Fact-checked by Grok 2 weeks ago

Enolate

An enolate is an anionic species formed by the of a , such as an , , or , at the alpha carbon position adjacent to the , resulting in a resonance-stabilized structure where the negative charge is delocalized between the alpha carbon and the oxygen atom. Enolates are generated under basic conditions using bases ranging from moderate ones like or alkoxides, which establish an with the parent carbonyl, to strong, non-nucleophilic bases such as (LDA) or , which drive complete . The stability of enolates arises from this delocalization, making the alpha C–H bond more acidic (pKa ≈ 20 for simple ketones like acetone) compared to typical hydrocarbons, and further enhanced in beta-dicarbonyl compounds (pKa 9–13) due to additional conjugation. In , enolates serve as versatile nucleophiles, enabling key carbon-carbon bond-forming reactions that are foundational to constructing complex molecules. Notable reactions include the , where an enolate adds to another carbonyl to form β-hydroxy carbonyls or α,β-unsaturated carbonyls after dehydration; the , involving ester enolates to produce β-keto esters; and alpha-alkylation, where enolates react with alkyl halides to introduce substituents at the alpha position. These processes, along with variants like the Dieckmann cyclization and , underscore enolates' role in stereoselective synthesis and the preparation of pharmaceuticals, natural products, and other fine chemicals.

Definition and Properties

Basic Definition

An enolate is the conjugate base formed by the of a , such as a , , or , at the alpha carbon position adjacent to the . This alpha generates an anion stabilized by the adjacent carbonyl functionality, distinguishing enolates as key reactive intermediates in . The stability of the enolate arises from delocalization of the negative charge between two primary contributing structures: one with the charge on the alpha carbon ( form) and the other with the charge on the oxygen atom (alkoxide-like enolate form), accompanied by a carbon-carbon in the latter. For simple cases, these resonance structures are represented as: \ce{R-C(O^-)=CH-R' <-> R-C(=O)-CH^--R'} The enolate concept emerged in early 20th-century , particularly through mechanistic studies of aldol reactions by chemists like Arthur Lapworth, who employed early notations to describe enolate formation. This foundational understanding has since underscored enolates' importance in carbon-carbon bond-forming reactions central to synthesis.

Stability and Solvation

The stability of enolates is fundamentally tied to the acidity of the α-hydrogens in their parent carbonyl compounds, which determines the ease of and the resulting position. For ketones, these pKa values typically fall in the range of 19–21, while aldehydes exhibit slightly higher acidity with pKa values around 16–18, owing to reduced steric bulk and enhanced electron-withdrawing effects of the formyl group compared to alkyl-substituted carbonyls in ketones. This difference in acidity directly influences enolate stability, as lower pKa values indicate a greater thermodynamic favorability for the enolate form in aldehydes versus ketones. In solution, enolate and behavior are profoundly affected by aggregation phenomena, where metal enolates—particularly those of —frequently form dimers, tetramers, or higher oligomers as contact pairs. These aggregates arise due to the coordination of the metal cation to multiple enolate oxygens, as observed in the lithium enolate of p-phenylisobutyrophenone in (THF), where both dimeric and monomeric species coexist, with the surprisingly dominating reactivity in certain alkylations. Such aggregation reduces the effective nucleophilicity of the enolate by shielding the carbanionic center but can also impart kinetic by limiting rates. Solvation plays a pivotal role in modulating enolate stability by influencing ion pair and disruption. In less polar aprotic solvents like THF, enolates maintain tight ion pairs, preserving aggregation and somewhat attenuating reactivity, whereas more polar aprotic solvents such as (DMSO) promote solvent-separated ion pairs through stronger of the cation, thereby increasing the free enolate concentration and enhancing nucleophilicity. This -dependent is evident in conductometric studies of alkali enolates, where DMSO facilitates greater ion separation compared to THF, leading to improved thermodynamic stability for the dissociated species. Enolates generally exhibit kinetic rather than thermodynamic stability, functioning as transient intermediates prone to rapid or due to their strong basicity, though careful control of conditions can extend their persistence. Under kinetic protocols in aprotic media at low temperatures, enolates achieve short-term stability, while thermodynamic equilibration may favor the more substituted (stable) . In exceptional cases, certain enolates, such as those derived from anthracen-9-yl ketones, have been isolated as crystalline aggregates under strictly inert, conditions, allowing characterization of their solid-state structures and confirming their viability beyond solution-phase transients.

Structure and Bonding

Bonding and Resonance

Enolates are described as resonance hybrids of two primary canonical structures: the form, where the negative charge resides on the α-carbon adjacent to the , and the form, where the charge is localized on the oxygen atom. The form dominates the hybrid due to oxygen's higher , resulting in greater on oxygen, while the carbanion character enables preferential electrophilic attack at the carbon site. This delocalization results in partial negative charge on both the α-carbon and oxygen. Bond lengths in enolates reflect this delocalization, with the Cα-C(carbonyl) bond shortened relative to the parent carbonyl (indicating partial double-bond character) and the C=O bond elongated (indicating partial single-bond character). These changes are confirmed by crystallographic studies of metal-coordinated enolates and corroborated by computational geometry optimizations. further supports this, showing the C=O stretching frequency shifted to lower values (typically around 1550-1650 cm⁻¹ compared to 1710-1720 cm⁻¹ for ketones), due to the weakened carbonyl . In molecular orbital terms, the highest occupied (HOMO) of the enolate is a π-type orbital derived from the carbonyl π* orbital, but lowered in energy upon α-deprotonation, with significant on the α-carbon (larger coefficient than on oxygen). This HOMO distribution explains the enolate's ambident nucleophilicity, favoring C-alkylation under kinetic conditions. stabilizes the system by populating this delocalized orbital, enhancing reactivity compared to the neutral carbonyl precursor. Compared to enols, enolates represent the deprotonated anionic analogs, exhibiting greater charge delocalization without the charge-separated inherent in enols (where the minor contributor involves a C⁺-O⁻ ). This leads to enolates having enhanced stability and nucleophilicity, as the negative charge is more evenly distributed across the C=C-O framework, facilitating diverse synthetic transformations.

Molecular Geometry

Enolates adopt planar geometries at the α-carbon due to sp² hybridization, with stereochemical configurations designated as or based on the relative positions of the α-substituent and the carbonyl oxygen across the partial Cα–C(carbonyl) . This arises from delocalization, which imparts double-bond character and restricts rotation. In enolates of ketones, the Z configuration predominates, stabilized by wherein the lithium cation coordinates to both the enolate oxygen and the carbonyl oxygen within oligomeric s. For ester-derived enolates, the E configuration is typically favored under standard conditions due to minimized steric interactions between the alkoxy group and the α-substituent, though addition of (HMPA) shifts selectivity toward the Z isomer by solvating the lithium and disrupting aggregate formation. The model provides a framework for understanding through a cyclic chair-like during , where the bulky base approaches the α-proton anti to the larger substituent, predicting Z selectivity for most enolates and E for esters when the α-substituent is small. However, this monomeric model has limitations in modern contexts, as experimental evidence highlights the role of enolate aggregation and solvent effects in overriding simple predictions. Spectroscopic techniques confirm these configurations: multinuclear NMR reveals distinct chemical shifts and coupling constants for Z and E isomers, with the Z-lithium enolate of exhibiting dimeric or tetrameric aggregation in solvents at low temperatures, evidenced by line broadening and diffusion-ordered spectroscopy. further supports Z geometry in solid-state structures of lithium enolates, showing planar C–C–O units with bridged between oxygens in a chelated arrangement. Substituent effects introduce torsional around the Cα–C(carbonyl) , influencing planarity; bulky α- increase out-of-plane and pyramidality at the α-carbon, deviating from ideal trigonal geometry and affecting aggregate stability, as quantified by dihedral angles in computational models (e.g., O–C–C–H ≈ 10–15° in fused systems).

Formation

Deprotonation Methods

Enolates are typically generated through the of carbonyl compounds at the alpha position using bases of varying strength, which influences the extent of deprotonation and the of the process. Strong, non-nucleophilic bases such as (LDA) are commonly employed to achieve quantitative under kinetic control conditions, typically at low temperatures like -78°C in aprotic solvents such as (THF). This approach favors the formation of the less substituted enolate due to the steric bulk of LDA and the irreversibility of the , as the conjugate acid has a pKa around 36, higher than that of most alpha protons in ketones (pKa ~20). A representative example is the of acetone: \ce{CH3C(O)CH3 + LDA -> CH3C(O)CH2^- Li^+ + HN(iPr)2} For thermodynamic control, weaker bases like sodium ethoxide (NaOEt) or potassium tert-butoxide are used at higher temperatures in protic solvents, allowing equilibration to the more stable, substituted enolate through reversible proton transfer. These conditions fully deprotonate the substrate only partially due to their basicity (pKa of EtOH ~16), promoting the thermodynamically favored species in solvents where enolate stability can be influenced by solvation effects. Weaker bases, such as alkoxides (e.g., , NaOEt), enable partial in protic solvents like , leading to an mixture biased toward the thermodynamic enolate. The lower basicity of NaOEt ( of EtOH ~16) results in only a small of enolate formation, but the reversible nature allows interconversion, favoring the more substituted under ambient conditions. Metal enolates of , sodium, and are formed by with the corresponding metal amides or hydrides, yielding salts that differ in reactivity due to cation size and coordination properties; lithium enolates are often monomeric or dimeric in THF, while sodium and potassium variants tend to aggregate more extensively. Organocatalytic methods, such as phase-transfer using chiral quaternary ammonium salts, facilitate enolate generation from aqueous bases like NaOH by transferring the anion into organic phases, enabling efficient without strong aprotic bases.

Regioselectivity

In unsymmetrical carbonyl compounds, regioselectivity during enolate formation presents a significant synthetic challenge, as multiple alpha protons may be available, leading to mixtures of regioisomeric enolates. This control is essential for directing subsequent reactivity toward desired products in . The choice of conditions determines whether deprotonation favors the less substituted (kinetic) or more substituted (thermodynamic) enolate, exploiting differences in reaction rates or stabilities. Kinetic enolates, which are less substituted and form more rapidly due to lower steric hindrance at the alpha site, are generated under conditions that promote irreversible . Strong, sterically hindered bases such as (LDA) at low temperatures (typically -78 °C in ) selectively abstract the more accessible proton, preventing equilibration. In contrast, thermodynamic enolates, which are more substituted and stabilized by greater alkyl substitution on the enolate double bond, predominate under equilibrating conditions, such as higher temperatures or weaker bases like , allowing proton exchange between regioisomers. The small differences between alpha positions—approximately 1-2 units, as seen in 2-butanone where the methylene group's alpha proton has a of ~26.5 in DMSO compared to ~27.6 for the —enable this kinetic bias, since the thermodynamic preference is modest and can be overridden by rate control. A representative example is 2-butanone (CH3C(O)CH2CH3), where kinetic deprotonation with LDA yields primarily the less substituted terminal enolate (CH2=C(O⁻)CH2CH3), while thermodynamic conditions favor the more substituted internal enolate (CH3C(O⁻)=CHCH3). In particularly hindered substrates, where standard bases like LDA may encounter steric issues, lithium 2,2,6,6-tetramethylpiperidide (LTMP)—a bulkier non-nucleophilic base—improves regioselectivity for the kinetic enolate by enhancing approach to congested alpha protons without promoting side reactions.

Role of Lewis Acids and Additives

Lewis acids, such as BF₃·OEt₂ and MgBr₂, coordinate to the oxygen atom of the carbonyl group in ketones and esters, thereby enhancing the acidity of the α-protons and enabling enolate formation using milder, weaker bases like Et₃N or i-Pr₂NEt under ambient conditions. This coordination activates the substrate by polarizing the C=O bond, lowering the pK_a of the α-hydrogen and promoting selective deprotonation without requiring strong bases like LDA, which can lead to over-deprotonation or side reactions. Such Lewis acid-mediated approaches are particularly valuable for generating "soft" enolates from esters, where the coordinated species reduce the tendency for self-condensation by stabilizing the enolate and suppressing nucleophilic attack on the ester carbonyl. Solvent choice plays a crucial role in modulating enolate reactivity through effects on and ion pairing. Polar aprotic solvents like HMPA or solvate or other cations strongly, promoting desolvation of the enolate anion and increasing its nucleophilicity by disrupting tight ion pairs or aggregates in THF. This enhancement is evident in reactions, where HMPA additives can shift selectivity toward C-alkylation by making the enolate more "naked" and reactive. Similarly, crown ethers such as 12-crown-4 or 18-crown-6 sequester cations, breaking down dimeric or higher-order enolate aggregates into more reactive monomeric species, thereby accelerating reactions and improving yields in non-polar media. A representative application involves the formation of enolates using TiCl₄ in conjunction with a like Bu₃N, which generates titanium-coordinated enolates suitable as precursors for crossed-Claisen condensations. In this process, TiCl₄ activates the by coordination, allowing selective and subsequent reaction with acid chlorides to afford β-keto esters in high yields (up to 95%) with minimal self-condensation, demonstrating the utility of such additives in synthetic planning.

Reactivity

Nucleophilic Additions

Enolates function as potent carbon nucleophiles in addition reactions with electrophilic π-systems, such as s and conjugated alkenes, while their ambident character allows for competing oxygen-centered reactivity depending on the electrophile's nature. These additions are fundamental in for constructing carbon-carbon bonds, with the enolate's nucleophilicity enhanced by the adjacent that stabilizes the negative charge. The reactivity often proceeds through a concerted or stepwise involving deprotonation of the initial . The exemplifies enolate , wherein the enolate attacks the carbonyl carbon of an to yield a β-hydroxy carbonyl compound. This process typically occurs under basic conditions and can be represented by the equation: \ce{^{-}CH2C(O)R' + RCHO ->[H+] RCH(OH)CH2C(O)R'} The stereochemical outcome of the aldol addition is rationalized by the Zimmerman-Traxler model, which posits a chair-like, six-membered cyclic structure coordinating the enolate's metal with the aldehyde oxygen. For or enolates, (Z)-enolates favor syn diastereoselectivity through a minimizing steric interactions, while (E)-enolates lead to anti products; experimental studies confirm preferences up to 50:1 for chair-like pathways over boat alternatives. In Michael additions, enolates undergo conjugate addition to the β-position of α,β-unsaturated carbonyl acceptors, forming enolates that protonate to 1,5-dicarbonyl compounds. This 1,4-addition exploits the electrophilic activation by the , with enolates typically adding efficiently to enones or enals under mild conditions. Seminal work demonstrated high yields for such additions using preformed enolates, establishing the reaction's utility for extending carbon chains. The ambident nature of enolates manifests in during nucleophilic additions, where the carbon terminus (softer nucleophilic site) predominates with soft electrophiles like primary alkyl halides in conjugate systems, yielding C-addition products. Conversely, hard electrophiles such as Meerwein trialkyloxonium salts promote O-attack, forming ethers. This dichotomy follows the hard-soft acid-base , with metal counterions and polarity modulating the C/O —lithium enolates in aprotic media favor C-addition by over 90% in many cases.

Alkylation and Acylation

Enolates act as carbon nucleophiles in alpha-alkylation reactions with primary alkyl halides, proceeding via an SN2 mechanism to forge new C-C bonds selectively at the alpha carbon. This process replaces an alpha hydrogen with an alkyl group and is favored with unhindered primary electrophiles such as methyl or ethyl iodides, which undergo clean backside displacement. To suppress polyalkylation—a common issue arising from the increased acidity of the monoalkylated product relative to the parent carbonyl—an excess of enolate is employed, ensuring high yields of the monoalkylated ketone or ester. Acylation of enolates with acid chlorides or anhydrides provides a direct route to 1,3-dicarbonyl compounds, particularly 1,3-diketones from ketone-derived enolates. The enolate carbon attacks the electrophilic carbonyl of the , displacing chloride or the to form the C-acylated product. These reactions exhibit excellent under kinetic control, often using lithium enolates to favor C-acylation over O-acylation, and are tolerant of various functional groups on the acyl component. The resulting 1,3-diketones are versatile synthons due to their enhanced acidity and content. Stereocontrol in these reactions is governed by the enolate geometry and conformational preferences, enabling diastereoselective outcomes. In enolates, such as those derived from 2-methyl, the preferentially approaches from the axial face of the planar enolate, leading to trans diastereomers in the 2,6-disubstituted products. This kinetic preference arises from minimized steric repulsion in the , where the incoming group aligns with the pseudo-axial trajectory, achieving diastereoselectivities often exceeding 90:10. Such control is critical for constructing stereodefined frameworks in synthesis. A classic illustration is the alkylation of the enolate of ethyl acetoacetate with methyl iodide, yielding ethyl 2-methyl-3-oxobutanoate as the alpha-methylated product. Generated using sodium ethoxide in ethanol, this enolate undergoes efficient SN2 alkylation at the activated methylene, setting the stage for further elaboration in the acetoacetic ester synthesis.

Synthetic Applications

Enolates play a pivotal role in target-oriented synthesis, particularly through masked variants such as silyl enol ethers, which enable regiospecific enone formation via Robinson annulation. In this approach, the silyl enol ether of a ketone undergoes Lewis acid-catalyzed Michael addition to methyl vinyl ketone (MVK), yielding a 1,5-diketone intermediate that cyclizes via intramolecular aldol condensation followed by dehydration to afford the α,β-unsaturated enone. This variation provides superior regioselectivity over classical base-mediated methods, avoiding competitive self-condensation and allowing access to fused cyclohexenone systems in high yields, as demonstrated in the synthesis of octalones from cyclohexanone-derived silyl enol ethers. A landmark application of enolate chemistry is found in the of progesterone, where sequential enolate alkylations build the steroidal carbon framework, complemented by an intramolecular to form the A-ring. Reported by and coworkers in 1971, this biomimetic route employs enolates for stereocontrolled C-C bond formations, culminating in a polyene cyclization and steps to deliver racemic progesterone in 18 steps from simple precursors, highlighting enolates' efficiency in constructing complex polycyclic structures. Modern synthetic applications leverage enolates for asymmetric synthesis, notably through chiral auxiliaries in aldol reactions. Evans' methodology uses N-acyl oxazolidinones to generate boron enolates, which add to aldehydes with high diastereoselectivity via a , enabling the preparation of syn-β-hydroxy carbonyls in enantiopure form after auxiliary . This approach has been instrumental in synthesizing fragments and natural products like discodermolide, achieving >95% in many cases. Complementing this, organocatalytic enolate equivalents, such as ammonium enolates from chiral tertiary amines, facilitate asymmetric alkylations and additions without metal mediators, as reviewed in methods generating centers with up to 99% . Enone synthesis can also proceed via kinetic enolates reacting with α-halo ketones to form 1,4-dicarbonyl compounds, followed by intramolecular . Under kinetic conditions with LDA at low temperature, the enolate of a methyl displaces the halide in an α-bromoacetone equivalent, affording the 1,4-dicarbonyl compound; subsequent acid-catalyzed dehydration yields the α,β-enone, providing a regioselective route to conjugated systems as utilized in precursors.

Aza-Enolates

Aza-enolates are nitrogen-containing analogs of enolates, specifically the conjugate bases of imines, exhibiting resonance structures such as R-CH⁻-CR'=N-R'' ↔ R-CH=CR'-N⁻-R''. These species share similarities with traditional oxygen-based enolates, allowing delocalization of the negative charge between the alpha carbon and the atom. Preparation of aza-enolates typically involves at the α-carbon of the corresponding precursor using strong bases such as (LDA). Alternative methods include transmetallation or activation with Lewis acids like . In terms of reactivity, aza-enolates preferentially undergo C-alkylation at the α-carbon due to the nucleophilic character of the carbanionic form, making them valuable for forming new carbon-carbon bonds. They are particularly effective in reactions with allyl halides, such as , yielding alkylated products in high yields (e.g., 62–91% with activated systems). Additionally, lithium aza-enolates can ring-open epoxides, including strained heterocycles like , which are often less reactive toward standard enolates. A representative synthetic application is the one-pot formation and of aza-enolates derived from secondary amines, such as , which has been employed in the expeditious synthesis of the male aggregation of the Oulema melanopus. In this process, oxidation of the amine generates an intermediate, followed by in situ to form the aza-enolate, which is then alkylated with an appropriate and hydrolyzed to the target .

Silicon and Other Enolates

Silyl enol ethers represent neutral, masked forms of enolate anions, where the enolate oxygen is silylated with a trimethylsilyl (TMS) group, offering enhanced stability and ease of isolation compared to reactive ionic enolates. They are typically generated by quenching kinetically formed lithium enolates—prepared from ketones and (LDA) at low temperatures—with chlorotrimethylsilane (TMSCl) in an aprotic solvent like . This approach ensures high , favoring the less substituted (kinetic) isomer due to the irreversible trapping under low-temperature conditions. An alternative metal-free protocol for kinetic employs N,O-bis(trimethylsilyl) (BSA) with catalytic tetrabutylammonium (TBAF), which promotes selective and silylation without strong bases, accommodating sensitive substrates. The primary advantages of silyl enol ethers lie in their greater thermal and chemical stability, allowing storage and purification by distillation or chromatography, unlike enolates which require generation in situ to avoid decomposition or side reactions. Upon activation with Lewis acids, they regenerate the enolate nucleophilicity in a controlled manner, minimizing competitive pathways like self-aldol condensation. A seminal application is the Mukaiyama aldol reaction, where silyl enol ethers couple with aldehydes under catalysis by titanium tetrachloride (TiCl4) or boron trifluoride etherate (BF3•OEt2) to afford β-hydroxy ketones with predictable syn/anti stereochemistry, often superior to traditional enolate aldols due to the neutral conditions. Beyond silicon-based masks, phosphonate-stabilized carbanions function as enolate equivalents in the Horner-Wadsworth-Emmons (HWE) olefination, providing a robust route to (E)-α,β-unsaturated esters from aldehydes and ketones. The reaction proceeds via base-mediated of a phosphonoacetate (e.g., with ), generating a resonance-stabilized anion that undergoes Wittig-like addition-elimination, with the leaving group ensuring high E-selectivity and yields often exceeding 80% for electron-deficient systems. This method's stability advantages stem from the electron-withdrawing , allowing milder conditions than unstabilized enolates and broad tolerance in synthesis. Sulfur ylides, such as dimethylsulfoxonium methylide, serve as non-carbonyl enolate mimics through their carbanionic reactivity, particularly in the , where they act as nucleophilic methylene donors to aldehydes or ketones, forming epoxides as masked 1,2-diols analogous to enolate addition products. Generated from sulfoxonium salts and bases like , these ylides exhibit tunable reactivity—semi-stabilized variants favor of α,β-unsaturated carbonyls, extending carbon frameworks with high diastereocontrol—and offer handling ease due to their neutral precursors, though they require careful exclusion of protic impurities to prevent decomposition.

References

  1. [1]
    enolate ions - csbsju
    An enolate ion is the anion that forms when a proton is removed next to a carbonyl. The carbon next to the carbonyl is called the α-position (alpha position).Missing: definition | Show results with:definition
  2. [2]
    None
    ### Summary of Enolate Ions and Enols (Chapter 18, Neuman Organic Chemistry)
  3. [3]
    Enolates - Formation, Stability, and Simple Reactions
    Aug 16, 2022 · Enolates can be thought of as the conjugate bases of aldehydes and ketones (among others) but they are also the conjugate bases of enols.
  4. [4]
    The iconic curly arrow | Feature - Chemistry World
    Mar 31, 2010 · Sir Robert Robinson (1886-1975) was one of the giants of early 20th century organic chemistry. ... enolate formation. The Robinson and ...
  5. [5]
    Thermodynamic and Kinetic Controlled Enolates: A Project for a ...
    This work involves regioselective alkylations of 2-methylcyclohexanone via thermodynamically and kinetically controlled enolates.
  6. [6]
    Ch21: Acidity of alpha hydrogens - University of Calgary
    Let's compare pKa of the common systems: aldehyde pKa = 17, ketone pKa = 19 and an ester pKa = 25, and try to justify the trend. comparing aldehydes and ketones
  7. [7]
    6.1 The Acidity of the α-Hydrogens – Organic Chemistry II
    The α-hydrogens in the carbonyl compound, which also bond to the sp 3 carbon, are much more acidic with pK a in the range of 16 to 20.
  8. [8]
    Aggregation of the Lithium Enolate of p-Phenylisobutyrophenone in ...
    Aggregation of the Lithium Enolate of p-Phenylisobutyrophenone in THF: The Unexpected Importance of Monomer | Journal of the American Chemical Society.
  9. [9]
    Aggregation and Alkylation Kinetics of the Lithium Enolate of p ...
    Ion pair aggregates and reactions; experiment and theory. ... Aggregation and ion pair acidity in THF: lithium and cesium enolates of α-phenylcyclohexanone.
  10. [10]
    Ion pairing and reactivity of enolate anions. 5. Thermodynamics of ...
    Thermodynamics of ionization of .beta.-di- and tricarbonyl compounds in dimethyl sulfoxide solution and ion pairing of their alkali salts.Missing: THF | Show results with:THF
  11. [11]
    Studies of Solvation Phenomena of Ions and Ion Pairs in ...
    Counterion and solvation effects on the conformations of the radical anions of 2,5-dicyanovinyl substituted furan and thiophene. An EPR and ENDOR study. J ...
  12. [12]
    Aggregation and Alkylation of the Cesium Enolate of 2-p ...
    The cesium enolates correspond to ion pair pK's (referred to cesium salts of delocalized indicators) that are only slightly smaller (∼1−2 units) than the ionic ...
  13. [13]
    Lithium, sodium and potassium enolate aggregates and monomers ...
    Feb 12, 2024 · But the structural studies of isolated alkali metal enolates are not as well-established as one may expect: the state-of-the-art is covered in ...
  14. [14]
    Concerted or Stepwise: How Much Do Free-Energy Landscapes Tell ...
    In this study, both 1- and 2-dimensional free energy surfaces are generated for these compounds with various substituents, using density functional theory.
  15. [15]
    Recent Advances in the Chemistry of Heavier Group 14 Enolates
    Jul 19, 2019 · Selected examples of single-crystal X-ray structures of HG 14 metal enolates. ... Selected bond lengths d [Å] and selected sum of valence ...
  16. [16]
    The Infrared Spectra of Enolate Ions
    ### Summary of IR Spectra of Enolate Ions (C=O Stretch Frequencies, Bond Lengths, Resonance)
  17. [17]
    [PDF] Reactions at α-Position In preceding chapters on carbonyl chemistry ...
    Consider the HOMO for the enolate nucleophile: Enolate structure. HOMO of enolate. The charge in the HOMO for the unsymmetrical enolate is far greater on the ...
  18. [18]
    Enols and Enolates of Carbonyl Compounds and Their Reactions
    Enols are isomers of aldehydes or ketones in which one alpha hydrogen has been removed and replaced on the oxygen atom of the carbonyl group.Missing: definition | Show results with:definition
  19. [19]
    [PDF] Structure and Reactivity of Lithium Enolates. From Pinacolone to ...
    Addition of LiX to an enolate solution could lead to formation of a mixed aggregate (A-B), and therefore change the product ratio P'IP'. A chiral additive ...Missing: EZ | Show results with:EZ
  20. [20]
    [PDF] University of Groningen Addition of Enolates and Azaenolates to α,β ...
    A correlation between enolate geometry and product stereochemistry is found, with (Z)-eno- lates producing anti-adducts and (E)-enolates yielding syn-adducts ...Missing: EZ | Show results with:EZ
  21. [21]
    None
    ### Summary of Enolate Geometries, E/Z for Ketones and Esters, Ireland Model, Effect of HMPA, Torsional Strain or Planarity
  22. [22]
    Conformationally Locked Pyramidality Explains the ...
    Apr 27, 2020 · To assess the relation between the torsional strain induced in the enolate along the two methylation pathways and the degree of pyramidalization ...
  23. [23]
  24. [24]
    Lithium Diisopropylamide-Mediated Enolizations: Solvent ...
    We have carried out a two-part investigation of LDA-mediated enolizations. In this paper we will describe detailed rate studies under pseudo-first-order ...Missing: original | Show results with:original
  25. [25]
    [PDF] gcw.Enolate Seminar 3.12.8pm - Macmillan Group
    Mar 12, 2008 · Chelate provides organizational role in fixing orientation between resident asymmetric center and enolate system. Evans, D. A. "Stereoselective ...
  26. [26]
    A Rational Approach to Catalytic Enantioselective Enolate Alkylation ...
    A Systematic Investigation of Quaternary Ammonium Ions as Asymmetric Phase-Transfer Catalysts. Synthesis of Catalyst Libraries and Evaluation of Catalyst ...
  27. [27]
    5.5: Alkylation of Enolate Ions
    ### Summary of Regioselectivity, Kinetic vs Thermodynamic Enolates, Conditions, Examples, and pKa Differences
  28. [28]
    Bordwell pKa Table - Organic Chemistry Data
    Below are tables that include determined pKa values for various acids as determined in water, DMSO and in the gas Phase. These tables are compiled in PDF files ...
  29. [29]
  30. [30]
    Catalytic Asymmetric Conjugate Addition and Allylic Alkylation with ...
    The results indicate that Mg2+ ions not only activate the enone (or enoate) via coordination to the oxygen atom of the substrate (Lewis acid effect) but also ...
  31. [31]
    [PDF] CHEM 330 Topics Discussed on Oct 2 Effect of solvent: C-reactivity ...
    O-reactivity is more pronounced in enolates prepared in strongly Lewis basic solvents (=strongly coordinating), e.g., DMSO, HMPA, DMPU and the like.Missing: crown cation sequestration
  32. [32]
    Formation, Reactivity and Decomposition of Aryl Phospha‐Enolates
    Nov 11, 2022 · Addition of 12-crown-4 to 1 a or 1 b effectively sequesters the lithium cation, breaking up the dimeric structure to form monomeric salts 2 a ...
  33. [33]
    Chemical Reactivity - MSU chemistry
    The ambident nature of enolate anions also enables electrophilic attack at both oxygen and carbon, but in most synthesis applications bonding to carbon is ...
  34. [34]
    Advances in the Chemistry of Ambident Enolate and Phenoxide Ions
    The reactivities of ambident enolate and phenoxide ions in alkylation reactions are examined. It is shown that the mode of C- and O-alkylation of an ambident ...
  35. [35]
    Chemistry of carbanions. XIX. Alkylation of enolates from ...
    Article August 1, 1971 Chemistry of carbanions. XIX. Alkylation of enolates from unsymmetrical ketones
  36. [36]
    Alkylation of enolate ions generated regiospecifically via ...
    Alkylation of enolate ions generated regiospecifically via organocopper reactions. Synthesis of decalin sesquiterpene valerane and of prostaglandin model ...Missing: cyclohexanone axial
  37. [37]
    Acylation. I. The Mechanisms of Enol Ester and 1,3-Diketone ...
    A novel synthesis of 1,3-diketones by reaction of an α-bromoketone with acyl chlorides promoted by Gallium triiodide.Missing: yielding | Show results with:yielding
  38. [38]
    The Alkylation of Esters and Nitriles - ResearchGate
    This chapter is concerned with the reactions of metal salts (enolates) of active methylene compounds with alkylating agents such as alkyl halides to produce ...
  39. [39]
  40. [40]
    Imine Azaenolates: Synthesis, Reactivity, and Outlook - 2022
    Jul 5, 2022 · Azaenolates are, quite simply, the aza variant of enolates. Compared to their oxygen counterparts, additional control of the reactivity of ...
  41. [41]
    One-pot formation of aza-enolates from secondary amines and ...
    This process has been applied to an expeditious synthesis of the male aggregation pheromone of the cereal leaf beetle Oulema melanopus. Graphical Abstract.
  42. [42]
    Silyl enol ether synthesis by silylation - Organic Chemistry Portal
    A remote functionalization strategy enables a Z-selective synthesis of silyl enol ethers of (hetero)aromatic and aliphatic ketones via Ni-catalyzed chain ...
  43. [43]
    New cross-aldol reactions. Reactions of silyl enol ethers with ...
    The Mechanism of Iron(II)-Catalyzed Asymmetric Mukaiyama Aldol Reaction in Aqueous Media: Density Functional Theory and Artificial Force-Induced Reaction Study.Missing: seminal | Show results with:seminal