Fact-checked by Grok 2 weeks ago

Nucleophilic addition

Nucleophilic addition refers to a fundamental class of chemical reactions in which a , an electron-rich , attacks and forms a bond with the electrophilic carbon atom of a multiple bond, most commonly the (C=O) in aldehydes and ketones, resulting in the formation of a tetrahedral without the departure of a . This process typically proceeds in two main steps: the adds to the carbon, rehybridizing it from sp² to sp³ and generating an , followed by of the oxygen to yield a stable product. can be anionic, such as (H⁻) or (CN⁻), or neutral, like or amines, with the latter often requiring or to facilitate the reaction. The reactivity of carbonyl compounds in nucleophilic addition stems from the of the C=O bond, where the electronegative oxygen withdraws , rendering the carbon partially positive and thus susceptible to nucleophilic attack. Aldehydes generally undergo these reactions more readily than ketones due to less steric hindrance from a single alkyl substituent on the carbonyl carbon compared to two in ketones, as well as greater of the C=O bond in aldehydes. The typically approaches the carbonyl at an angle of approximately 105° from the plane of the C=O bond, opposite the oxygen lone pairs, which influences the of the product; for instance, addition to a prochiral carbonyl can generate a new chiral center, often resulting in racemic mixtures unless controlled conditions are used. acids, such as metal cations, can coordinate to the carbonyl oxygen to enhance electrophilicity and accelerate the reaction rate. These reactions are central to synthetic , enabling the construction of complex carbon skeletons from simple carbonyl precursors, and they underpin natural processes such as , where nucleophilic addition occurs in enzymatic transformations of aldehydes. Notable examples include the , in which organomagnesium reagents add to carbonyls to form alcohols, and the cyanohydrin formation via addition, which introduces a hydroxyl and functionality useful for further synthesis. Variations extend beyond simple carbonyls to include conjugate additions (1,4-additions) to α,β-unsaturated carbonyls, where the targets the β-carbon, leading to intermediates that can be protonated to saturated products. Overall, nucleophilic additions highlight the versatility of carbonyl compounds as electrophiles in building blocks for pharmaceuticals, materials, and biochemical pathways.

Fundamentals

Definition and scope

A nucleophilic addition is an in which a , an electron-rich , donates a pair of electrons to form a new with an electrophilic center, typically the electron-deficient atom in a multiple bond such as a π-bond, resulting in the conversion of the multiple bond to a and often the formation of a tetrahedral . This is a cornerstone of , enabling the construction of carbon-carbon and carbon-heteroatom bonds without the displacement of a . The scope of nucleophilic addition encompasses reactions at polarized multiple bonds, including the carbon-oxygen double bond (C=O) in carbonyl compounds, the carbon-nitrogen double bond (C=N) in imines, the carbon-nitrogen triple bond (C≡N) in nitriles, and activated carbon-carbon multiple bonds (C=C or C≡C) conjugated with electron-withdrawing groups. Unlike nucleophilic substitution reactions, where the nucleophile replaces a leaving group attached to the electrophilic center, addition reactions preserve the original substituents and simply incorporate the nucleophile across the multiple bond. Key characteristics of nucleophilic additions include their frequent irreversibility under conditions for simple cases, such as those involving strong nucleophiles like organometallics, due to the poor leaving group ability of the resulting or similar species. These reactions are particularly prevalent in carbonyl chemistry, where they facilitate the synthesis of alcohols, amines, and extended carbon chains essential for and pharmaceutical development. Historically, the reaction type was first exemplified in the through the addition to aldehydes, forming cyanohydrins, as reported by Winkler in 1832.

Nucleophiles and electrophiles

Nucleophiles are electron-rich that donate a pair of electrons to form a with an during nucleophilic addition reactions. These typically feature lone pairs on heteroatoms or excess in π-bonds, allowing them to seek out electron-deficient centers. They are classified into anionic nucleophiles, such as the ion (\ce{CN^-}) and ion (\ce{H^-}); neutral nucleophiles, exemplified by secondary amines (\ce{R2NH}) and alcohols (\ce{ROH}); and organometallic nucleophiles, including Grignard reagents (\ce{RMgBr}). Electrophiles, in contrast, are electron-deficient species that accept electron pairs from nucleophiles, with the reactive site often being a carbon atom in polarized multiple bonds. In carbonyl compounds (\ce{C=O}), the carbon atom carries a partial positive charge (\delta^+) owing to oxygen's higher , rendering the bond polarized and the carbon highly susceptible to nucleophilic attack. Carbon-carbon double bonds (\ce{C=C}) can also serve as electrophiles when activated by electron-withdrawing groups (EWGs), such as a conjugated carbonyl, which depletes from the β-carbon and enhances its electrophilicity. Nucleophilicity—the tendency of a to donate electrons—is influenced by basicity, where for structurally similar nucleophiles, stronger Brønsted bases tend to be more nucleophilic; , favoring softer, more diffuse electron clouds in larger atoms or anions; and . In protic solvents (e.g., ), effects reduce the nucleophilicity of small, strongly basic anions more than larger, weaker bases, leading to trends opposite to basicity (e.g., \ce{I^- > Br^- > Cl^- > F^-}). aprotic solvents (e.g., DMSO) boost anionic nucleophilicity by minimizing ion . Electrophilicity, meanwhile, depends on bond , with stronger increasing the partial positive charge on the electrophilic carbon; steric hindrance, which impedes nucleophile approach in crowded environments; and the stabilizing effect of EWGs, which further enhance . Common nucleophile-electrophile pairings illustrate these principles: the (\ce{H^-}) from (\ce{NaBH4}) adds to the electrophilic carbonyl carbon of ketones, yielding secondary alcohols after ; similarly, \ce{CN^-} attacks the polarized \ce{C=O} of aldehydes to form cyanohydrins, useful in .

Reaction mechanisms

General mechanism

Nucleophilic addition reactions involve the attack of a on an electrophilic multiple bond, typically a carbon- double bond such as C=O or C=NR, leading to the formation of a new carbon-nucleophile σ-bond. The process occurs in a stepwise manner, with the electrophilic carbon serving as the primary site of reactivity due to its partial positive charge from the polarization of the π-bond. In the initial step, the nucleophile approaches and bonds to this carbon, simultaneously breaking the π-component of the multiple bond and generating a tetrahedral (or analogous trigonal pyramidal) intermediate. This intermediate features the original substituents on the carbon, the added nucleophile, and a negatively charged heteroatom (e.g., O⁻ or N⁻R), which stabilizes the structure through or charge delocalization. The general reaction pathway can be depicted as follows: \mathrm{R_2C=XR' + Nu^- \rightarrow R_2C(Nu)-X^-R' \quad (tetrahedral \ intermediate)} \mathrm{R_2C(Nu)-X^-R' + H^+ \rightarrow R_2C(Nu)-XHR'} where X represents a like or , and R' may be H or an . The second step involves neutralization of the anionic , commonly via of the to yield the addition product. For instance, in the case of oxygen-based substrates, the (O⁻) abstracts a proton to form a hydroxyl group. This protonation can occur from solvent, a catalyst, or an external acid source, ensuring the reaction proceeds to completion under neutral or basic conditions. The geometry of the nucleophilic approach is governed by the Bürgi-Dunitz trajectory, in which the advances toward the electrophilic carbon at an of approximately 107° relative to the C–X bond axis. This oblique path maximizes orbital overlap between the nucleophile's highest occupied (HOMO) and the electrophile's lowest unoccupied (LUMO), facilitating efficient bond formation while minimizing steric repulsion. Experimental and computational studies of crystal structures and transition states have consistently validated this across various nucleophilic additions to carbonyls and analogous systems. Thermodynamically, the step involves breaking the π-bond and forming a new σ-bond; the overall process can be favorable depending on the and subsequent , though some (e.g., ) are reversible with favoring the carbonyl. The process contrasts with reactions by retaining the in the product, emphasizing across the multiple .

Acid- versus base-catalyzed mechanisms

In base-catalyzed nucleophilic addition, the nucleophile directly attacks the neutral electrophile, such as a carbonyl carbon, forming an anionic tetrahedral intermediate, often an alkoxide in the case of carbon-oxygen double bonds. This mechanism is favored when strong, anionic nucleophiles are employed, such as organometallics or hydroxide ions, as the basic conditions enhance the nucleophile's reactivity without altering the electrophile's inherent polarity. The intermediate then protonates in a subsequent step to yield the neutral product, typically requiring an acidic workup if no internal proton source is available. In contrast, acid-catalyzed nucleophilic addition begins with of the on the , such as the oxygen in a , forming a resonance-stabilized cationic species like R₂C=OH⁺ that significantly increases the 's reactivity toward even weak, nucleophiles like or alcohols. The then adds to this activated , generating a tetrahedral intermediate, such as R₂C(Nu)-OH₂⁺, which undergoes to afford the product R₂C(Nu)-OH and regenerate the acid catalyst.
R₂C=OH⁺ + Nu → R₂C(Nu)-OH₂⁺ → R₂C(Nu)-OH + H⁺
This stepwise process involves equilibrium , making the overall reaction reversible and pH-dependent, with optimal rates at moderate acidity to balance activation and availability. Key differences between the two mechanisms include the charge of the intermediates— anionic in base catalysis versus neutral or cationic in —and the types of nucleophiles suitable, with base conditions suiting strong anions and acid conditions enabling weaker neutrals. enhances the reaction rate by several orders of magnitude for additions to carbonyls, by increasing polarity, but it risks side reactions such as enolization in substrates with alpha hydrogens, potentially leading to competing pathways like .

Addition to carbon-oxygen double bonds

Aldehydes and ketones

Nucleophilic addition reactions to aldehydes and ketones primarily target the electrophilic carbonyl carbon, forming tetrahedral intermediates that lead to alcohols or other functionalized products. Aldehydes exhibit greater reactivity than ketones toward nucleophiles due to reduced steric hindrance from the smaller substituent compared to the bulkier alkyl groups in ketones, as well as slightly greater electrophilicity of the carbonyl carbon in aldehydes arising from the electron-withdrawing of . This difference influences reaction rates and selectivity, with aldehydes often reacting faster under milder conditions. A prominent example is the addition of (RMgX), which involves nucleophilic attack by the carbon nucleophile of the Grignard reagent on the carbonyl carbon, yielding an alkoxide intermediate that is hydrolyzed during to produce secondary alcohols from or tertiary alcohols from ketones. The general reaction for an aldehyde is depicted as: \mathrm{RCHO + R'MgX \rightarrow RCH(OMgX)R' \xrightarrow{H_3O^+} RCH(OH)R'} This transformation is highly versatile for carbon-carbon bond formation, with the alkoxide preventing over-addition by deactivating the intermediate toward further nucleophilic attack. However, the reaction typically yields racemic mixtures at the new chiral center unless chiral auxiliaries or catalysts are employed for stereocontrol. Hydride reductions using (NaBH₄) or lithium aluminum hydride (LiAlH₄) deliver H⁻ as the , converting aldehydes to primary alcohols and ketones to secondary alcohols through a that can involve either acid- or base-catalyzed pathways depending on solvent and conditions. is milder and selective for aldehydes and ketones at or below, while LiAlH₄ is more reactive and requires conditions but achieves similar outcomes. These reductions preserve at existing chiral centers but produce racemic products at the carbinol carbon in achiral substrates. Cyanohydrin formation involves addition of (HCN), where CN⁻ acts as the under basic conditions, adding to the carbonyl to yield β-hydroxy nitriles (RCH(OH)CN from aldehydes or RC(OH)(CN)R' from ketones). This reaction is reversible and equilibrium-driven, often requiring a catalyst like KCN to generate CN⁻ , and is particularly useful for aldehydes due to their higher reactivity. The process highlights the carbonyl's susceptibility to soft nucleophiles like , with the product serving as a synthetic for further transformations.

Carboxylic acid derivatives

Carboxylic acid derivatives, including acid chlorides, anhydrides, esters, and amides, undergo nucleophilic addition reactions at the carbonyl carbon, but these processes typically proceed via an addition-elimination rather than forming stable addition products. Unlike aldehydes and ketones, where the tetrahedral is relatively stable, the presence of a good in these derivatives (such as Cl⁻ in acid chlorides or OR⁻ in esters) facilitates the expulsion of that group, leading to . This reactivity makes true nucleophilic addition rare without specific stabilization of the , as the elimination step regenerates the carbonyl. The reactivity of these derivatives toward nucleophiles decreases in the order acid chlorides > anhydrides > esters > amides, influenced by both the quality of the and effects. Acid chlorides are the most reactive due to the excellent leaving ability of and minimal stabilization of the carbonyl. Anhydrides follow closely, with as the leaving group. Esters are less reactive because is a poorer leaving group, while amides are the least reactive owing to donation from the , which delocalizes the carbonyl π electrons and reduces electrophilicity. This order reflects slower addition rates compared to simple carbonyls, where no leaving group is available to drive elimination. The general mechanism begins with nucleophilic attack on the carbonyl carbon, forming a tetrahedral , followed by elimination of the to restore planarity. For example, in the reaction of an acid chloride with a , the initial -elimination yields a : \mathrm{RC(O)Cl + R'MgBr \rightarrow RC(O)R' + MgBrCl} However, the resulting can undergo further unless conditions are controlled, such as using organocadmium or organocopper reagents to selectively form the . With esters, s typically add twice—first displacing the alkoxide to form a intermediate, then adding again to produce a tertiary alcohol after —highlighting the competition between and . Amides, due to their low reactivity, primarily undergo under acidic or basic conditions, where the step is rate-determining but still leads to products like carboxylic acids and amines.

Addition to carbon-nitrogen multiple bonds

Imines and enamines

Imines are compounds characterized by a carbon-nitrogen , with the general R₂C=NR', where R and R' represent alkyl, aryl, or other groups. They are typically synthesized through the acid-catalyzed of aldehydes or ketones with primary , a process that involves nucleophilic addition of the to the carbonyl carbon, followed by to eliminate . This is reversible and often requires removal of to drive imine formation forward. Compared to carbonyl compounds, imines exhibit reduced electrophilicity at the C=N bond due to the lower of relative to oxygen, which results in less effective polarization of the and diminished attraction for nucleophiles. As a result, nucleophilic additions to imines generally require activation, such as to form more electrophilic ions, or the use of acids. This contrasts with the more straightforward reactivity of carbonyls in nucleophilic addition mechanisms. Nucleophilic additions to s commonly involve donors or organometallic reagents, leading to products after . For instance, reduction using (NaBH₄) adds H⁻ to the imine carbon, generating a resonance-stabilized anion that protonates to yield a secondary or tertiary , depending on the substituents. The proceeds as follows: \ce{R2C=NR' + H- -> R2CH-NR'^- ->[H+] R2CH-NHR'} This reduction is particularly efficient for imines derived from aldehydes. Similarly, organometallic reagents like Grignard (RMgX) or organolithium (RLi) compounds add their organic group to the imine carbon, forming amines upon aqueous ; these reactions often benefit from Lewis acid coordination to enhance imine reactivity. When secondary amines react with aldehydes or ketones possessing α-hydrogens, the initial carbinolamine intermediate cannot eliminate water to form an but instead undergoes dehydration and tautomerization to produce an , R₂N-CR=CR₂. Enamines function as equivalents, with the β-carbon acting as a in subsequent reactions, though their formation itself stems from nucleophilic addition to the carbonyl. In the Stork enamine reaction, these enamines alkylate at the α-position of the original carbonyl upon , enabling selective C-C formation. The mechanism of nucleophilic addition to imines parallels that of carbonyl additions, featuring nucleophilic attack at the electrophilic carbon to form an anionic intermediate, followed by proton transfer. However, the intermediate anion on protonates more rapidly than the oxygen analog, stabilizing the . Imines often display (cis-trans) isomerism due to restricted rotation around the C=N bond, which can influence the stereochemical outcome of additions, particularly in asymmetric catalysis. A key application of imine reactivity is in , where the imine intermediate formed from a carbonyl and is directly reduced to produce , avoiding isolation of the often unstable imine; this method is widely used in pharmaceutical for its efficiency and selectivity.

Nitriles

Nitriles, represented by the formula R–C≡N, exhibit reactivity toward addition primarily at the electrophilic carbon atom of the , driven by the electron-withdrawing . This addition typically forms imine-like intermediates, often referred to as iminates, where the nucleophile bonds to the carbon while the gains a proton or . Unlike additions to carbon-oxygen or carbon-nitrogen double bonds, the in nitriles necessitates two successive nucleophilic additions for complete saturation, rendering these reactions generally slower and requiring harsher conditions. The general mechanism begins with the attacking the carbon, yielding an intermediate of the form R–C(Nu)=NH, which may tautomerize to an or undergo further transformation depending on conditions. In acid- or base-catalyzed processes, this intermediate resembles those seen in chemistry but involves an initial step across the . These intermediates can be hydrolyzed to derivatives, highlighting the versatility of additions in synthetic pathways. A primary application is the of to or , proceeding via nucleophilic addition of under acidic or conditions. The overall process can be summarized as: \ce{RCN + H2O ->[H+ or OH-] RCONH2} Further of the yields the , with high yields (e.g., 92–95% under acidic conditions with ). Another key reaction involves Grignard reagents, where the organomagnesium species adds to the to form a after : \text{RCN} + \text{R'MgBr} \rightarrow \text{RC(}= \text{NMgBr)}\text{R'} \xrightarrow{\text{H}_3\text{O}^+} \text{RCOR'} This proceeds through an imine magnesium salt intermediate, providing a route to ketones from nitriles. Additions of cyanide to nitriles are rare due to unfavorable thermodynamics without metal coordination.

Addition to carbon-carbon multiple bonds

Isolated alkenes and alkynes

Nucleophilic additions to isolated alkenes and alkynes are rare because the carbon-carbon multiple bonds in unactivated systems are non-polar and exhibit low electrophilicity, necessitating strained structures or exceptionally strong nucleophiles to proceed. Such direct additions to simple unactivated alkenes are essentially unknown in standard and are limited to highly specialized cases. Unlike polarized bonds such as C=O, these reactions lack a strong driving force for nucleophilic attack without activation. A prominent example for alkenes involves strained systems like fullerene, where the Bingel reaction enables through nucleophilic addition. In this process, diethyl bromomalonate is deprotonated by a base such as to generate a stabilized α-halocarbanion, which adds to a [6,6]- of the fullerene. The resulting fullerene anion then undergoes intramolecular displacement of the bromide, closing to form a methanofullerene with yields up to 40%. This reaction highlights how curvature-induced strain in fullerenes enhances the electrophilicity of the C=C bonds. For alkynes, the sp-hybridized carbons confer greater electrophilicity than in alkenes, allowing additions with strong nucleophiles under basic conditions. Terminal alkynes, such as , undergo reaction with alkoxides at elevated temperatures (e.g., 150 °C) to form vinyl ethers, or with to yield acrylonitriles, proceeding via nucleophilic attack at the terminal carbon followed by of the vinyl anion . However, treatment of terminal alkynes with even stronger bases like favors (pKa ≈ 25) over addition, as the acetylide anion forms preferentially. These transformations remain largely confined to academic research due to their demanding conditions and from alternative pathways, with minimal industrial relevance relative to additions involving conjugated or activated multiple bonds.

Conjugated systems

In conjugated systems, such as α,β-unsaturated carbonyl compounds, can occur either directly at the carbonyl carbon (1,2-addition) or at the β-carbon (1,4- or conjugate addition), with the latter enabled by the delocalization of electrons across the conjugated π-system. The conjugate addition, often termed the Michael addition when involving nucleophiles, proceeds via attack of the at the β-carbon, generating an intermediate at the carbonyl oxygen; subsequent at the α-carbon yields the saturated 1,4-adduct. This arises from the partial positive charge at the β-carbon due to conjugation, making it an electrophilic site for nucleophilic attack. The competition between 1,2- and 1,4-addition is governed by kinetic and thermodynamic factors, as well as the nature of the . The 1,2-addition is typically kinetically favored, occurring rapidly at the more electrophilic carbonyl carbon, particularly with hard nucleophiles like organolithium or Grignard reagents under low-temperature conditions. In contrast, the 1,4-addition is thermodynamically preferred due to the stability of the resulting and , and it predominates with soft nucleophiles or under equilibrating conditions such as higher temperatures or protic solvents. According to hard-soft acid-base (HSAB) principles, hard s prefer the hard carbonyl carbon (1,2-pathway), while soft nucleophiles target the softer β-carbon (1,4-pathway). Representative examples include the use of organocopper reagents, such as Gilman reagents (R₂CuLi), which selectively promote 1,4-addition to enones, delivering the R group to the β-position with high efficiency and minimal 1,2-product formation. Another key case is the Michael addition of enolates (often from β-dicarbonyl compounds) to α,β-unsaturated carbonyl acceptors, forming new C-C bonds at the β-position and enabling subsequent cyclizations like the . The general mechanism for conjugate addition can be represented as: \mathrm{RCH=CH-C(=O)R' + Nu^- \rightarrow RCH(Nu)-CH_2-C(^-O^-)R' \rightarrow RCH(Nu)-CH_2-C(=O)R'} where the enolate intermediate is protonated to afford the 1,4-adduct. Stereochemistry in conjugate additions is often syn, with the nucleophile and the ensuing enolate geometry aligning on the same face of the original double bond, potentially generating chiral centers at the α- and β-positions. Asymmetric variants achieve high enantioselectivity through chiral auxiliaries, ligands, or catalysts, enabling stereocontrol in the formation of these centers.

References

  1. [1]
  2. [2]
    19.4 Nucleophilic Addition Reactions of Aldehydes and Ketones
    Sep 20, 2023 · A nucleophilic addition reaction to an aldehyde or ketone. The nucleophile approaches the carbonyl group from an angle of approximately 75° to ...Missing: definition | Show results with:definition
  3. [3]
    Nucleophilic Addition To Carbonyls - Master Organic Chemistry
    Sep 9, 2022 · This reaction is called, “nucleophilic addition”, or sometimes, “1,2-addition”. In this reaction a C-Nu bond is formed and the CO pi bond breaks.
  4. [4]
    Nucleophilic Additions to Aldehydes, Ketones, Imines, and Nitriles
    Jun 15, 2022 · The rate-determining step of the reaction is the nucleophilic addition of the nitrogen nucleophile to the protonated carbonyl. ... Examples of the ...
  5. [5]
    Alpha-Hydroxynitrile - an overview | ScienceDirect Topics
    The addition of cyanide to carbonyl compounds giving α-hydroxynitriles (cyanohydrins) was discovered as early as 1832 when Winkler reacted HCN with benzaldehyde ...
  6. [6]
    Amine Reactivity
    ### Summary of Carbonyl Compounds from https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/chapt16.htm
  7. [7]
  8. [8]
    What defines electrophilicity in carbonyl compounds - PMC - NIH
    We find that this electrophilicity is mainly determined by the electrostatic attractions between the carbonyl compound and the nucleophile (cyanide)
  9. [9]
    8.3. Factors affecting rate of nucleophilic substitution reactions
    Nucleophile strength · Charge – negatively charged => stronger nucleophile · Within a row – more electronegative atom => weaker nucleophile · Within a column, size ...
  10. [10]
    Chapter 7: Nucleophilic attack at the carbonyl carbon: – OCLUE
    All of these reactions are the result of a nucleophilic addition to the carbonyl group, during the course of which the carbonyl carbon rehybridizes from sp2 to ...Missing: definition | Show results with:definition
  11. [11]
    Geometrical reaction coordinates. II. Nucleophilic addition to a ...
    Nucleophilic addition to a carbonyl group. Click to copy article linkArticle ... Bürgi–Dunitz Trajectory. The Journal of Organic Chemistry 2015, 80 (15) ...<|control11|><|separator|>
  12. [12]
    4.6: Nucleophilic Addition Reactions - Chemistry LibreTexts
    Sep 21, 2023 · It is called nucleophilic addition reaction because the overall result is the addition of a nucleophile ( HNu ) to a carbonyl ( C = O ) group.Learning Objectives · Base-promoted nucleophilic... · Acid-catalyzed nucleophilic...
  13. [13]
    [PDF] Aldehydes and Ketones I. Nucleophilic Addition to the Carbonyl Group
    Organic Chemistry II / CHEM. 252. Chapter 16 – Aldehydes and. Ketones I. Nucleophilic Addition to the Carbonyl Group. Bela Torok. Department of Chemistry.Missing: definition | Show results with:definition
  14. [14]
    CO6. Relative Reactivity of Carbonyls - carbonyl addition
    There are two kinds of electrophiles that commonly undergo nucleophilic addition at the carbonyl group: aldehydes and ketones. Of these, aldehydes are a little ...
  15. [15]
    Chapter 17 notes
    aldehydes more reactive than ketones (due to steric hindrance with ketones); aromatic carbonyls less reactive than aliphatic (due to conjugation effects); a ...
  16. [16]
    19.7: Nucleophilic Addition of Hydride and Grignard Reagents
    Jan 19, 2025 · Aldehydes and ketones will undergo nucleophilic addition with organolithium and Grignard reagent nucleophiles. The nucleophilic carbon in ...
  17. [17]
    Ch17: RLi or RMgX with Aldehydes or Ketones - University of Calgary
    Organolithium or Grignard reagents react with the carbonyl group, C=O, in aldehydes or ketones to give alcohols. The substituents on the carbonyl dictate the ...<|separator|>
  18. [18]
    Asymmetric addition of Grignard reagents to ketones - PubMed Central
    In conclusion, we have developed a new class of N,N,O-tridentate chiral ligands for single-metal controlled asymmetric addition of Grignard reagents to ketones.
  19. [19]
    19.3: Reductions using NaBH4, LiAlH4 - Chemistry LibreTexts
    Jul 1, 2020 · Note! LiAlH4 and NaBH4 are both capable of reducing aldehydes and ketones to the corresponding alcohol.Reduction of aldehydes and... · Exercises · Alcohols from carbonyl...
  20. [20]
    LiAlH4 and NaBH4 Carbonyl Reduction Mechanism - Chemistry Steps
    LiAlH4 and NaBH4 reduce carbonyls by hydride addition, catalyzed by lithium or hydrogen bonding, followed by protonation to form alcohols.
  21. [21]
    19.6: Nucleophilic Addition of HCN - Cyanohydrin Formation
    Jan 19, 2025 · Hydrogen cyanide (HC≡N), adds reversibly to aldehydes and many ketones forming hydroxyalkanenitrile adducts commonly known and as cyanohydrins.Study Notes · Mechanism of Cyanohydrin... · Other Cyanohydrins
  22. [22]
    Ch17: CN- to Cyanohydrin - University of Calgary
    Cyanide adds to aldehydes/ketones to form a cyanohydrin. This involves cyanide adding to the carbonyl group, followed by protonation of the alkoxide oxygen.
  23. [23]
    [PDF] Carboxylic Acid Derivatives: Nucleophilic Acyl Substitution 20.1
    Mechanism occurs in two stages. The first is addition of the nucleophile to the carbonyl carbon to form a tetrahedral intermediate. The second stage in collapse ...
  24. [24]
    Derivatives of Carboxylic Acids - MSU chemistry
    Inductive electron withdrawal by Y increases the electrophilic character of the carbonyl carbon, and increases its reactivity toward nucleophiles.
  25. [25]
    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 ...
  26. [26]
    Recent advances in the chemistry of imine-based multicomponent ...
    The reversible condensation between amines and carbonyl compounds to form imines is one of the most fundamental and ubiquitous reactions in chemistry. The ...
  27. [27]
    Lewis Acid Promoted Addition of Organometallics to Transient Imines
    May 22, 2019 · Activation of the transient imines by Lewis acids that are compatible with the presence of lithium alkoxides was found to be crucial to accommodate a broad ...
  28. [28]
    Enamines from Aldehydes and Ketones with Secondary Amines
    The reaction of aldehydes and ketones with secondary amines produces enamines. Enamines are amines with a double bond on the adjacent carbon.
  29. [29]
    A Practical Guide for Predicting the Stereochemistry of Bifunctional ...
    Apr 29, 2016 · This Account describes investigations into the mechanism of the phosphoric acid catalyzed addition of nucleophiles to imines, focusing on binaphthol-based ...
  30. [30]
    Reductive Amination - Chemistry Steps
    Reductive amination is used for preparing amines from aldehydes and ketones by reacting them in the presence of NaBH3CN at lower pH:
  31. [31]
  32. [32]
    [2 + 1] Cycloaddition reactions of fullerene C60 based on diazo ...
    They are synthesized in two main reactions: nucleophilic cyclopropanation according to the Bingel method and thermal addition of diazo compounds. This ...
  33. [33]
  34. [34]
    [PDF] α,β-Unsaturated Aldehydes and
    α,β-unsaturated aldehydes and ketones have a double bond conjugated with the carbonyl group, are more polar, and have two nucleophile attack sites.Missing: mechanism | Show results with:mechanism
  35. [35]
    The Michael Addition - Oregon State University
    The Michael addition is simply the conjugate addition of an enolate nucleophile to an α, β-unsaturated carbonyl compound. Protonation of the initial ...Missing: mechanism | Show results with:mechanism
  36. [36]
    CO22b. Michael Addition and Robinson Annulation - csbsju
    A Michael addition reaction. In a Michael addition, a doubly-activated enolate undergoes conjugate addition or 1,4-addition to an α,β-unsaturated carbonyl.
  37. [37]
    [PDF] The α-Carbon Atom and its pKa The inductive effect of the carbonyl ...
    β-carbon of α,β-unsaturated ketone and aldehydes. This is referred to a 1,4-addition or conjugate addition. When a reaction can take two possible path, it is ...
  38. [38]
    [PDF] Enol Content and Enolization
    In the case of 1,2- versus 1,4 addition of an α,β-unsaturated carbonyl,. 1,2 ... The Michael Reaction –The conjugate addition of a enolate ion to an α,β- ...
  39. [39]
    Illustrated Glossary of Organic Chemistry - Nucleophilic addition ...
    In a Michael addition reaction, a carbon nucleophile attacks the carbon-carbon pi bond of an α,β-unsaturated carbonyl compound. In this example, lithium ...<|control11|><|separator|>
  40. [40]
    Alkyl Halide Reactivity - MSU chemistry
    Later we shall find that Gilman reagents also display useful carbon-carbon bond forming reactions with conjugated enones and with acyl chlorides.
  41. [41]
    [PDF] Stereoselective conjugate addition of organolithium reagents to ...
    Following this procedure, the conjugate addition reaction was obtained instead of the products of direct addition to the carbonyl group usually found in other.