Fact-checked by Grok 2 weeks ago

Amide reduction

Amide reduction is a key transformation in wherein the of an is cleaved and reduced to yield either amines or aldehydes, enabling the construction of these vital functional groups from abundant amide starting materials. The traditional approach utilizes strong donors such as lithium aluminum (LiAlH₄) in solvents, which reduces primary, secondary, and tertiary amides to the corresponding primary, secondary, and tertiary amines through initial of to the carbonyl, formation of an ion intermediate, and final reduction, often achieving high yields like 95% for N-methyldodecanamide to N-methyldodecylamine. Selective reduction of amides to aldehydes, rather than over-reduction to amines, can be accomplished with milder like (Sia₂BH) or (Cp₂Zr(H)Cl), which proceed via or hydrozirconation mechanisms that halt at the aldehydic stage after , offering utility in sensitive substrates. Contemporary methods emphasize for enhanced and , including nickel- or zinc-catalyzed hydrosilylation with silanes under mild conditions that tolerate esters and groups, as well as ruthenium pincer complex-mediated with molecular H₂ at moderate pressures (30–50 bar) and temperatures (up to 160°C), producing amines via with preservation of the C–N bond while generating minimal waste. These catalytic protocols, often employing earth-abundant metals, address limitations of stoichiometric reductants like hazardous byproducts and poor tolerance, thereby supporting efficient of pharmaceuticals and fine chemicals in line with goals.

Fundamentals

Definition and Scope

reduction encompasses the chemical transformation of the in , characterized by the general structure R-C(O)-NR_2, into either a methylene (CH_2) unit yielding (R-CH_2-NR_2) or a formyl (CHO) group producing (R-CHO). This reaction serves as a cornerstone in , enabling the conversion of stable precursors into valuable or derivatives. The process applies broadly to primary (NR_2 = NH_2), secondary (NR_2 = NHR'), and tertiary (NR_2 = NR'R'') , with selectivity determining the product outcome—full reduction typically affords , while partial reduction targets . The inherent stability of amides stems from resonance delocalization of the lone pair into the carbonyl π-system, which imparts partial double-bond character to the C-N bond and diminishes the electrophilicity of the carbonyl carbon relative to other carbonyl compounds such as aldehydes, ketones, or esters. This resonance stabilization necessitates the use of potent reducing agents to overcome the kinetic and thermodynamic barriers, distinguishing amide reduction from more facile reductions of less stabilized carbonyls. Historically, the reduction of amides using metal s was first documented in the late 1940s, marking the advent of reliable methods for this transformation shortly after the discovery of aluminum hydride in 1947. Amide reduction holds significant importance in the of pharmaceuticals, agrochemicals, and natural products, where it facilitates the late-stage of or moieties essential for and structural diversity.

Types of Products

Amide reduction reactions can produce distinct classes of products depending on the degree of and the substitution pattern of the . Full of primary amides (RCONH₂) proceeds to primary amines (RCH₂NH₂), involving the of the C=O bond and addition of four equivalents of , as represented by the general equation: \text{RCONH}_2 + 4[\text{H}] \rightarrow \text{RCH}_2\text{NH}_2 + \text{H}_2\text{O} This transformation effectively reduces the carbonyl to a . Full of secondary amides (RCONHR') proceeds to secondary amines (RCH₂NHR'), involving the of the C=O bond and addition of four equivalents of , as represented by the general equation: \text{RCONHR'} + 4[\text{H}] \rightarrow \text{RCH}_2\text{NHR'} + \text{H}_2\text{O} This transformation effectively reduces the carbonyl to a methylene group while retaining the nitrogen substituent. Full reduction of tertiary amides (RCONR'R'') proceeds to tertiary amines (RCH₂NR'R''), following a similar pattern. In contrast, partial reduction of primary amides (RCONH₂) can generate aldehydes (RCHO) and ammonia (NH₃), requiring only two equivalents of hydrogen and halting at the aldehydic stage: \text{RC(O)NH}_2 + 2[\text{H}] \rightarrow \text{RCHO} + \text{NH}_3 Partial reduction targets N,N-disubstituted (tertiary) amides (RC(O)NR₂) to generate aldehydes (RCHO) and the corresponding amine (HNR₂), requiring only two equivalents of hydrogen and halting at the aldehydic stage after hydrolysis: \text{RC(O)NR}_2 + 2[\text{H}] \rightarrow \text{RCHO} + \text{HNR}_2 This selective process exploits the amide's resonance stabilization to limit over-reduction. Achieving selectivity between and products hinges on factors such as steric hindrance around the carbonyl, the nature of the , and conditions like temperature and solvent, which collectively mitigate risks of over-reduction to hydrocarbons or formation of byproducts. Rare special cases include complete to hydrocarbons (e.g., RCH₃ from RCONHR), typically requiring specialized systems beyond standard protocols. Key challenges in these reductions encompass preventing side products like alcohols from competing pathways and ensuring high yield in -directed syntheses by avoiding further reduction to primary alcohols.

Reduction to Amines

Catalytic Hydrogenation

Catalytic hydrogenation represents a classical method for the complete reduction of amides to amines, leveraging molecular hydrogen and heterogeneous catalysts to overcome the inherent stability of the amide bond due to resonance delocalization between the carbonyl and nitrogen lone pair. This process typically proceeds via a stepwise mechanism involving initial activation of the carbonyl group on the catalyst surface, followed by sequential addition of hydrogen to form intermediates such as hemiaminals or iminols, and ultimately cleavage of the C-N bond to yield the amine product and water. The overall transformation can be represented by the general equation: \mathrm{RC(O)NR'_2 + 2H_2 \rightarrow RCH_2NR'_2 + H_2O} where \mathrm{R} and \mathrm{R'} denote alkyl or aryl substituents. The reaction requires elevated temperatures and pressures to achieve practical rates, with typical conditions ranging from 150–250°C and 100–300 atm of hydrogen pressure. Common catalysts include copper-chromium oxide (CuCr₂O₄, often promoted with barium or manganese oxides), Raney nickel, and rhenium heptoxide (Re₂O₇), which facilitate hydrogen activation and substrate adsorption. For instance, copper-chromium oxide catalysts, pioneered by Homer Adkins in the 1930s, effectively reduce primary, secondary, and tertiary amides under these harsh conditions, though selectivity can vary with the amide substitution pattern. Raney nickel operates similarly at around 250°C, offering good activity for aliphatic amides but requiring careful handling due to its pyrophoric nature. A representative example is the reduction of (CH₃CONHC₆H₅) to N-ethylaniline (CH₃CH₂NHC₆H₅), achieved using copper-chromium oxide catalyst at 200–225°C and 200–300 atm, yielding the product in high conversion after several hours. This method has found industrial application in the synthesis of amine precursors for polymers, particularly through the of fatty acid amides to aliphatic amines used in production. The primary advantages of catalytic lie in its scalability for large-scale manufacturing, as it employs inexpensive gas and avoids stoichiometric reductants, producing only water as a . However, the need for high-pressure equipment and elevated temperatures limits its use in synthesis, while catalysts like copper-chromium oxide pose environmental concerns due to , necessitating careful disposal and recovery protocols. Developed primarily in and 1940s by researchers such as Adkins for the reduction of amides, this approach laid the foundation for subsequent milder catalytic innovations.

Hydride Reduction Methods

Hydride reduction methods employ stoichiometric metal s to convert amides into amines by delivering hydride ions to the , resulting in complete and reduction of the C-N bond to a . aluminum hydride (LiAlH₄) is the archetypal for this transformation, applicable to primary, secondary, and tertiary amides, yielding the corresponding amines in good to excellent yields under refluxing conditions in ether or solvents. The general reaction can be represented as: \text{RC(O)NR}_2 + 4[\text{H}] \xrightarrow{\text{LiAlH}_4} \text{RCH}_2\text{NR}_2 + \text{H}_2\text{O} This method, first demonstrated in the late 1940s, provides a straightforward route for laboratory-scale synthesis but requires careful handling due to the reagent's reactivity. The mechanism begins with nucleophilic attack by hydride on the amide carbonyl, forming a tetrahedral intermediate, followed by elimination of the amine leaving group to generate an iminium ion intermediate. Subsequent hydride additions reduce the iminium to the final amine product. Primary amides exhibit high selectivity, often proceeding without significant side reactions, as exemplified by the reduction of benzamide to benzylamine in diethyl ether, affording the product in approximately 80% yield after acidic workup. Variations include borane complexes such as , which offer faster reaction rates, particularly for tertiary amides, due to the reagent's milder conditions and compatibility with certain functional groups like olefins. Borane reductions proceed via coordination to the carbonyl oxygen, enhancing electrophilicity and facilitating transfer, often completing in hours at . For milder conditions, (NaBH₄) combined with additives like iodine or titanium(IV) chloride enables selective reduction of secondary and tertiary amides to amines, avoiding the harshness of LiAlH₄ while achieving yields up to 76% in . Diisobutylaluminum hydride (DIBAL-H) is occasionally used for formation with excess reagent but is primarily noted for partial to aldehydes (see Reduction to Aldehydes section). Limitations of these hydride methods include risks of over- for substrates with sensitive groups like or cyano functionalities, as well as the necessity for rigorous aqueous workup to hydrolyze aluminum or salts, which can complicate purification.

Alternative Catalytic Approaches

One prominent class of alternative catalytic approaches involves the use of as hydride donors in the presence of earth-abundant metal catalysts, such as iron or , to achieve deoxygenative reduction of amides to amines under mild conditions. In a seminal 2009 report, Beller and coworkers demonstrated the first general iron-catalyzed hydrosilylation of secondary and tertiary amides using (PMHS) as the reducing agent and Fe₃(CO)₁₂ as the precatalyst, proceeding at 65–100 °C in to afford amines in yields up to 99% for a range of substrates including aliphatic and aromatic amides. The reaction leverages the activation of the silane by low-valent iron species, facilitating C=O bond cleavage and transfer, with byproducts forming siloxanes that can be easily separated. Similarly, in 2017, and colleagues developed a nickel-catalyzed protocol employing NiCl₂(dme) with phenylsilane (PhSiH₃) at 115 °C, enabling selective reduction of secondary and tertiary amides, including lactams, to amines in yields exceeding 90% while tolerating esters and preserving stereocenters. These silane-mediated methods offer advantages over traditional by operating at lower pressures and temperatures, avoiding the need for high-pressure gas, and utilizing inexpensive, non-toxic reducing agents. A distinct two-step catalytic route involves initial thionation of amides to thioamides using , followed by desulfurization with to yield amines under ambient conditions. This method, detailed in a 2021 review by Yousuf et al., proceeds via reflux in for thionation (yields >80% for diverse amides) and subsequent room-temperature treatment with in or acetone, affording amines without harsh reductants. The process benefits from the mildness of , which selectively removes while preserving sensitive functionalities, making it suitable for complex molecule synthesis. In the 2020s, cobalt-based gained traction for hydroamination-like reductions of amides. Beller and coworkers reported a homogeneous Co(NTf₂)₂/(p-anisyl)triphos system with [Me₃SiOTf] cocatalyst, enabling deoxygenative at 100 °C and 30 bar H₂ to produce amines from secondary and tertiary amides in yields up to 95%, with broad substrate scope including cyclic lactams. This method underscores cobalt's efficacy in promoting C-O cleavage via intermediates, offering scalability and tolerance to functional groups like halides and alkenes. Overall, these alternatives provide versatile, operationally simple pathways that expand beyond conventional , emphasizing through base metals and mild reagents.

Reduction to Aldehydes

Noncatalytic Methods

One of the cornerstone noncatalytic approaches to partial amide reduction involves the use of diisobutylaluminum hydride (DIBAL-H) as a stoichiometric , particularly effective for converting amides to aldehydes under low-temperature conditions, typically at -78°C in solvents like or . The reaction proceeds via a single transfer, represented by the equation: \ce{RC(O)NR'_2 + DIBAL-H -> RCHO + (iBu)_2AlNR'_2} This selectivity arises because the aluminum coordinates to both the emerging aldehyde oxygen and the nitrogen lone pair, forming a stable five-membered chelate that inhibits further hydride addition or over-reduction to the alcohol. The method is broadly applicable to aliphatic, aromatic, and heteroaromatic tertiary amides, often delivering yields exceeding 80% when 1-1.2 equivalents of DIBAL-H are employed. A specialized application leverages Weinreb amides (N-methoxy-N-methylamides), which undergo clean reduction to aldehydes with DIBAL-H, even at temperatures up to -40°C, owing to enhanced chelation involving the methoxy oxygen that stabilizes the intermediate and blocks excess reagent reactivity. This variant has found utility in complex syntheses. The Sonn-Müller method represents an earlier stoichiometric route, primarily for N-aryl amides, entailing activation with phosphorus pentachloride (PCl5) to generate an imidoyl chloride intermediate, followed by reduction with tin(II) chloride in concentrated hydrochloric acid to form an iminium salt, and mild hydrolysis to the aldehyde. Though less prevalent in modern practice due to its harsh conditions and lower functional group tolerance, it remains relevant for specific aromatic systems where hydride methods falter. Under carefully controlled conditions, such as modified formulations or specific solvent systems, aluminum (LiAlH₄) can partially reduce N,N-dialkylamides to aldehydes, for example, via 1-acylaziridine intermediates with LiAlH₄, or other variants like tri-tert-butoxyaluminum , to enhance selectivity and prevent progression to amines. These adaptations deliver moderate to good yields for dialkyl substrates but demand precise stoichiometry to minimize over-reduction. Despite their efficacy, these noncatalytic methods share limitations, including high sensitivity to over-reduction if temperatures exceed -78°C or excess is used, as the chelate can dissociate, allowing secondary additions leading to alcohols in up to 20-30% yields under suboptimal conditions. All require rigorously environments, as the organoaluminum or metal s react violently with moisture, complicating handling and necessitating inert atmospheres.

Catalytic Methods

Catalytic methods for the reduction of to primarily rely on hydrosilylation strategies, which employ catalysts to facilitate the addition of silanes to the carbonyl, yielding silyl-protected intermediates that to . These approaches offer enhanced selectivity compared to traditional stoichiometric reductions, minimizing over-reduction to or alcohols. One seminal method involves hydrozirconation using , Cp₂ZrHCl, which reacts with at to form a zirconacycle intermediate, followed by to the . This process achieves high yields (up to 95%) for aromatic and aliphatic , including Weinreb amides, while tolerating esters and other functional groups. Although typically stoichiometric in the reagent, its mild conditions (0–25°C, solvent, 1–2 hours) and have inspired catalytic variants. The mechanism proceeds via η²-coordination of the to , followed by migration and elimination of the . Titanium-catalyzed hydrosilylation represents a truly catalytic advancement, particularly from developments in the . Using Ti(OiPr)₄ (5–10 mol%) with 1,1,3,3-tetramethyldisiloxane (TMDS) as the hydrosilane, secondary and tertiary amides are reduced to aldehydes in good yields (70–90%) under mild conditions (, 24 hours, solvent). For example, N,N-diethylbenzamide yields selectively without over-reduction. This system extends to primary aromatic amides, though with moderate efficiency, and avoids the toxicity of diphenylsilane by employing the cheaper TMDS. The reaction likely involves formation of a titanium-silyl species that adds across the C=O bond, generating a silyl that eliminates to the aldehyde upon . Ruthenium-catalyzed hydrosilylation provides an alternative organometallic route, often via activation of the amide to an iminoyl chloride intermediate. Treatment of secondary amides with oxalyl chloride generates the iminoyl chloride, which undergoes selective hydrosilylation with PhMe₂SiH using a ruthenium catalyst (e.g., RuH₂(PPh₃)₄, 1–5 mol%) at 50°C, yielding the aldehyde after hydrolysis (yields 60–85%). This two-step process adapts principles from classical reductions like the Rosenmund reaction for acid chlorides, achieving selectivity for aldehyde formation from aliphatic and aromatic substrates while handling sensitive groups. Conditions remain mild (0–50°C overall), but require careful handling of the activation step. These catalytic hydrosilylation methods excel in avoiding over-reduction to amines, a common issue in noncatalytic approaches, due to the controlled delivery of equivalents and stabilization of intermediates. However, challenges include the relatively high cost of catalysts and the need to manage byproducts from consumption. Ongoing refinements focus on earth-abundant metals to improve . The general equation for hydrosilylation-based reduction is: \mathrm{RC(O)NR'_2 + HSiR''_3 \xrightarrow{\text{catalyst}} RCH(OSiR''_3)NHR'_2 \xrightarrow{\mathrm{H_2O}} RCHO + HNR'_2 + (R''_3Si)_2O This pathway underscores the efficiency of catalytic systems in achieving partial reduction.

Recent Developments

Transition Metal Innovations

Recent innovations in transition metal catalysis have expanded the scope of amide reductions, particularly for challenging unactivated substrates, by leveraging copper(I) systems for site-selective hydrogenation. A bifunctional catalyst combining Cu(I) with an N-heterocyclic carbene and a guanidine organocatalyst enables the first Cu(I)-catalyzed hydrogenation of amides using molecular hydrogen, targeting difficult-to-reduce unactivated alkyl and heterocycle-derived amides with high chemoselectivity via hydrogen-bonding recognition. The reaction proceeds under mild conditions of 70°C and 100 bar H₂ in 1,4-dioxane, employing 10–20 mol% catalyst loading and 1.3–2.5 equiv NaOᵗBu as base, affording amines in good yields while tolerating sensitive functional groups. The transformation follows the general scheme for deoxygenative reduction: \text{RC(O)NHR'} + \text{H}_2 \xrightarrow{\text{Cu(I) cat., base}} \text{RCH}_2\text{NHR'} This approach demonstrates improved temperature control compared to classical catalytic hydrogenation, which typically demands >150°C and pressures exceeding 100 bar for unactivated amides. Nickel and cobalt catalysts have also seen post-2022 progress in reductive transformations involving amides, often integrating reduction steps into broader synthetic sequences. Nickel catalysis facilitates ester-to-amide conversion followed by reduction, as in a 2023 Ni-based reductive amidation of esters with nitroarenes under hydrogen, yielding amides that undergo subsequent Ni-promoted deoxygenation to amines with high efficiency and broad substrate tolerance. Complementing this, a 2024 cobalt carbonyl system enables light-promoted hydroaminocarbonylation of alkenes with amines at low pressure (1–5 bar CO/H₂) and room temperature, producing amides that can be directly reduced in tandem to amines using silane additives, streamlining access to complex amine scaffolds. Zirconium-based has advanced the of tertiary amides to aldehydes through a 2023 catalytic protocol using 5 mol% Cp₂ZrCl₂ to generate zirconocene hydride intermediates. This mild, divergent semireduction converts tertiary amides to imines at in , followed by acidic to afford aldehydes in yields up to 94%, with exceptional over esters, nitriles, and other reducible groups. The method outperforms stoichiometric (Cp₂Zr(H)Cl) by enabling turnover and avoiding over-reduction. These innovations have found application in pharmaceutical , such as preparing intermediates for - and morpholine-based candidates by site-selective Cu(I) of polyfunctional amides, enhancing step economy in late-stage diversification. Overall, they reduce reliance on harsh conditions, with pressures as low as 1 in Co systems and temperatures below 100°C, contrasting classical methods and improving in and production.

Metal-Free and Sustainable Strategies

Electrochemical methods represent a cornerstone of metal-free amide reduction, leveraging as a clean reductant to generate species , often in protic solvents like or alcohols for enhanced . These approaches typically involve cathodic reduction, where electrons facilitate of the carbonyl to form amines, bypassing the need for hazardous reducing agents or high-pressure gas. A seminal example is the hydrosilane-mediated electrochemical reduction reported in 2021, which uses phenylsilane as a hydrogen donor on a carbon , achieving up to 99% yield for the conversion of amides to amines under mild conditions (, undivided cell). The process proceeds via a ketyl , enabling selective C-O cleavage. The general reaction can be represented as: \text{RC(O)NR}_2 + 2e^- + 2\text{H}^+ \rightarrow \text{RCH}_2\text{NR}_2 + \text{H}_2\text{O} This equation illustrates the two-electron, two-proton transfer essential for deoxygenative reduction. Reviews from 2018 to 2024 underscore the versatility of such electrosyntheses for carbonyl compounds, including amides, with applications in pharmaceutical synthesis due to their atom economy and compatibility with renewable energy sources. Advantages include minimal waste generation and scalability, as demonstrated by continuous-flow setups that enhance efficiency while using earth-abundant electrodes like graphite or lead. Organocatalytic strategies further advance sustainable amide reduction by employing non-metallic Lewis acids to activate the amide for hydrosilylation with inexpensive silanes, yielding amines without toxic byproducts. Boron-based catalysts, such as tris(pentafluorophenyl)borane [B(C₆F₅)₃], have emerged as highly effective, promoting the reduction of secondary and tertiary amides under mild conditions (e.g., 60–80°C, toluene or solvent-free). A 2008 seminal study established the chemoselectivity of this approach for tertiary amides, achieving >90% yields while sparing esters and nitro groups. More recent advancements, detailed in a 2023 review, highlight extensions to primary amides using polymethylhydrosiloxane (PMHS) or tetramethyldisiloxane (TMDS) as green, air-stable reductants, with catalyst loadings as low as 5 mol% and turnover numbers exceeding 100. For instance, B(C₆F₅)₃ facilitates the activation of the amide nitrogen, enabling nucleophilic attack by silane to form silylated intermediates that hydrolyze to amines. Calcium(II) triflimide [Ca(NTf₂)₂], an earth-abundant organocatalyst, has been explored in 2023 for related hydrosilylation activations of imine intermediates, offering high functional group tolerance. These methods prioritize sustainability through recyclable catalysts and avoidance of precious metals, with E-factors often below 5, making them suitable for industrial-scale amine production. Biocatalytic reductions of amides remain an emerging frontier, adapting enzymes like reductases to handle amide-derived substrates in aqueous media for enantioselective synthesis. A report demonstrated engineered reductive aminases (RedAms) for the stereoselective reduction of cyclic s in processes, achieving >95% ee in systems at ambient and 7–9. This approach leverages cofactor recycling (e.g., NADPH via glucose dehydrogenase) to minimize waste, with conversions up to 80% for secondary products. While direct amide reduction is limited, these adaptations fill a gap in by enabling mild, metal-free conditions compatible with bio-based feedstocks. Green chemistry principles are integrated across these strategies through the use of water or bio-derived solvents, such as . Overall, these metal-free methods offer significant advantages over conventional techniques, including reduced toxicity, lower environmental impact, and improved selectivity, addressing key challenges in amide reduction.

References

  1. [1]
    21.7 Chemistry of Amides - OpenStax
    Sep 20, 2023 · Amide reduction occurs by nucleophilic addition of hydride ion to the amide carbonyl group, followed by expulsion of the oxygen atom as an ...<|control11|><|separator|>
  2. [2]
    Amine synthesis by amide reduction - Organic Chemistry Portal
    The combination of triethylborane with an alkali metal base catalyzes the reduction of amides with silanes to form amines under mild conditions. In addition, a ...
  3. [3]
    Reduction of Amides to Amines and Aldehydes - Chemistry Steps
    Amides can be reduced to amines using LiAlH4. Tertiary amides can be reduced to tertiary amines and aldehydes with 9-BBN and Sia2BH respectively.
  4. [4]
    Reduction of carboxyl compounds to aldehydes
    A highly efficient in situ generation of the Schwartz reagent provides a convenient method for the reduction of amides to aldehydes and the regioselective ...
  5. [5]
    Catalytic Hydrogenation of Carboxylic Acid Esters, Amides, and ...
    This review describes the catalytic reduction of amides, carboxylic acid esters and nitriles with homogeneous catalysts using molecular hydrogen as an ...<|control11|><|separator|>
  6. [6]
    Amide Reduction - Wordpress
    Oct 30, 2025 · Most imides can be reduced to amines with the reagents used for amide reduction. The greenest methods available have lowest utility – see later.
  7. [7]
    Highly Chemoselective Reduction of Amides (Primary, Secondary ...
    Jan 24, 2014 · Highly chemoselective direct reduction of primary, secondary, and tertiary amides to alcohols using SmI2/amine/H2O is reported.
  8. [8]
    Amide Activation in Ground and Excited States - PMC - NIH
    The chemical and biochemical stability and reactivity is determined by the contribution of the minor resonance structure (B in Figure 1), in which formally a ...
  9. [9]
    Mild and Selective Hydrozirconation of Amides to Aldehydes Using ...
    The mild reduction of tertiary amides to aldehydes was efficiently carried out on both aromatic and alkyl substrates with the Schwartz reagent as shown in ...
  10. [10]
    Reductive Functionalization of Amides in Synthesis and ... - Frontiers
    Apr 25, 2021 · In this review, applications of the amide reductive functionalization methodology for the synthesis and modification of bioactive compounds are covered.<|control11|><|separator|>
  11. [11]
    Chemoselective reduction of carboxamides - RSC Publishing
    Oct 24, 2016 · The reduction of amides gives access to a wide variety of important compounds such as amines, imines, enamines, nitriles, aldehydes and ...
  12. [12]
    Homogeneous and heterogeneous catalytic reduction of amides ...
    Aug 4, 2020 · This review describes the most recent advances in the area of amide hydrogenation using H 2 exclusively and molecularly defined homogeneous as well as nano- ...<|control11|><|separator|>
  13. [13]
    Hydrogenation of amides - US2143751A - Google Patents
    The hydrogenating metalsreferred to, comprise copper, zinc, cadmium and silver. The acid-forming metals include chromium, vanadium and molybdenum. It has been ...
  14. [14]
    Reduction of Organic Compounds by Lithium Aluminum Hydride. I ...
    A safety guidance document for lithium aluminum hydride (LAH) reduction: a resource for developing specific SOPs on LAH manipulations.
  15. [15]
    Amide Reduction Mechanism by LiAlH4 - Chemistry Steps
    Amide reduction by LiAlH4 involves protonation, hydride addition, C=O oxygen elimination, and C=N bond reduction, resulting in amines.
  16. [16]
    Convenient methods for the reduction of amides, nitriles, carboxylic ...
    Reaction of amides with NaBH4-I2 system in THF gives the corresponding amines in 70–76% yields. Reduction of nitriles yields the corresponding amines in ...
  17. [17]
    A Convenient and General Iron‐Catalyzed Reduction of Amides to ...
    Nov 30, 2009 · While the iron is hot: The first general and efficient iron-catalyzed reduction of secondary and tertiary amides into amines using ...
  18. [18]
    Base Metal Catalysts for Deoxygenative Reduction of Amides to ...
    Base metal catalysts (Mn, Fe, Co, Ni) are used for deoxygenative reduction of amides to amines via hydrogenation, hydrosilylation, and hydroboration, though ...
  19. [19]
    [PDF] Catalytic Amide Reductions under Hydrosilylation Conditions
    Apr 27, 2016 · This thesis covers the development of catalytic methodologies for the mild and chemoselective reductions of amides.
  20. [20]
    A Focused Review of Synthetic Applications of Lawesson's Reagent ...
    Nov 17, 2021 · The conversion of oxygen of amides to its thio analogs is the basis of thioamide formations. ... RANEY® nickel-induced hydrodesulfurization ...Missing: desulfurization | Show results with:desulfurization
  21. [21]
    Homogeneous cobalt-catalyzed deoxygenative hydrogenation of ...
    The first general and efficient cobalt-catalyzed deoxygenative hydrogenation of amides to amines is presented.
  22. [22]
    Controlled reduction of activated primary and secondary amides into ...
    A practical method uses diisobutylaluminum hydride (DIBAL-H) in toluene to reduce activated amides into aldehydes, with good to excellent yields.Missing: original | Show results with:original
  23. [23]
    DIBALH: from known fundamental to an unusual reaction
    Oct 18, 2021 · This study presents a quick and reliable approach to the chemoselective partial reduction of tertiary amides to aldehydes in the presence of readily reducible ...Missing: original | Show results with:original
  24. [24]
    DIBAL-H, Diisobutylaluminium hydride - Organic Chemistry Portal
    Kim, H. Lee, H. Lee, Org. Lett., 2018, 20, 5478-5481. Esters and Weinreb amides undergo reduction to the corresponding aldehydes using DIBAL-H followed by ...Missing: first report reference
  25. [25]
    First asymmetric synthesis of both enantiomers of Tropional® and ...
    Subsequent reduction of the resulting nitriles with diisobutyl aluminium hydride (DIBAL-H) led to the desired aldehydes in good overall yields (52–53%) and ...
  26. [26]
    Sonn‐Müller Reaction - Major Reference Works - Wiley Online Library
    Sep 15, 2010 · The transformation of aromatic anilides into aromatic aldehydes involves multistep process and is generally referred to as the Sonn–Müller reaction.Missing: original | Show results with:original
  27. [27]
    Selective Reductions. I. The Partial Reduction of Tertiary Amides ...
    Reduction of Weinreb amides to aldehydes under ambient conditions with magnesium borohydride reagents. Tetrahedron Letters 2015, 56 (5) , 706-709. https ...Missing: Sonn- method
  28. [28]
    chemoselective partial reduction of tertiary amides in the presence ...
    This study presents a quick and reliable approach to the chemoselective partial reduction of tertiary amides to aldehydes in the presence of readily reducible ...
  29. [29]
    Unexpected selectivity in ruthenium-catalyzed hydrosilylation of ...
    Selective ruthenium-catalyzed reductive coupling of primary amides under hydrosilylation conditions is achieved using an one pot procedure.Missing: aldehydes | Show results with:aldehydes
  30. [30]
    Site-Selective Copper(I)-Catalyzed Hydrogenation of Amides
    Jan 3, 2025 · A site-selective reduction of amides poses a formidable challenge for method development, as generally highly nucleophilic hydrides (such as ...
  31. [31]
    Streamlining the synthesis of amides using Nickel-based ... - Nature
    Aug 17, 2023 · Here, we present a robust methodology for amide synthesis compared to traditional amidation reactions: the reductive amidation of esters with ...
  32. [32]
    Cobalt-catalyzed synthesis of amides from alkenes and amines ...
    Jan 4, 2024 · We report an alkene hydroaminocarbonylation catalyzed by unmodified, inexpensive cobalt carbonyl under mild conditions and low pressure promoted by light.
  33. [33]
    Mild Divergent Semireductive Transformations of Secondary and Tertiary Amides via Zirconocene Hydride Catalysis
    ### Abstract and Key Points on Cp2Zr(H)Cl for Tertiary Amides to Aldehydes (2023)
  34. [34]
    Hydrosilane-Mediated Electrochemical Reduction of Amides
    Jun 21, 2021 · Electrochemical reduction of amides was achieved by using a hydrosilane without any toxic or expensive metals.Missing: photoelectrochemical | Show results with:photoelectrochemical
  35. [35]
    Electrosynthesis of amides: Achievements since 2018 and prospects
    Jun 15, 2024 · Moreover, amides are vital intermediates for the synthesis of other types of molecules, including amines, isonitriles, and N-heterocycles.
  36. [36]
    Highly Chemoselective Metal-Free Reduction of Tertiary Amides
    This communication describes the chemoselective metal-free reduction of tertiary amides to the corresponding amines.
  37. [37]
    Metal‐Free Catalytic Reduction of Amides: Recent Progress - 2023
    Mar 27, 2023 · This review covers recent advances in the catalytic reduction of amides to amines without the need for metal catalysts, thus overcoming such problems.Missing: organocatalytic | Show results with:organocatalytic
  38. [38]
    Catalyzed Reductive Amination of Biomass-Derived Keto Acids to ...
    Mar 31, 2023 · The two versatile reactions were triggered by an earth-abundant calcium-based catalyst using hydrosilane as a reducing reagent under solvent- ...
  39. [39]
    Engineered Biocatalysts for Enantioselective Reductive Aminations ...
    Mar 22, 2023 · Here we show that the scope of reductive aminases (RedAms) can be extended to allow selective reductive aminations of cyclic secondary amines.Missing: amide | Show results with:amide
  40. [40]
    Deoxygenative photochemical alkylation of secondary amides ...
    Jan 22, 2025 · Our method leverages triflic anhydride-mediated semi-reduction of amides to imines, followed by a photochemical radical alkylation step.