Fact-checked by Grok 2 weeks ago

Silyl enol ether

Silyl enol ethers are a class of organosilicon compounds derived from , in which the hydroxyl group is replaced by a silyloxy moiety, typically -OSi(CH₃)₃ (trimethylsilyloxy), resulting in a structure featuring a adjacent to the oxygen-bound atom, generally represented as R₂C=CR-OSiR'₃. These compounds serve as stable, isolable equivalents of enolate anions, offering enhanced reactivity and selectivity in synthetic transformations due to the silicon-oxygen bond, which imparts stability under neutral conditions while allowing activation toward electrophiles. In , silyl enol ethers are prepared through the of carbonyl compounds, such as ketones or aldehydes, often using silyl chlorides like chlorotrimethylsilane in the presence of a base to generate the intermediate, or via alternative methods including ruthenium-catalyzed dehydrogenative with triorganosilanes and nickel-catalyzed processes from olefins. Their geometric isomerism (E or Z) can be controlled, influencing in downstream reactions, and they are particularly valued for avoiding the handling challenges of reactive . The primary applications of silyl enol ethers lie in carbon-carbon bond-forming reactions, most notably the , where they act as nucleophiles with aldehydes or other electrophiles under to produce β-hydroxy carbonyl compounds with high regio- and stereocontrol. They also feature prominently in alkylations, Michael additions, and strategies, enabling the construction of complex molecular frameworks in natural product synthesis and pharmaceutical development.

Structure and Properties

Chemical structure and nomenclature

Silyl enol ethers are organosilicon compounds featuring a silyl ether linked to an moiety, with the general molecular R₃Si–O–CR=CR₂, where the R groups on are commonly alkyl substituents such as methyl in trimethylsilyl (TMS) derivatives or tert-butyl and methyl in tert-butyldimethylsilyl (TBDMS) variants, and the enol segment originates from the of ketones or aldehydes. This structure incorporates a characteristic carbon-carbon conjugated to the oxygen-silicon linkage, rendering silyl enol ethers stable, isolable equivalents of metal enolates that enable reactivity by inverting the inherent electrophilicity of the parent carbonyl carbon to nucleophilicity at the β-position. Nomenclature for silyl enol ethers typically employs common descriptive terms based on the parent carbonyl compound, such as "trimethylsilyl ether of ," which highlights their derivation and utility in synthesis. Systematic IUPAC naming treats them as substituted , constructing names like trimethyl[(1-methylethenyl)oxy]silane for the enol ether derived from acetone, where the alkenyloxy group specifies the enol configuration attached to the silyloxy unit. These names prioritize the parent chain while detailing the unsaturated alkoxy , ensuring precise identification of the molecular framework. Due to the C=C , silyl enol ethers exhibit geometric isomerism, manifesting as E and Z configurations that influence reactivity and selectivity in subsequent transformations. For unsymmetrical carbonyl precursors, such as methyl alkyl ketones, regioisomeric forms arise depending on the site of enolization, exemplified by the 1-regioisomer (terminal , CH₂=CR–OSiR₃) versus the 2-regioisomer (internal , R'CH=CR–OSiR₃), with kinetic conditions often favoring the less substituted variant.

Physical and chemical properties

Silyl enol ethers are generally colorless liquids or low-melting solids at . For instance, simple trimethylsilyl derivatives, such as (isopropenyloxy)trimethylsilane, exhibit boiling points in the range of 80–120°C under . These compounds are highly soluble in common organic solvents including , , and , but they are insoluble in and display to hydrolytic conditions even in moist environments. Chemically, silyl enol ethers demonstrate sufficient thermal stability to allow under reduced pressure, typically enduring temperatures up to 100–150°C without significant decomposition, depending on the . They are prone to under either acidic or basic aqueous conditions, reverting to the corresponding carbonyl compounds and byproducts, a process accelerated by protic media. However, under conditions, they remain inert toward many nucleophilic and electrophilic reagents, facilitating their isolation, purification, and short-term storage in sealed containers. Spectroscopic methods provide reliable identification of silyl enol ethers. Infrared (IR) spectroscopy reveals characteristic absorptions for the alkene C=C stretch at approximately 1640–1680 cm⁻¹ and the Si–O linkage at around 1000 cm⁻¹, with additional Si–CH₃ deformation bands near 1250 cm⁻¹. In ¹H nuclear magnetic resonance (NMR) spectra, the vinylic protons adjacent to the oxygen appear in the deshielded region of 4.0–5.5 ppm, often as multiplets reflecting geometric isomerism. Confirmation via ²⁹Si NMR typically shows a resonance for the trimethylsilyl group between 15 and 25 ppm, distinct from other silane functionalities. Safety considerations for handling silyl enol ethers include their flammability, with flash points often below 20°C, classifying them as highly flammable liquids that require storage away from ignition sources. They may cause and eye irritation upon contact, attributable to volatile silicon-containing species, and should be manipulated in a with appropriate protective equipment.

Synthesis

Preparation from carbonyl compounds

Silyl enol ethers are commonly prepared from ketones and aldehydes through the generation of s followed by silylation. The standard procedure involves of the carbonyl compound using a strong, non-nucleophilic base such as (LDA) or (NaH), which forms the enolate intermediate, subsequently trapped by a silyl chloride like chlorotrimethylsilane (TMSCl). This method, pioneered by and coworkers, typically employs aprotic solvents such as tetrahydrofuran (THF) or dimethylformamide (DMF), often with additives like (HMPA) to enhance solubility and yields. Reactions are conducted under inert atmosphere at low temperatures (e.g., -78°C for kinetic control with LDA) to minimize side reactions, affording silyl enol ethers in 70–95% yields for simple substrates like acetone or . The overall transformation can be represented as: \ce{R2C=O + base ->{{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}} R2C(-) - O(-)} \quad \ce{R2C(-) - O(-) + R'3SiCl ->{{grok:render&&&type=render_inline_citation&&&citation_id=2&&&citation_type=wikipedia}} R2C=CR - OSiR'3 + MCl} where step 1 is enolate formation and step 2 is silylation, with M denoting the metal counterion from the base. Alternative routes enable direct silylation without isolating the enolate. For instance, N,O-bis(trimethylsilyl)acetamide (BSA) reacts with carbonyl compounds in the presence of a base or catalyst, such as in ionic liquids, providing a milder, often solvent-free approach with good yields (typically >80%) for a range of aldehydes and ketones. Similarly, trimethylsilyl trifluoromethanesulfonate (TMSOTf) combined with a tertiary amine base like triethylamine facilitates efficient silylation under ambient conditions in dichloromethane, suitable for sensitive substrates and yielding 75–90% for aryl alkyl ketones. Another pathway utilizes the rearrangement, involving 1,2-silyl . From acylsilanes, addition of a to the carbonyl generates an that rearranges to a , which upon yields the silyl enol ether; this regiospecific method is particularly useful for accessing specific isomers from α-silyl carbonyl precursors. α-Silyl alcohols can also undergo to alkoxides, triggering rearrangement to the corresponding silyl enol ether after , often in THF with strong bases like n-BuLi. These conditions maintain aprotic environments to ensure clean , with yields around 80% for aliphatic systems. The choice of base in enolate-based methods can influence , though detailed control is addressed in specialized contexts. Catalytic methods have also been developed for the synthesis of silyl enol ethers. Ruthenium-catalyzed dehydrogenative allows the direct coupling of enolizable carbonyl compounds with triorganosilanes, such as using a tethered complex to form silyl enol ethers from ketones and hydrosilanes under neutral conditions, often achieving high yields (>80%) for various substrates. Similarly, nickel-catalyzed processes enable the from ketones bearing remote olefins via chain-walking functionalization, providing Z-selective silyl enol ethers with up to 95% yield and high , as reported in 2021.

Regioselectivity and stereoselectivity control

The in the formation of silyl enol ethers from unsymmetrical ketones is governed by whether the reaction proceeds under kinetic or thermodynamic control, which is dictated by the choice of base, temperature, and reaction time. Under kinetic conditions, a strong, sterically hindered base such as (LDA) at low temperature (−78 °C) in (THF) selectively deprotonates the less hindered α-carbon to form the less substituted ; subsequent trapping with chlorotrimethylsilane (TMSCl) yields the kinetic silyl enol ether with high (>95:5 in many cases). In contrast, thermodynamic control is achieved using milder bases like triethylamine (Et₃N) with TMSCl, often in (DMF) at or higher, allowing proton equilibration to favor the more substituted, conjugated and thus the thermodynamic silyl enol ether ( up to 90:10 favoring the more stable ). A representative example is the preparation from (PhC(O)CH₂CH₃). Kinetic conditions with LDA/TMSCl at −78 °C afford predominantly the less substituted 1-phenylprop-1-en-1-yloxytrimethylsilane (Ph–C(OSiMe₃)=CH–CH₃, >95% ), while thermodynamic conditions with Et₃N/TMSCl yield the more substituted (E)-1-phenylprop-1-en-2-yloxytrimethylsilane (Ph–C(CH₃)=CH–OSiMe₃, ~85:15 E/Z). These differences arise from the kinetic enolate's irreversible formation due to the bulky base's preference for the less sterically encumbered site, whereas thermodynamic conditions permit reversible to the more stable stabilized by conjugation or substitution. Stereoselectivity in silyl enol ether formation is particularly relevant for acyclic and cyclic ketones, where the geometry (E/Z) or approach direction influences the product. In cyclic ketones such as 4-tert-butylcyclohexanone, kinetic deprotonation with LDA favors axial proton abstraction from the equatorial conformation, leading to the enolate (and thus silyl enol ether) with the double bond oriented away from the axial hydrogens, achieving >90% selectivity for the less hindered isomer. For asymmetric induction, chiral lithium amide bases derived from (-)-sparteine, such as (–)-sparteine/LiTMP (lithium 2,2,6,6-tetramethylpiperidide), enable enantioenriched kinetic enolates from achiral ketones; trapping with TMSCl provides silyl enol ethers with up to 95% ee, as demonstrated in the deprotonation of 4-phenylcyclohexanone to yield the (R)-enriched silyl enol ether. Additional factors influencing selectivity include solvent and silyl group sterics. Solvents like 1,2-dimethoxyethane (DME) enhance kinetic control by coordinating the cation less effectively than THF, promoting faster, irreversible formation and improving to >98:2 in LDA-mediated reactions of methyl ketones. The size of the silyl group also affects E/Z ratios in acyclic silyl enol ethers; bulkier groups like triisopropylsilyl () favor the Z isomer (up to 15:1 Z/E) due to minimized A^{1,3} allylic strain in the , compared to trimethylsilyl (TMS), which gives more balanced mixtures (~3:1 E/Z under thermodynamic conditions).

Reactions

General reactivity as nucleophiles

Silyl enol ethers function as neutral reagents, inverting the inherent electrophilicity of carbonyl compounds by rendering the β-carbon nucleophilic through enolate-like delocalized across the C=C . This nucleophilicity arises from the push-pull electronic structure, where the oxygen lone pairs donate into the π-system, increasing at the β-position. The substituent further modulates this reactivity by stabilizing the during nucleophilic attack, particularly through involving Si–C σ-bonds and the developing positive charge on the or the intermediate. In general, silyl enol ethers engage electrophiles such as aldehydes and alkyl halides via addition at the β-carbon, typically under Lewis acid catalysis. Common activators include TiCl₄ and BF₃·OEt₂, which coordinate to the electrophile's electron-withdrawing groups or to the enol ether's oxygen, lowering the activation barrier for bond formation. The reaction proceeds through a concerted or stepwise mechanism, culminating in a β-silyloxy carbonyl adduct; the silyl group often migrates to the newly formed oxygen or is eliminated as a silanol or silyl triflate during aqueous workup. Kinetic studies reveal their nucleophilicity to be moderate, falling between that of allylsilanes and allylstannanes toward carbenium ions, enabling selective reactions with activated electrophiles. Compared to traditional anions, silyl enol ethers offer significant advantages as nucleophiles, including neutrality that prevents self-condensation and proton abstraction side reactions, enhanced solubility in nonpolar organic media, and reduced basicity for better tolerance. These properties allow for milder reaction conditions and improved in mixtures of regioisomeric enol ethers. However, their limitations include inherent sensitivity to protic solvents and moisture, which can hydrolyze them back to carbonyls, and the need for stoichiometric acids to activate less electrophilic partners, potentially complicating or introducing metal residues.

Carbon-carbon bond formation

Silyl enol ethers act as versatile nucleophiles in carbon-carbon bond forming reactions, particularly under , where they engage at the β-carbon position. One of the most widely used transformations is the Mukaiyama aldol reaction, in which a silyl enol ether derived from a or adds to an or electrophile, typically promoted by (TiCl₄) or boron trifluoride diethyl etherate (BF₃·OEt₂), to yield a silylated aldol adduct that hydrolyzes to the β-hydroxy carbonyl product. This reaction avoids self-condensation issues common in direct enolate aldol additions and proceeds via initial coordination of the Lewis acid to the carbonyl oxygen of the electrophile, followed by nucleophilic attack and silyl group migration to form an oxocarbenium intermediate. A representative example involves the addition of the trimethylsilyl enol ether of acetone to , affording the β-hydroxy 4-hydroxy-4-phenylbutan-2-one in high yield after . \begin{align*} &\ce{CH2=C(OSiMe3)CH3 + PhCHO ->[TiCl4] PhCH(OSiMe3)CH2C(O)CH3} \\ &\ce{->[H3O+] PhCH(OH)CH2C(O)CH3} \end{align*} The Mukaiyama aldol has been extended to asymmetric variants using chiral acids, such as those derived from BINOL-titanium complexes, achieving enantioselectivities exceeding 90% ee for various substrates. In Michael additions, silyl enol ethers serve as carbon nucleophiles in conjugate additions to α,β-unsaturated carbonyl compounds, generating 1,5-dicarbonyl products after desilylation. This Mukaiyama-Michael reaction, often catalyzed by Lewis acids like TiCl₄ or salts, involves activation of the β-carbon of the acceptor, enabling regioselective 1,4-addition. For instance, the silyl enol ether of adds to under BF₃·OEt₂ to produce 2-(3-oxobutyl)cyclohexan-1-one, a key motif in precursors. Asymmetric versions employing chiral -bis() complexes deliver the adducts with up to 98% ee, highlighting the method's utility in stereocontrolled synthesis. Alkylation reactions of silyl enol ethers with allylic halides or epoxides provide another route to C-C bond formation, often under or . With allylic halides, such as , silyl enol ethers undergo SN2'-type allylation promoted by Pd(0) catalysts and ligands, yielding α-allylated carbonyl compounds with high after . A notable example is the palladium-catalyzed coupling of the silyl enol ether of with crotyl chloride, affording the γ-methylated product in 85% yield. Similarly, epoxides react with silyl enol ethers in the presence of TiCl₄ to effect regioselective ring opening at the less substituted carbon, forming β-(silyloxy) alcohols that convert to 1,4-dicarbonyls upon . This transformation yields γ-hydroxy carbonyl compounds efficiently. Stereochemistry in these reactions is critically influenced by the choice of catalyst and enol ether geometry. In the Mukaiyama aldol, an open transition state predominates with most Lewis acids, favoring anti diastereoselectivity due to minimized steric interactions between substituents, as observed in additions yielding >10:1 anti:syn ratios with bulky silyl groups. However, with TiCl₄ or certain chiral auxiliaries, a closed, Zimmerman-Traxler-like six-membered transition state can form, enforcing syn selectivity through chelation and enabling asymmetric induction; for example, (E)-silyl enol ethers of Evans' auxiliaries deliver syn aldols with 95:5 dr and 92% ee using a chiral titanium catalyst. In Michael additions, chiral organocatalysts like bifunctional thioureas promote enantioselective protonation of the enolate intermediate, achieving up to 99% ee in additions to enals. A specialized ring contraction involving silyl enol ether-derived precursors is the Eschenmoser–Tanabe fragmentation of α,β- ketones, where the epoxy ketone is treated with tosylhydrazine to afford alkynyl carbonyl fragments, effectively reducing ring size by one carbon unit through C-C bond reorganization. This method, applied to cyclohexanone-derived epoxy ketones, yields homopropargylic aldehydes in 70-80% overall yield from the initial aldol step.

Electrophilic functionalizations

Silyl enol ethers undergo electrophilic functionalizations where heteroatom-based electrophiles add to the electron-rich , typically at the β-carbon, leading to α-functionalized carbonyl derivatives after . These reactions leverage the nucleophilic of silyl enol ethers, the introduction of , oxygen, , , or at the α-position with high . The resulting intermediates often involve α-silyloxy carbocations that are quenched hydrolytically to regenerate the , avoiding over-functionalization common in direct chemistry. Halogenation is a cornerstone of these transformations, with N-chlorosuccinimide (NCS) or (Br₂) serving as effective electrophiles to afford α-chloro or α-bromo carbonyl compounds, respectively. For instance, treatment of cyclohexyl trimethylsilyl enol ether with NCS in at yields the α-chloroketone in high yield after aqueous , proceeding via initial chloronium addition followed by silyl group departure. Similarly, reaction with Br₂ in provides the α-bromoketone regioselectively, even for unsymmetrical ketones where the less substituted enol ether regiomer is favored under kinetic control. These methods are mild, tolerant of various functional groups, and widely used for preparing α-halo carbonyls as synthons in further chain extensions. Oxidative functionalizations, particularly the Saegusa–Ito reaction, convert silyl enol ethers directly to α,β-unsaturated carbonyls using palladium(II) catalysis. In this seminal process, the silyl enol ether coordinates to Pd(OAc)₂, followed by dehydrosilylation and reoxidation by p-benzoquinone, yielding enones in good yields for both cyclic and acyclic substrates. For example: \ce{(CH3)2C=CH-OSiMe3 ->[Pd(OAc)2, p-BQ][AcOH, rt] (CH3)2C=CH-C(O)-CH3} The reaction is stereospecific, preserving the enol ether geometry in the product, and has been pivotal in syntheses due to its operational simplicity and avoidance of over-oxidation. Sulfenylation introduces at the α-position using sulfenyl s like phenylsulfenyl (PhSCl), generating α-phenylthio carbonyls after . The electrophilic adds to the β-carbon, forming an α-(phenylthio)-α-silyloxy intermediate that hydrolyzes cleanly; yields exceed 80% for ketone-derived silyl enol ethers in ether solvents at low temperatures. Modern variants employ chiral bases for enantioselective sulfenylation, achieving up to 95% with N-(p-tolylsulfinyl)iminoarylsulfuranes as electrophiles, highlighting the versatility for asymmetric . Additional electrophilic aminations utilize azodicarboxylates (e.g., di-tert-butyl azodicarboxylate) as nitrogen sources, often under silver or copper catalysis, to produce α-hydrazino carbonyls that serve as precursors to α-amino ketones. For example, silver triflate-catalyzed reaction of silyl enol ethers with dibenzyl azodicarboxylate in THF affords the α-aminated products in 70–90% yields, with asymmetric induction possible using chiral ligands. These transformations proceed via nucleophilic attack on the azo , followed by silyl transfer and , enabling efficient C–N bond formation. Electrophilic , though less common, can be achieved with dialkyl phosphorochloridates under Lewis acid activation, yielding α-phosphorylated carbonyls, but requires careful control to prevent side reactions. The mechanisms across these reactions share a common motif: the E⁺ (where E is , oxygen, , etc.) attacks the β-carbon of the silyl enol ether, generating a resonance-stabilized α-carbocation bound to the -OSiR₃ group. This intermediate may undergo silyl migration or direct loss of the silyl group during , restoring the carbonyl and installing the at the α-position. For the Saegusa–Ito oxidation, Pd(II) insertion into the C–O bond facilitates β-hydride elimination instead, uniquely yielding the enone. These pathways ensure high , with the silyl group modulating reactivity to prevent or elimination side products.

Hydrolysis and desilylation

Silyl enol ethers undergo under mild acidic conditions to regenerate the parent carbonyl compounds along with the corresponding . Typical protocols employ acetic acid in aqueous (THF/H₂O) or similar solvent mixtures, often proceeding quantitatively for most - and aldehyde-derived substrates. For instance, treatment with 1:1 AcOH/H₂O at silyl enol ethers derived from cyclic in high yields without affecting other acid-sensitive functionalities. This process follows the general equation: \text{R}_2\text{C}=\text{CR-OSiR}_3' + \text{H}_3\text{O}^+ \rightarrow \text{R}_2\text{CH-C(=O)R} + \text{R}_3'\text{SiOH} The mild nature of these conditions makes acidic hydrolysis suitable for deprotecting silyl enol ethers in multi-step syntheses where selective removal is required, such as in the presence of silyl ether protecting groups on alcohols. Basic desilylation of silyl enol ethers is achieved using fluoride ion sources, such as tetrabutylammonium fluoride (TBAF), which cleave the O-Si bond to generate enolates in situ. These enolates, often as quaternary ammonium salts, can be directly utilized in subsequent reactions like alkylations or aldol additions without the need for metal counterions. For example, TBAF in THF at low temperatures converts trimethylsilyl enol ethers to the corresponding enolates in excellent yields, enabling further transformations while avoiding protonation to the carbonyl. This method is particularly valuable for generating reactive enolates under anhydrous conditions. Oxidative desilylation methods transform silyl enol ethers into α,β-unsaturated carbonyl compounds, combining deprotection with dehydrogenation. In the Saegusa-Ito oxidation, palladium(II) acetate catalyzes the reaction in the presence of benzoquinone, affording enones from silyl enol ethers regioselectively and in good yields, as demonstrated in the original report for acyclic and cyclic substrates. These transformations highlight the utility of silyl enol ethers as protected forms in synthetic routes requiring unsaturation, such as total syntheses of natural products.

Use in aldol and Mukaiyama reactions

Silyl enol ethers act as stable nucleophilic equivalents of enolates in the Mukaiyama , enabling the stereoselective formation of β-hydroxy carbonyl compounds through Lewis acid-catalyzed addition to aldehydes and ketones. First reported in 1973, this variant of the addresses limitations of traditional enolate-based methods by preventing self-condensation of the carbonyl partner and accommodating acid-sensitive functional groups under mild conditions. The reaction typically proceeds with high efficiency, often delivering yields exceeding 80% and enabling precise control over based on the of the silyl enol ether precursor. A range of Lewis acids facilitate the Mukaiyama aldol, with titanium tetrachloride (TiCl₄) serving as the original promoter for activating both ketone- and aldehyde-derived silyl enol ethers toward a broad scope of electrophiles, including aliphatic and aromatic aldehydes. Subsequent developments expanded to boron trifluoride diethyl etherate (BF₃·OEt₂) and tin(IV) chloride (SnCl₄), which enhance reactivity with sterically hindered substrates and improve compatibility with functionalized carbonyls, such as α-alkoxy aldehydes common in natural product synthesis. These variants allow for anti-selective diastereocontrol in certain cases, particularly when using chelating Lewis acids or bulky substituents on the silyl enol ether, favoring the unlike (anti) relative stereochemistry via open transition states. Intramolecular Mukaiyama aldol reactions of tethered to aldehydes provide a powerful strategy for constructing 5- or 6-membered carbocycles and heterocycles, exploiting the proximity effect to achieve high and efficiency in ring closure. For instance, under BF₃·OEt₂ catalysis, such directed cyclizations form cyclopentanols or tetrahydropyrans with diastereoselectivities often greater than 10:1, making them valuable for assembling complex frameworks. The utility of in Mukaiyama aldol reactions extends to synthesis, where they enable the iterative construction of chains with defined , as demonstrated in the assembly of precursors through sequential cross-aldol couplings. Stereocontrol is further enhanced by chiral auxiliaries, such as oxazolidinones adapted from Evans' methodology, or chiral Lewis acids like those derived from BINOL, routinely affording products with enantiomeric excesses above 80% in asymmetric variants. These adaptations underscore the reaction's versatility in generating enantioenriched building blocks for bioactive molecules while maintaining compatibility with diverse functional groups.

Applications in total synthesis and recent developments

Silyl enol ethers have played a pivotal role in the of complex s, particularly through stereoselective aldol couplings that enable the construction of polyoxygenated frameworks. These examples highlight how silyl enol ethers enable efficient, convergent strategies for assembling densely functionalized carbon skeletons in natural product . In pharmaceutical applications, silyl enol ethers are employed as versatile intermediates in scalable processes for active pharmaceutical ingredients (APIs), notably statins. The of 7-amino-3,5-dihydroxy-6-heptenoates, key side-chain precursors for inhibitors like simvastatin, utilizes Lewis acid-catalyzed aldol reactions of 8-keto silyl enol ethers to introduce the requisite β-hydroxy acid motif with high diastereoselectivity. This approach supports industrial-scale production by providing robust, regioselective carbon-carbon bond formation under mild conditions, minimizing side reactions in multi-kilogram syntheses. Recent developments from 2015 to 2025 have expanded the utility of silyl enol ethers through innovative catalytic methods. (PET) reactions have enabled radical couplings, allowing silyl enol ethers to participate in single-electron transfer processes for the of diversified carbonyl derivatives under visible-light , as reviewed in applications to cyclopropyl silyl ethers and acetals. In 2019, a photoredox-Brønsted base hybrid system achieved direct allylic C-H of silyl ethers using an and 2,4,6-collidine, generating nucleophilic allylic s for selective C-C bond formation with alkyl halides, broadening access to branched ketones. Catalyst- and additive-free thiofunctionalization emerged in 2024, where silyl enol ethers react with disulfides to form β-keto sulfides in high yields, offering a sustainable route to sulfur-containing motifs without metal mediators. Asymmetric variants have also advanced, with a 2023 cobalt-catalyzed method enabling regio- and stereoselective generation of silyl enol ethers from aldehydes, expanding substrate scope to include aliphatic and aromatic systems for enantioenriched aldol precursors. In 2025, catalytic asymmetric / of silyl enol ethers was reported, providing access to chiral boronic esters via remote functionalization. Additionally, an -catalyzed alkynylogous allylic substitution using alpha-alkynyl silyl enol ethers enabled of enantioenriched alpha-alkynyl ketones as of November 2025. Looking ahead, integration of silyl enol ethers with advanced catalysis promises greener synthesis protocols, such as or iron-based systems, to reduce waste and enable late-stage functionalization in while maintaining high efficiency and selectivity.

Silyl ketene acetals

Silyl ketene acetals are ester-derived analogs of silyl enol ethers, featuring the general structure R₂C=C(OR')OSiR₃, where the silyloxy group is attached to the β-carbon relative to the original ester carbonyl, forming a ketene acetal motif. This structural difference imparts higher nucleophilicity at the α-carbon compared to ketone-derived silyl enol ethers, as the alkoxy substituent stabilizes developing positive charge during electrophilic attack, enabling more efficient reactions with carbonyl electrophiles. Unlike silyl enol ethers from ketones, silyl ketene acetals serve primarily as precursors to carboxylic acid derivatives, facilitating the synthesis of β-functionalized esters. Their synthesis parallels that of silyl enol ethers but starts from enolates generated under strong base conditions, followed by trapping with a silyl chloride. A common method involves deprotonation of an with (LDA) at low temperature, typically -78°C in THF, and subsequent addition of chlorotrimethylsilane (TMSCl) to afford the silyl ketene acetal. This approach, first detailed by Rathke and Sullivan, yields the desired compounds in high efficiency and allows control over E/Z through base and silyl group selection. The general reaction scheme is: \text{R-CH}_2\text{-COOR'} + \text{LDA} + \text{TMSCl} \rightarrow \text{R-CH=C(OSiMe}_3\text{)OR'} + \text{byproducts} This process generates lithium enolates that are silylated on oxygen, avoiding C-silylation under kinetic conditions. In terms of reactivity, silyl ketene acetals exhibit enhanced nucleophilicity, making them suitable for Reformatsky-like additions to aldehydes and ketones, often promoted by Lewis acids such as TiCl₄, to produce β-hydroxy esters with good diastereocontrol. They also participate in Claisen condensations, where crossed variants with methyl esters under NaOH catalysis yield α-monoalkylated β-keto esters selectively, bypassing self-condensation issues common in traditional methods. These transformations highlight their utility in constructing carbon-carbon bonds at the α-position of carboxylic derivatives, with the silyl group providing mild, functional group-compatible activation.

Other silyl-protected enol derivatives

Aldehyde-derived silyl ethers, such as those from , exhibit lower stability compared to their counterparts, often necessitating generation to prevent or side reactions like self-aldolization. These compounds are particularly valuable in stereoselective transformations, including iridium-catalyzed isomerizations that afford fully substituted variants with high Z-selectivity (up to 95:5 Z/E) for subsequent aldol additions or allylations. For instance, cobalt-catalyzed one-carbon extensions enable regio- and stereoselective synthesis from aryl aldehydes, yielding (Z)-silyl ethers in high yields (up to 87%), which serve as precursors for carbon centers in complex molecules. Amide and imine analogs, known as N-silyl enol ethers or silyl ketene aminals, represent nitrogen-protected variants that display reduced nucleophilicity relative to oxygen-based silyl enol ethers due to the lower electron-donating ability of nitrogen. These compounds are widely employed in the of heterocycles, such as in the of marine alkaloids like palau'amine, where silyl ketene aminals facilitate selective alkylations and cyclizations under mild conditions. Their lower reactivity enables precise control in group transfer polymerizations and conjugate additions. Specialized C-silyl enol ethers, or vinylsilanes, function as carbon-silylated analogs and serve as key precursors in the , where β-hydroxysilanes undergo syn-elimination to form alkenes with controllable . These vinylsilanes exhibit utility in cross-coupling reactions, such as the Hiyama-Denmark coupling, enabling Pd-catalyzed C-C bond formation with aryl or vinyl halides in aqueous media with high efficiency. Compared to O- or N-protected variants, C-silyl enol ethers offer enhanced stability toward but lower inherent nucleophilicity at the vinyl position, making them suitable for selective functionalizations in natural product synthesis.

References

  1. [1]
    [PDF] Silyl Enol Ethers - S. Kobayashi, K. Manabe, H. Ishitani, and J.
    Apr 4, 2016 · most important application of silyl enol ethers in organic synthesis. Under the influence of Lewis acids, silyl enol ethers react with not ...
  2. [2]
    Reactive enolates from enol silyl ethers - ACS Publications
    Rhenium-Catalyzed Decarboxylative Coupling of Cyclic Enol Carbonates with Silyl Enol Ethers and Ketene Silyl Acetals.
  3. [3]
    Silyl enol ether synthesis by silylation - Organic Chemistry Portal
    Silyl enol ethers can be synthesized using ionic liquids, ruthenium complexes, N-heterocyclic carbenes, Ni-catalyzed chain walking, and salt-free ...
  4. [4]
  5. [5]
    Trimethyl[(1-methylethenyl)oxy]silane - CAS Common Chemistry
    Compound Properties. Boiling Point (1). 94-95 °C. Melting Point (2). 40-45 °C. Density (1). 0.786 g/cm³. Source(s). (1) Metal-Organics Catalog physical property ...
  6. [6]
  7. [7]
    [PDF] 2-METHYL-1-(TRIMETHYLSILOXY)-1-PROPENE - Gelest, Inc.
    Jan 24, 2017 · Boiling point. : 82 - 86 °C @ 100 mm Hg. Flash point. : 14 °C. Auto ... Solubility. : Insoluble in water. Reacts slowly with water. Log Pow.
  8. [8]
    Silyl Enol Ether Prins Cyclization: Diastereoselective Formation of ...
    Sep 9, 2014 · The cyclization step forms new carbon–carbon and carbon–oxygen bonds, as well as a quaternary center with good diastereoselectivity. The method ...
  9. [9]
    Expanding the chemical space of enol silyl ethers - RSC Publishing
    Oct 20, 2023 · Enol silyl ethers are versatile, robust, and readily accessible substrates widely used in chemical synthesis. However, the conventional ...
  10. [10]
    Impact of Silyl Enol Ether Stability on Palladium-Catalyzed Arylations
    Mar 16, 2010 · Another important feature of this reaction is the loading of the catalyst that was reduced with the increased stability of the silyl enol ether.
  11. [11]
    [(1-Methoxyhex-1-en-1-yl)oxy](trimethyl)silane - Vulcanchem
    IR (KBr, cm⁻¹): Strong absorption at 1640–1680 (C=C stretch). Peaks at 1250 (Si ... Silyl enol ether derivatives are explored as precursors for silicone ...
  12. [12]
    A highly diastereoselective “super silyl” governed aldol reaction
    Oct 6, 2015 · Iodobenzene was then added to the NMR tube and a second 29Si NMR spectrum was recorded a er 45 min at room ... silyl enol ether by substitution of ...
  13. [13]
    Chemistry of carbanions. XVIII. Preparation of trimethylsilyl enol ethers
    A cooperative cobalt-driven system for one-carbon extension in the synthesis of (Z)-silyl enol ethers from aldehydes: unlocking regio- and stereoselectivity.
  14. [14]
    Synthesis of a Silyl Enol Ether Using Trimethylsilyl Triflate
    We are proud to present the synthesis of the silyl enol ether from α-tetralone using trimethylsilyl triflate. ... Trimethylsilyl Triflate (= TMSOTf) [T0871] · α- ...
  15. [15]
    Silyl ketone chemistry. A new regiospecific route to silyl enol ethers
    ... silyl enol ethers through brook rearrangement of the 1,2-addition products. Tetrahedron Letters 1990, 31 (6) , 831-834. https://doi.org/10.1016/S0040-4039 ...
  16. [16]
    Product Subclass 16: Silyl Enol Ethers - Thieme E-Books
    Silyl enol ethers can be prepared readily from the parent carbonyl compounds by silylation of the corresponding enolate anions. In the synthesis of silyl enol ...<|control11|><|separator|>
  17. [17]
  18. [18]
  19. [19]
    New cross-aldol reactions. Reactions of silyl enol ethers with ...
    New cross-aldol reactions. Reactions of silyl enol ethers with carbonyl compounds activated by titanium tetrachloride.
  20. [20]
    Mukaiyama Aldol Addition - Organic Chemistry Portal
    The use of silyl enol ethers as an enolate equivalent in Lewis acid-catalyzed aldol additions. The trimethylsilyl group is thought of as a sterically demanding ...<|control11|><|separator|>
  21. [21]
    Mukaiyama aldol reaction: an effective asymmetric approach to ...
    Nov 8, 2023 · The Mukaiyama aldol reaction is generally a Lewis-acid catalyzed cross-aldol reaction between an aldehyde or ketone and silyl enol ether.
  22. [22]
    Mukaiyama Michael Reaction - an overview | ScienceDirect Topics
    The Mukaiyama–Michael reaction is defined as a reaction involving the transfer of a silyl group from a silyl enol ether to a carbonyl group of an enone, ...Missing: original | Show results with:original
  23. [23]
    The Michael reaction of silyl enol ethers or ketene silyl acetals with ...
    The Michael reaction of silyl enol ethers or ketene silyl acetals with conjugated nitro olefins activated by Lewis acids: new synthesis of 1,4-diketones and .Missing: original | Show results with:original
  24. [24]
    Catalytic asymmetric defluorinative allylation of silyl enol ethers
    May 24, 2023 · We report a new catalytic methodology able to construct α-allyl ketones via defluorinative allylation of silyl enol ethers in a regio-, diastereo- and ...Missing: halides | Show results with:halides
  25. [25]
    The Allylic Alkylation of Ketone Enolates - PMC - NIH
    Sep 10, 2020 · 11 One of the advantages of silyl enol ethers is that the reaction conditions can in principal be maintained completely neutral, thus allowing ...
  26. [26]
    promoted reaction of trimethylsilyl enol ethers with epoxides
    Epoxide ring opening with organometallic nucleophiles is a well-known reaction and a synthetically useful method of C–C bond formation.1 However, the ...
  27. [27]
    Carbon-carbon bond formation in reactions of PhIO.cntdot.HBF4 ...
    Carbon-carbon bond formation in reactions of PhIO.cntdot.HBF4-silyl enol ether adduct with alkenes or silyl enol ethers. Click to copy article linkArticle ...
  28. [28]
    4.4: The aldol reaction - Chemistry LibreTexts
    Aug 15, 2020 · The Mukaiyama aldol reaction proceeds through an open transition state, which can be represented in a newman projection.Enolate Geometry in Aldol... · Boron aldol reaction · Absolute stereochemistry
  29. [29]
    Synthetic methods. IV. Halogenation of carbonyl compounds via silyl ...
    Reactivity of electrophilic reagents towards silyl enol ether groups ... Halogenation of PGI2-enol ether, with N-halosuccinimide: Synthesis of new ...
  30. [30]
    Reactions of enol silyl ethers with N-halosuccinimide - a stepwise ...
    General experimental procedure: To a stirred, refluxing solution of 5.0 g of the enol silyl ether in 150 ml CH2Cl2, an equivalent amount of NCS in 250 ml CH2Cl2 ...Missing: primary | Show results with:primary
  31. [31]
    Catalytic, Enantioselective Sulfenylation of Ketone-Derived ...
    Sep 5, 2014 · ... silyl enol ether substrates, a sulfenylating agent tha ... (31d) The configurational stability of III is unknown, but its ...
  32. [32]
    Silver-catalyzed asymmetric amination of silyl enol ethers
    In a mixture of toluene or mesitylene and THF, silyl enol ethers reacted with dibenzyl azodicarboxylate (DBnAD) smoothly to afford the corresponding amination ...
  33. [33]
    Total synthesis and development of bioactive natural products - NIH
    Total synthesis of erythromycin A. We also developed several other ... enol silyl ether and an O-silyl secondary alcohol. The SnCl4-promoted aldol ...
  34. [34]
    Total Syntheses of Vancomycin and Eremomycin Aglycons - Evans
    Oct 16, 1998 · Controlling the elements of planar and axial chirality are the principal challenges in the synthesis of the aglycon of vancomycin.
  35. [35]
    3-Hydroxy-3-methylglutaryl-coenzyme A reductase inhibitors. 9. The ...
    The synthetic strategy employed epoxidation or Lewis acid-catalyzed aldol reaction of the 8-keto silyl enol ether as a key reactive intermediate.
  36. [36]
    The photoinduced electron transfer (PET) reactions of silyl ethers ...
    The E1/2ox of a typical trimethylsilyl enol ether is usually 1 V lower than that of a simple trimethylsilyl ether. On the basis of this difference, in 1988, ...Missing: IUPAC name
  37. [37]
    Direct allylic C–H alkylation of enol silyl ethers enabled by ... - Nature
    Jun 20, 2019 · Catalytic asymmetric C−H activation of silyl enol ethers as an equivalent of an asymmetric Michael reaction. J. Am. Chem. Soc. 123, 2070 ...
  38. [38]
    Thiofunctionalization of Silyl Enol Ether: An Efficient Approach ... - NIH
    Herein, we report a highly efficient synthesis of β-keto sulfides via a catalyst- and additive-free reaction between silyl enol ethers and sulfur-containing ...
  39. [39]
    Silyl Enol Ethers from Aldehydes: Unlocking Regio- and ...
    Dec 12, 2023 · A variety of cyclic silyl enol ethers can be used, and the yields of the desired products range from 52 to 93%, while the enantiomeric excesses ...
  40. [40]
    Expanding the chemical space of enol silyl ethers - RSC Publishing
    Herein we describe a mild, fast, and operationally simple one-step protocol that combines readily available fluoroalkyl halides, silyl enol ethers, and, for ...
  41. [41]
    The chemistry of O-silylated ketene acetals. Stereocontrolled ...
    The chemistry of O-silylated ketene acetals. Stereocontrolled synthesis of 2-deoxy- and 2-deoxy-2-C-alkyl-erythro-pentoses.Missing: stabilization | Show results with:stabilization
  42. [42]
    NaOH-catalyzed crossed Claisen condensation between ketene ...
    We have developed a practical crossed Claisen condensation between ketene silyl acetals and methyl esters using catalytic NaOH to obtain α-monoalkylated β-keto ...
  43. [43]
    Recent developments in organic synthesis for constructing carbon ...
    Mar 19, 2025 · ... stability and reaction reversibility. ... Marek, Stereoselective Access to Fully Substituted Aldehyde-Derived Silyl Enol Ethers by Iridium- ...
  44. [44]
    [PDF] synthesis, study and application of silyl ketene imines in lewis - CORE
    Chapter 1: Theory and Applications of Lewis Base Catalysis in Organic Synthesis ... Addition of amide derived silyl ketene aminal 111 to hydrocinnamaldehyde. (c) ...
  45. [45]
    Total synthesis of palau'amine | Nature Communications
    Nov 4, 2015 · Treatment of 12 with triethylsilyl trifluoromethanesulfonate and 2,6-di-tert-butyl pyridine at −78 °C readily afforded silyl ketene aminal ...
  46. [46]
    Organocatalytic Group Transfer Polymerization of α-Methylene-N ...
    In the GTP of α-MMP and SKAR, the silyl ketene aminal derived from α-MMP (SKA MMP R) was initially formed by the Mukaiyama–Michael addition between α-MMP and ...Introduction · Results And Discussion · Organocatalytic Gtp Of...<|control11|><|separator|>
  47. [47]
    Peterson Olefination - Organic Chemistry Portal
    The Peterson Olefination offers the possibility of improving the yield of the desired alkene stereoisomer by careful separation of the two diastereomeric β- ...
  48. [48]
    Peterson Olefination - an overview | ScienceDirect Topics
    Silyl-substituted enol ethers may be accesssed from vinylsilanes by an addition–elimination sequence <1997TL6763>. Alcohols and iodine may be added to ...
  49. [49]
    Reactions of an Isolable Dialkylsilylene with Ketones † | Request PDF
    Aug 6, 2025 · ... silirane, sila‐aza/‐phospha/‐oxa/‐thia cyclopropanes. The Si ... rearrange to the corresponding silyl and disilyl enol ethers, respectively.