Fact-checked by Grok 2 weeks ago

Ether cleavage

Ether cleavage is a fundamental reaction in organic chemistry involving the breaking of the carbon-oxygen bond in an ether (R–O–R'). The most common method is acidic cleavage under strongly acidic conditions, yielding an alcohol and an alkyl halide. Other approaches include base-mediated cleavage with strong nucleophiles like organolithium or Grignard reagents, and alternative methods using Lewis acids, as detailed in later sections. Acidic cleavage typically requires concentrated strong acids such as hydrobromic acid (HBr) or hydroiodic acid (HI), with heating, and proceeds via protonation of the ether oxygen to generate a good leaving group (a protonated alcohol). Unlike milder acids like HCl, which are ineffective due to the poor nucleophilicity of chloride ion, HI and HBr succeed because iodide and bromide are better nucleophiles, and the reactions often involve excess acid to convert the initial alcohol product into a second alkyl halide. The mechanism of acidic ether cleavage depends on the nature of the alkyl groups attached to the oxygen and follows either an SN2 or SN1 pathway. In the SN2 mechanism, common for ethers with primary or methyl alkyl groups, the ion attacks the less sterically hindered carbon, displacing the protonated leaving group in a concerted fashion; for example, reacts with HBr to form and initially, with excess HBr yielding two equivalents of . Conversely, the SN1 mechanism predominates for , benzylic, or allylic alkyl groups, where leads to formation and subsequent trapping by ; a classic case is tert-butyl methyl with HI, producing and 2-iodo-2-methylpropane, as the is stable. Secondary ethers may exhibit mixed mechanisms, while elimination (E1) can compete under certain conditions, such as with groups and weaker nucleophiles, yielding alkenes alongside . Regioselectivity in unsymmetrical ether cleavage is governed by the mechanism: SN2 favors attack at the less substituted carbon, while SN1 directs the halide to the more stable carbocation site. Aryl alkyl ethers, such as anisole (methoxybenzene), cleave to phenols and alkyl halides because the aryl-oxygen bond resists breaking due to partial double-bond character from resonance, preventing aryl halide formation. Diaryl ethers, like diphenyl ether, are particularly stable and do not undergo cleavage under acidic conditions. This reaction's utility extends to synthetic applications, such as deprotecting alcohols masked as ethers, though it highlights ethers' general inertness under neutral or basic conditions, making them valuable solvents and protecting groups in organic synthesis.

General concepts

Definition and scope

Ether cleavage refers to the chemical transformation involving the rupture of one or more carbon-oxygen (C-O) bonds in an , represented generally as R-O-R', where R and R' are alkyl or aryl groups, resulting in products such as alcohols, , or alkyl halides. Ethers exhibit high due to their structure, featuring sp³ hybridized C-O bonds and two lone pairs on the oxygen atom, which allow for basicity but resist nucleophilic attack under mild conditions, often necessitating strong acids or organometallic for cleavage. The scope of ether cleavage spans various ether classes, including dialkyl ethers, which typically yield an and an alkyl halide, or two alkyl halides with excess reagent, depending on reaction conditions; alkyl aryl ethers, which produce a and an alkyl derivative; and diaryl ethers, which are notably resistant to standard cleavage and require specialized methods such as catalytic or electrochemical processes. For instance, symmetric dialkyl ethers like can cleave to derivatives, while unsymmetric variants exhibit selectivity based on the nature of the alkyl groups involved. This reaction holds significant importance in , particularly for deprotecting hydroxyl groups masked as , such as in the removal of benzyl or methyl protecting groups during multi-step syntheses of complex molecules. Additionally, it facilitates the preparation of from alkyl aryl and serves analytical purposes in identifying structures through controlled degradation.

Factors influencing reactivity

The reactivity of ethers toward cleavage is profoundly influenced by steric and electronic factors associated with the alkyl or aryl substituents attached to the oxygen atom. Primary alkyl ethers predominantly undergo cleavage via an SN2 pathway due to minimal steric hindrance at the carbon center, allowing direct nucleophilic attack, whereas tertiary alkyl ethers favor an SN1 pathway because of significant steric bulk that promotes formation after departure of the . In aryl alkyl ethers, cleavage occurs selectively at the alkyl C-O bond rather than the aryl C-O bond, because the aryl C-O bond is strengthened by , exhibiting partial double-bond character. The choice of reagent plays a critical role in determining the efficiency and products of ether cleavage. Strong acids such as and HBr are particularly effective for promoting formation, with exhibiting higher reactivity than HBr due to the greater nucleophilicity of the . further modulate reactivity; polar protic solvents stabilize ionic intermediates and transition states in SN1 pathways, enhancing cleavage for ethers, while polar aprotic solvents accelerate SN2 processes by solvating cations without hydrogen bonding to the , thus preserving its reactivity for primary ethers. Thermodynamically, the C-O in ethers has a of approximately 350 /, rendering cleavage endothermic under standard conditions and requiring to proceed. of the ether oxygen or coordination with acids lowers the barrier by converting the poor () into a more stable oxonium or coordinated , thereby facilitating bond breaking. Certain structural motifs impose limitations on ether cleavage reactivity. Symmetrical diaryl ethers, such as , rarely undergo cleavage under standard acidic conditions due to the low basicity of the oxygen and the high stability of the aryl C-O bonds, often requiring forcing conditions like alkali metals in liquid ammonia. Similarly, cyclic ethers like (THF) exhibit reduced reactivity and demand elevated temperatures or concentrated reagents to achieve cleavage, typically yielding 1,4-dihalobutanes with hydrogen halides.

Acid-mediated cleavage

Protonation and general mechanism

In acid-mediated ether cleavage, the reaction begins with the of the ether's oxygen by a strong acid, such as (HBr) or (HI). This step generates an alkyloxonium ion intermediate, denoted as \ce{R-O^+H-R'}, which polarizes the C-O bonds and activates the ether toward by rendering the moiety a viable . Following , the general proceeds with nucleophilic attack by the conjugate (\ce{Br^-} or \ce{I^-}) on one of the alkyl carbons, displacing the neutral . The overall transformation is summarized by the equation: \ce{R-O-R' + HX -> R-X + R'-OH} (or the reverse, \ce{R-OH + R'-X}, based on structural factors). This cleavage exhibits , where the preferentially bonds to the less sterically hindered alkyl group or the one capable of forming a more stable intermediate. These reactions are typically conducted under conditions with concentrated aqueous HX to ensure complete conversion, as the initially formed may further react to yield a second equivalent of alkyl halide. is particularly favored for cleaving methyl ethers owing to the superior nucleophilicity of compared to bromide. The pathway after may follow either an SN2 or SN1 route depending on the alkyl substituents involved.

SN2 pathway for methyl and primary alkyl ethers

In the SN2 pathway for the acid-mediated cleavage of methyl and primary alkyl ethers, the reaction begins with protonation of the ether oxygen atom by the strong acid, such as HI or HBr, to form a resonance-stabilized oxonium ion intermediate. This protonation enhances the electrophilicity of the alkyl carbon attached to the oxygen, making it susceptible to nucleophilic attack. The mechanism then proceeds via a concerted bimolecular nucleophilic substitution (SN2), where the halide ion (e.g., I⁻) performs a backside attack on the primary or methyl carbon, leading to inversion of configuration at that carbon and simultaneous displacement of the neutral alcohol leaving group. A representative example is the cleavage of with , which yields methyl iodide and . The overall reaction can be depicted as follows, with the protonated shown for clarity: \ce{CH3-O-CH3 + HI ->[H+] [CH3-OH-CH3]+ I- -> CH3-I + CH3-OH} This pathway is favored for unhindered ethers due to the low steric hindrance at the primary carbon, allowing efficient backside approach by the . The kinetics of this SN2 process are second-order overall, first-order in both the protonated ether concentration and the halide nucleophile, reflecting the bimolecular rate-determining step. The reaction rate increases with nucleophile strength, following the order I⁻ > Br⁻ > Cl⁻, as iodide is a better nucleophile in protic solvents; HCl typically requires harsher conditions for effective cleavage. The protonated alcohol serves as an excellent leaving group, further facilitating the substitution. Another illustrative case is the cleavage of diethyl ether with HI, producing iodoethane and ethanol via SN2 attack at one of the primary ethyl carbons.

SN1 pathway for secondary and tertiary alkyl ethers

The acid-mediated cleavage of secondary and alkyl ethers proceeds via an SN1 pathway, characterized by the formation of a intermediate after initial of the ether oxygen. Strong acids such as HBr or protonate the oxygen atom, generating a good in the form of an alkyloxonium (\ce{R2OH^{+}-R'}). This undergoes unimolecular dissociation, cleaving the C-O bond adjacent to the carbon that can best stabilize the positive charge, thereby producing an and a secondary or . The is then captured by the nucleophilic to yield the corresponding alkyl . This dissociative is predominant for secondary and ethers because the of the intermediate lowers the for bond breaking. Carbocation involvement in the SN1 pathway allows for potential rearrangements, such as or alkyl shifts, to form more stable intermediates during the dissociation step. For instance, the cleavage of isopropyl methyl with concentrated HBr yields isopropyl and , proceeding through a secondary isopropyl . A more pronounced example is the reaction of tert-butyl methyl with HBr, where the protonated species rapidly dissociates to the highly stable tertiary and , followed by attack to form . The overall process can be represented as: \ce{(CH3)3C-O-CH3 + HBr ->[(1) protonation] (CH3)3C-OH^{+}-CH3 \cdot Br^{-}} \ce{(CH3)3C-OH^{+}-CH3 ->[(2) dissociation] (CH3)3C^{+} + CH3OH} \ce{(CH3)3C^{+} + Br^{-} ->[(3) nucleophilic attack] (CH3)3CBr} The reaction rate follows first-order kinetics with respect to the concentration of the protonated ether, as the rate-determining step is the formation of the carbocation. Rate constants demonstrate a clear dependence on carbocation stability, with tertiary alkyl ethers reacting significantly faster than secondary ones (e.g., relative rates of approximately 10^5:1 for tertiary vs. secondary ethers in HI cleavage at 100°C). This order—tertiary > secondary—arises from the hyperconjugative and inductive stabilization of the carbocation by alkyl substituents. In unsymmetrical ethers, regioselectivity favors halide attachment to the more substituted carbon due to preferential carbocation formation at that site.

Base-mediated cleavage

Reaction with organolithium and Grignard reagents

Ether cleavage using organolithium and Grignard reagents represents a base-mediated approach that exploits the strong nucleophilicity of these organometallics to break C-O bonds under aprotic conditions. Common reagents include (n-BuLi) and (PhMgBr), typically employed in ethereal solvents such as or (THF). For Grignard reagents, reactions often require elevated temperatures of 160–180°C to achieve reasonable rates, while organolithium compounds react more readily at milder conditions, such as 0–50°C, sometimes with excess reagent to drive completion. These conditions facilitate selective cleavage, particularly at the less sterically hindered or more electrophilic carbon center. The scope of this reaction is broadest for activated acyclic ethers, including allyl, benzyl, and alkyl aryl ethers, where the benzylic or allylic positions enhance reactivity through stabilization of intermediates. For instance, (methyl phenyl ether) undergoes efficient cleavage with methylmagnesium iodide at 170°C to yield phenol in 78% yield, demonstrating high selectivity for the alkyl-oxygen bond in unsymmetrical cases. Simple dialkyl ethers, such as , exhibit lower reactivity and typically require excess base, prolonged heating, or strained variants for practical cleavage; however, allyl and benzyl dialkyl ethers proceed more readily due to their activated nature. Strained cyclic ethers like epoxides are highly susceptible but are often cleaved under similar conditions, though the focus remains on acyclic systems to avoid competing ring-opening pathways. Competing pathways, such as reductive cleavage, can produce alkanes from the organometallic (e.g., from n-BuLi) alongside alkenes and alkoxides, particularly in ethereal solvents like THF, which may decompose slowly. Products from these cleavages generally consist of an salt and an or derivative, depending on the pathway and ether structure. In a representative example, treatment of phenetole () with amylmagnesium bromide at 160°C produces phenol (as magnesium phenoxide) and , reflecting scission at the alkyl-oxygen bond via elimination. With organolithium reagents, such as n-BuLi reacting with benzyl methyl ether, the outcome often includes the benzyloxide and (from ), with possible methoxide or other fragments. These transformations are valuable for deprotecting allyl or benzyl groups in , yielding clean precursors upon aqueous . Historically, Grignard reagents were first reported to cleave ethers in 1910 by himself, using phenetole at high temperatures to generate . Organolithium-mediated cleavages followed in the 1930s, with Ziegler's work on , but systematic studies of THF cleavage by organolithiums emerged in the 1950s, highlighting solvent decomposition pathways that informed safer handling in .

Mechanism and

In base-mediated ether cleavage using organolithium or Grignard reagents, the process can proceed via alpha-deprotonation followed by elimination or direct C-O bond cleavage through coordination to the oxygen atom. The alpha-deprotonation pathway begins with the strong organometallic base abstracting a proton from the carbon alpha to the oxygen, forming a carbanion intermediate [R-CH(-)-O-R'] M+ (M = Li or MgBr). This carbanion subsequently undergoes elimination of the alkoxide, yielding an alkene (R=CH2) and the metal-bound alkoxide (M-O-R'). This route is particularly favored for ethers bearing activated alpha-hydrogens, such as those in allylic or benzylic positions, where the carbanion is stabilized. A representative for the deprotonation-elimination pathway is: \ce{R-CH2-O-R' + R''M -> [R-CH(-)-O-R']M+ + R''H} \ce{[R-CH(-)-O-R']M+ -> R=CH2 + M-O-R'} The overall result after aqueous is the R=CH2 and R'OH. The direct C-O cleavage involves coordination of the metal center ( or ) to the ether oxygen, which polarizes the C-O bond and facilitates nucleophilic attack by the organometallic alkyl group on the ether's carbon atom or elimination. This displaces the coordinated as a in a concerted manner, often via SN2 for primary or E2 for beta-H present. Unlike acid-mediated cleavage, where protonation activates the C-O bond for direct nucleophilic substitution, the base-mediated mechanism primarily targets the adjacent C-H bond in the deprotonation-elimination pathway for activated ethers, avoiding direct C-O attack unless coordination enables it; the deprotonation route is thus preferred for activated ethers. Regarding stereochemistry, the direct coordination-mediated cleavage typically proceeds with inversion of configuration at the attacked carbon due to the SN2-like backside attack. However, pathways involving carbanion intermediates from deprotonation can lead to racemization due to planar carbanions.

Alternative cleavage methods

Lewis acid-mediated cleavage

Lewis acid-mediated cleavage of ethers employs reagents such as (BBr₃) or (BCl₃), which activate the ether oxygen through coordination rather than , enabling selective C-O scission under milder conditions than protic acids. This approach is particularly effective for deprotecting aryl alkyl ethers, where the Lewis acid forms a with the oxygen, polarizing the C-O and facilitating nucleophilic attack by halide ions on the alkyl group. The begins with the coordination of BBr₃ to the oxygen, generating an that weakens the adjacent C-O . Subsequent of a by another equivalent of BBr₃ produces a cationic oxonium intermediate and tetrabromoborate (BBr₄⁻). This is followed by an SN2-like nucleophilic attack by from BBr₄⁻ on the alkyl carbon, displacing the aryloxyborane and yielding an alkyl . For example, in the demethylation of , the reaction proceeds as follows: \mathrm{Ph-O-CH_3 + BBr_3 \rightarrow [Ph-O-CH_3 \cdot BBr_3] \rightarrow Ph-O-BBr_2 + CH_3Br} Subsequent hydrolysis of the borane intermediate liberates the phenol. This process can cycle up to three times per BBr₃ molecule, allowing substoichiometric use (e.g., 0.33 equiv) for efficient cleavage at elevated temperatures like 100°C. BBr₃ exhibits high selectivity for methyl groups over ethyl or longer alkyl chains in mixed aryl alkyl ethers, owing to the favorable SN2 transition state for primary carbons, and preferentially cleaves aryl alkyl ethers over dialkyl ethers due to the stability of the resulting aryloxyborane. Yields for methyl ether cleavage typically range from 75-85%. In carbohydrate chemistry, BBr₃ enables exhaustive demethylation of permethylated sugars, such as 2,3,4,6-tetra-O-methylglucopyranose, at low temperatures (-80°C) in dichloromethane, cleaving methyl ethers without disrupting glycosidic bonds—a selectivity not achievable with protic acids like HCl. These methods offer advantages over traditional HI or HBr cleavage, including compatibility with sensitive functional groups, reduced harshness, and applicability in complex syntheses requiring orthogonal deprotection.

Cleavage in synthetic applications

Ether cleavage serves as a vital deprotection strategy in , particularly for liberating from methoxyarenes and alcohols from benzyl ethers. (BBr₃) is widely employed for the demethylation of aryl methyl ethers, proceeding under mild conditions at or below to yield without disrupting sensitive functionalities such as groups or diphenyl ethers. For instance, the demethylation of 3,3'-dimethoxybiphenyl affords 3,3'-dihydroxybiphenyl in 77–86% yield after extraction and recrystallization. Similarly, (HBr) in a two-phase system with as a enables selective debenzylation of ortho-substituted phenol benzyl ethers, providing the corresponding under conditions. In , ether facilitates key transformations in complex natural product assemblies, such as the aglycon. During Boger's synthesis, aluminum (AlBr₃) with cleaves methyl ethers in late-stage intermediates, reducing rotational barriers to yield the desired upon thermal equilibration. Reductive using naphthalenide offers a chemoselective alternative for debenzylating benzyl ethers to alcohols, achieving high yields (typically >90%) and operational simplicity in THF at low temperatures while tolerating esters, acetals, and alkenes. This method has been applied in and syntheses to unmask hydroxyl groups without affecting other protecting groups. Selectivity remains a critical challenge in multi-step syntheses involving orthogonal protection schemes, where ether cleavage must target specific groups amid diverse functionalities. Para-siletanylbenzyl (PSB) ethers demonstrate compatibility with methoxymethyl (MOM) and tert-butyldimethylsilyl () ethers, allowing selective oxidative cleavage of PSB under alkaline conditions while leaving MOM and intact, as shown in competition experiments yielding >95% selectivity. Such orthogonality minimizes side reactions, though yields can drop to 70–80% in sterically hindered substrates due to incomplete cleavage or byproduct formation. Recent advancements emphasize catalytic and greener methods for ether cleavage, aligning with sustainable goals. Heterogeneous catalysts enable hydrolytic cleavage of aromatic C–O bonds in alkyl aryl ethers at 160°C in , achieving >80% selectivity toward phenol formation with minimal over-reduction. For model β-O-4 ethers, Pd/C facilitates mild C–O bond scission using as hydrogen donor in air, producing in 92–98% yields and promoting recyclability for valorization. Visible-light with organic dyes debenzylates benzyl ethers in at , delivering alcohols in 80–95% yields while avoiding harsh acids or metals.

References

  1. [1]
    18.3: Reactions of Ethers - Acidic Cleavage - Chemistry LibreTexts
    Mar 19, 2024 · The most common reaction of ethers is cleavage of the C–O bond by using strong acids. During acidic cleavage the ether oxygen is protonated to form a good ...
  2. [2]
    Cleavage Of Ethers With Acid - Master Organic Chemistry
    Nov 19, 2014 · Ether cleavage with acid always begins with protonation of the ether oxygen, followed by either SN1 or SN2 reaction depending on the ...
  3. [3]
    Reactions of Ethers-Ether Cleavage - Chemistry Steps
    The ether cleavage is a substitution reaction where the OR group is replaced with a halogen by converting it into a good leaving group first. This leaving group ...
  4. [4]
  5. [5]
    Electrochemical C–O bond cleavage of diaryl ethers - RSC Publishing
    Selective and efficient cleavage of diaryl ether C–O bonds is significant and challenging for biomass conversion and organic synthesis.
  6. [6]
    18.3 Reactions of Ethers: Acidic Cleavage - Organic Chemistry
    Sep 20, 2023 · Acidic ether cleavages are typical nucleophilic substitution reactions and take place by either SN1 or SN2 mechanisms, depending on the ...
  7. [7]
    [PDF] Copper-catalyzed cleavage of benzyl ethers with ...
    The conversion of alcohols or phenols to ethers and their subsequent cleavage is very important in organic synthesis, particularly in multi-step synthesis of ...
  8. [8]
    Selective demethylation reactions of biomass-derived aromatic ether ...
    Oct 25, 2023 · Moreover, ether cleavage is known to be feasible with alkali organometallic compounds or alkali metals.
  9. [9]
    The Cleavage of Ethers. | Chemical Reviews - ACS Publications
    From 1,2-dialkoxyalkanes to 1,4-dioxanes. A transformation mediated by NbCl5via multiple C–O bond cleavage at room temperature.
  10. [10]
    The Role of Solvent in SN1, SN2, E1 and E2 Reactions
    To compare the effect of polar protic and aprotic solvents, we can say that the protic solvent puts the nucleophile in a cage, thus making it weaker, while the ...
  11. [11]
    Common Bond Energies (D - Wired Chemist
    Common Bond Energies (D. ) and Bond Lengths (r). Hydrogen. Bond, D (kJ/mol), r (pm). H-H, 432, 74. H-B, 389, 119. H-C, 411, 109. H-Si, 318, 148. H-Ge, 288, 153.
  12. [12]
    Ch16 : ROR + HX -> ROH + RX - University of Calgary
    Alkyl ethers are cleaved by the strong acids HI or HBr in a nucleophilic substitution reaction similar to that of alcohols. Protonation of the ethereal oxygen ...
  13. [13]
    [PDF] Chapter 18: Ethers and Epoxides; Thiols and Sulfides
    Ether bonds can be broken with strong acid (HI or HBr) and heat. The mechanism involves protonation of the ether followed by an SN2 reaction with the halide ...
  14. [14]
    THE CLEAVAGE OF SOME ETHERS BY ORGANOLITHIUM ...
    THE CLEAVAGE OF SOME ETHERS BY ORGANOLITHIUM COMPOUNDS. Click to copy ... Journal of Organometallic Chemistry 1988, 352 (1-2) , 1-46. https://doi.org ...
  15. [15]
    200 Years of Lithium and 100 Years of Organolithium Chemistry
    Dec 21, 2017 · In all solvents n-butyllithium was the most stable, tert-butyllithium showed the fastest progress of ether cleavage. As shown above, solutions ...
  16. [16]
    α- versus β-Elimination in the Cleavage of Ethers by Organoalkali ...
    Mechanism of Cleavage of Diethyl Ether by Alkyllithium Compounds: Differing Mechanisms of Decomposition of α‐ and β‐Deuterated Diethyl Ether. Angewandte ...Missing: deprotonation | Show results with:deprotonation
  17. [17]
  18. [18]
    Organic Syntheses Procedure
    ### Summary of BBr3 Demethylation of Aryl Methyl Ethers (CV5P0412)
  19. [19]
    A Simple Debenzylation of O-Substituted Phenol Ethers Using ...
    HBr in presence of tetrabutylamrnoniurn bromide the product obtained was the debenzylated product 2b (Fig. 1). To study the utility of this reagent various ...
  20. [20]
    [PDF] STUDIES TOWARDS THE TOTAL SYNTHESIS OF VANCOMYCIN ...
    Cleavage of the methyl ethers by AlBr3 and EtSH reduced the rotational barrier of the AB axis sufficiently and afforded the correct atropisomer 1.26 on thermal ...
  21. [21]
    Reductive cleavage of benzyl ethers with lithium naphthalenide. A ...
    Mar 31, 1997 · An operationally simple, high-yielding and highly chemoselective procedure has been developed for the conversion of benzyl ethers to the corresponding alcohols.
  22. [22]
    Synthesis, Reactivity, and Indirect Oxidative Cleavage of para ...
    Cleavage of the PSB ether is accomplished under mild conditions—involving alkaline hydrogen peroxide—that are unique among cleavage protocols for arylmethyl ...Missing: definition | Show results with:definition
  23. [23]
    [PDF] Palladium‐Catalyzed Hydrolytic Cleavage of Aromatic C−O Bonds
    Jan 18, 2017 · Reaction conditions for (A): Diphenyl ether (1.70 g, 0.010 mol), 0.2 wt% Pd/C (40.0 mg, 2.3W10@7 mol of Pdsurface, prepared by diluting 5 wt% Pd ...
  24. [24]
    Mild Heterogeneous Palladium‐Catalyzed Cleavage of β‐O‐4 ...
    Nov 4, 2013 · This reaction is applicable to cleaving the β-O-4′-ether bond found in lignin polymers of different origin.Missing: silanes | Show results with:silanes
  25. [25]
    Efficient Greener Visible‐Light‐Catalyzed Debenzylation of Benzyl ...
    May 13, 2022 · Benzyl deprotection from more crowded and rigid polycyclic diosgenin benzyl ether 8 found to take longer (4 h) time to yield 88% of diosgenin.