ...
  55. [55]
    Mining the cellular inventory of pyridoxal phosphate-dependent ...
    Pyridoxal phosphate (PLP) is an enzyme cofactor required for the chemical transformation of biological amines in numerous essential cellular processes.
  56. [56]
    Quantifying Reductive Amination in Nonenzymatic Amino Acid ... - NIH
    Amino acid biosynthesis initiates with the reductive amination of α‐ketoglutarate with ammonia to produce glutamate. However, the other α‐keto acids derived ...
  57. [57]
    Kinetic studies of glutamate dehydrogenase. The reductive ... - NIH
    Kinetic studies of the reductive amination of 2-oxoglutarate catalysed by glutamate dehydrogenase with NADH and NADPH as coenzyme were made at pH7.0 and pH 8.0.
  58. [58]
    The fungal α-aminoadipate pathway for lysine biosynthesis requires ...
    The fungal α-aminoadipate pathway for lysine biosynthesis requires two enzymes of the aconitase family for the isomerization of homocitrate to homoisocitrate.
  59. [59]
    Glutamate Metabolism in the Brain Focusing on Astrocytes - NIH
    1). The thermodynamic equilibrium constant (6 × 10−15 M) favors the reductive amination of glutamate (Engel & Dalziel 1967).
  60. [60]
    Amino Acid Neurotransmitter Synthesis and Removal | Neuroglia
    The GDH reaction has to operate in both directions for these shuttles to operate, that is, reductive amination in neurons and oxidative deamination in ...
  61. [61]
    Biocatalysis in Drug Design: Engineered Reductive Aminases ... - NIH
    Oct 2, 2023 · We report a biocatalytic process to access a specific diastereomer of a chiral amine building block used in drug discovery.<|control11|><|separator|>
  62. [62]
    Reductive Amination in the Synthesis of Pharmaceuticals - PubMed
    Dec 11, 2019 · This Review concisely compiles information on 71 medical substances that are synthesized by reductive amination.<|control11|><|separator|>
  63. [63]
    Total synthesis of fluoxetine and duloxetine through an in situ imine ...
    We report efficient, catalytic, asymmetric total syntheses of both (R)-fluoxetine and (S)-duloxetine from α,β-unsaturated aldehydes conducting five sequential ...Missing: amination | Show results with:amination
  64. [64]
    Synthesis of anti-allergic drugs - RSC Publishing
    Feb 4, 2020 · The nal step included the alkylation or reductive amination and the removal of Boc to obtain the nal product (compounds 5a–n, 6a–d and. 7a–h).
  65. [65]
    Synthesis and Performance of Biobased Surfactants Prepared ... - NIH
    Reductive amination of carbonyl-containing compounds is an effective method to form C–N bonds. This reaction can be used to selectively modify monosaccharides ...Missing: ethanolamines | Show results with:ethanolamines
  66. [66]
    Methods for making ethanolamine(s) and ethyleneamine(s) from ...
    The conversion of ethanolamines by the reductive amination reaction can be in the range of from 20% to 80%, preferably in the range of from 30% to 50%. In ...
  67. [67]
    Reductive aminations by imine reductases: from milligrams to tons
    In this perspective we trace the development of the IRED-catalyzed reductive amination reaction from its discovery to its industrial application on kg to ton ...Missing: surfactants | Show results with:surfactants
  68. [68]
    A new target region for changing the substrate specificity of amine ...
    Jun 1, 2015 · An improved R-ATA, ATA-117-Rd11, was successfully engineered for the manufacture of sitagliptin, a widely used therapeutic agent for type-2 diabetes.
  69. [69]
    Advances in One-Pot Chiral Amine Synthesis Enabled by Amine ...
    Apr 10, 2023 · The aims of these approaches are higher yields, decreased costs, and environmental advantages, preferably combined with high enantio- and ...Missing: savings | Show results with:savings
  70. [70]
    Amine synthesis by reductive amination (reductive alkylation)
    Reductive amination between aldehydes or ketones and amines occurs smoothly within the hydrophobic cores of nanomicelles in water. A broad range of substrates ...
  71. [71]
    Polymer Supported Triphenylphosphine-Palladium Acetate ... - MDPI
    Indirect reductive amination of aldehydes, catalyzed by polymer supported triphenylphosphine-palladium acetate complex PS-TPP-Pd(OAc)2 catalyst have been ...
  72. [72]
    Efficient reductive amination of HMF with well dispersed Pd ...
    The present study illustrates a new method for the design of metal–organic framework (MOF)/polymer composites containing well-defined metal nanoparticles in a ...<|separator|>
  73. [73]
    Synthesis of Secondary Amines from One-Pot Reductive Amination ...
    However, the one-pot reductive amination with formic acid as the hydrogen donor has been rarely studied, (23-25) and the reported methods were performed over ...
  74. [74]
    One-pot reductive amination of aldehydes with nitroarenes using ...
    One-pot reductive amination of aldehydes with nitroarenes using formic acid as the hydrogen donor and mesoporous graphitic carbon nitride supported AgPd alloy ...
  75. [75]
    Chemoselective Reductive Amination of Carbonyl Compounds for ...
    Chemoselective Reductive Amination of Carbonyl Compounds for the Synthesis of Tertiary Amines Using SnCl2·2H2O/PMHS/MeOH | The Journal of Organic Chemistry.
  76. [76]
    Efficient ruthenium catalyzed transfer hydrogenation of ...
    Transfer hydrogenation of various functionalized imines by isopropanol catalyzed by [Ru(CO)2(Ph4C4CO)]2 (3) has been studied. The use of either an oil bath ...
  77. [77]
    Engineered Biocatalysts for Enantioselective Reductive Aminations ...
    Mar 22, 2023 · Reductive aminases (RedAms) have recently emerged as promising biocatalysts for the synthesis of chiral secondary amines by coupling primary ...
  78. [78]
    Biocatalytic reductive aminations with NAD(P)H-dependent enzymes
    Dec 7, 2023 · In this review, we discuss advances in developing AmDHs and IREDs as biocatalysts and focus on evolutionary history, substrate scope and applications of the ...
  79. [79]
    Practical reduction of imines by NaBH 4 /alumina under solvent-free ...
    Sodium borohydride supported on alumina reduces imines to the corresponding secondary amines in high to excellent isolated yields under solvent-free conditions.
  80. [80]
    [PDF] One-pot Reductive Amination of Carbonyl Compounds with NaBH4 ...
    An efficient one-pot procedure for the direct reductive amination of aldehyde and ketones was achieved in the presence of sodium borohydride by using B(OSO3H)3/ ...Missing: seminal paper
  81. [81]
    Ligand-free reductive amination via Pd-coated mechanocatalysis
    Oct 23, 2025 · The application of H2-driven reductive amination under solvent-free conditions remains notably underexplored. Mechanochemistry has emerged ...<|separator|>
  82. [82]
    Continuous reductive amination to synthesize primary amines with ...
    In this work, we propose a continuous reductive amination process in a micropacked bed reactor using nickel catalysts.Missing: shift | Show results with:shift
  83. [83]
    Continuous flow reductive amination of cyclohexanone using Pd/C ...
    We developed an environmentally benign continuous flow reductive amination system employing a micro ... time yields (2.7 × 104 g L−1 h−1), and operates ...Missing: residence | Show results with:residence
  84. [84]
    Catalytic Continuous Reductive Amination with Hydrogen in Flow ...
    Jan 8, 2024 · The advantages of flow reactors, including temperature control, mass transfer intensification, and residence time distribution, have ...
  85. [85]
    Iridium-catalyzed direct asymmetric reductive amination utilizing ...
    Jun 10, 2022 · In this study we describe primary alkyl amines effectively serve as the N-sources in direct asymmetric reductive amination catalyzed by the iridium precursor.
  86. [86]
    Recent Advances in the Efficient Synthesis of Useful Amines ... - MDPI
    In this review, we summarize and discuss the recent significant progress in the generation of useful amines from bio-based furanic oxygenates with H2 and ...
  87. [87]
    Direct Reductive N‐alkylation of Amines with Carboxylic Esters
    Nov 29, 2024 · We report herein a facile direct method for the reductive N-alkylation of amines with carboxylic esters, using tert-butylmagnesium chloride ...
  88. [88]
    Toward Renewable Amines: Recent Advances in the Catalytic ...
    In this review, we will give a concise overview of the up-to-date methods for the production of industrially important amines from biomass-derived oxygenates.