Alcohols & Ethers

Alcohol Nomenclature

In the IUPAC system of nomenclature, functional groups are normally designated in one of two ways. The presence of the function may be indicated by a characteristic suffix and a location number. This is common for the carbon-carbon double and triple bonds which have the respective suffixes ene and yne. Halogens, on the other hand, do not have a suffix and are named as substituents, for example: (CH3)2C=CHCHClCH3 is 4-chloro-2-methyl-2-pentene. If you are uncertain about the IUPAC rules for nomenclature you should review them now.
Alcohols are usually named by the first procedure and are designated by an ol suffix, as in ethanol, CH3CH2OH (note that a locator number is not needed on a two-carbon chain). On longer chains the location of the hydroxyl group determines chain numbering. For example: (CH3)2C=CHCH(OH)CH3 is 4-methyl-3-penten-2-ol. Other examples of IUPAC nomenclature are shown below, together with the common names often used for some of the simpler compounds. For the monofunctional alcohols, this common system consists of naming the alkyl group followed by the word alcohol. Alcohols may also be classified as primary, ; secondary, & tertiary, , in the same manner as alkyl halides. This terminology refers to alkyl substitution of the carbon atom bearing the hydroxyl group (colored blue in the illustration).

Many functional groups have a characteristic IUPAC suffix designator, and only one such suffix (other than "ene" and "yne") may be used in a name.
For a list of characteristic IUPAC suffixes used to designate different functional groups Click Here.

When the hydroxyl functional group is present together with a function of higher nomenclature priority, it must be cited and located by the prefix hydroxy and an appropriate number. For example, lactic acid has the IUPAC name 2-hydroxypropanoic acid.

Compounds incorporating a C–S–H functional group are named thiols or mercaptans. The IUPAC name of (CH3)3C–SH is 2-methyl-2-propanethiol, commonly called tert-butyl mercaptan. The chemistry of thiols will not be described here, other than to note that they are stronger acids and more powerful nucleophiles than alcohols.


Reactions of Alcohols

The functional group of the alcohols is the hydroxyl group, –OH. Unlike the alkyl halides, this group has two reactive covalent bonds, the C–O bond and the O–H bond. The electronegativity of oxygen is substantially greater than that of carbon and hydrogen. Consequently, the covalent bonds of this functional group are polarized so that oxygen is electron rich and both carbon and hydrogen are electrophilic, as shown in the drawing on the right. Indeed, the dipolar nature of the O–H bond is such that alcohols are much stronger acids than alkanes (by roughly 1030 times), and nearly that much stronger than ethers (oxygen substituted alkanes that do not have an O–H group). The most reactive site in an alcohol molecule is the hydroxyl group, despite the fact that the O–H bond strength is significantly greater than that of the C–C, C–H and C–O bonds, demonstrating again the difference between thermodynamic and chemical stability.

For a discussion of how acidity is influenced by molecular structure Click Here.

      1. Electrophilic Substitution at Oxygen
Because of its enhanced acidity, the hydrogen atom on the hydroxyl group is rather easily replaced by other substituents. A simple example is the facile reaction of simple alcohols with sodium (and sodium hydride), as described in the first equation below. Another such substitution reaction is the isotopic exchange that occurs on mixing an alcohol with deuterium oxide (heavy water). This exchange, which is catalyzed by acid or base, is very fast under normal conditions, since it is difficult to avoid traces of such catalysts in most experimental systems.

2 R–O–H   +   2 Na     2 R–O(–)Na(+)   +   H2
R–O–H   +   D2O     R–O–D   +   D–O–H

The mechanism by which many substitution reactions of this kind take place is straightforward. The oxygen atom of an alcohol is nucleophilic and is therefore prone to attack by electrophiles. The resulting "onium" intermediate then loses a proton to a base, giving the substitution product. If a strong electrophile is not present, the nucleophilicity of the oxygen may be enhanced by conversion to its conjugate base (an alkoxide). This powerful nucleophile then attacks the weak electrophile. These two variations of the substitution mechanism are illustrated in the following diagram.

The preparation of tert-butyl hypochlorite from tert-butyl alcohol is an example of electrophilic halogenation of oxygen, but this reaction is restricted to 3º-alcohols because 1º and 2º-hypochlorites lose HCl to give aldehydes and ketones. In the following equation the electrophile may be regarded as Cl(+).

(CH3)3C–O–H   +   Cl2   +   NaOH     (CH3)3C–O–Cl   +   NaCl  +   H2O

Alkyl substitution of the hydroxyl group leads to ethers. This reaction provides examples of both strong electrophilic substitution (first equation below), and weak electrophilic substitution (second equation). The latter SN2 reaction is known as the Williamson Ether Synthesis, and is generally used only with 1º-alkyl halide reactants because the strong alkoxide base leads to E2 elimination of 2º and 3º-alkyl halides.

One of the most important substitution reactions at oxygen is ester formation resulting from the reaction of alcohols with electrophilic derivatives of carboxylic and sulfonic acids. The following illustration displays the general formulas of these reagents and their ester products, in which the R'–O– group represents the alcohol moiety. The electrophilic atom in the acid chlorides and anhydrides is colored red. Examples of specific esterification reactions may be selected from the menu below the diagram, and will be displayed in the same space.


      2. Nucleophilic Substitution of the Hydroxyl Group
Using the chemical behavior of alkyl halides as a reference, we are encouraged to look for analogous substitution and elimination reactions of alcohols. The chief difference, of course, is a change in the leaving anion from halide to hydroxide. Since oxygen is slightly more electronegative than chlorine (3.5 vs. 2.8 on the Pauling scale), we expect the C-O bond to be more polar than a C-Cl bond. Furthermore, an independent measure of the electrophilic character of carbon atoms from their nmr chemical shifts (both 13C & alpha protons), indicates that oxygen and chlorine substituents exert a similar electron-withdrawing influence when bonded to sp3 hybridized carbon atoms. Despite this promising background evidence, alcohols do not undergo the same SN2 reactions commonly observed with alkyl halides. For example, the rapid SN2 reaction of 1-bromobutane with sodium cyanide, shown below, has no parallel when 1-butanol is treated with sodium cyanide. In fact ethyl alcohol is often used as a solvent for alkyl halide substitution reactions such as this.

CH3CH2CH2CH2–Br +  Na(+) CN(–)     CH3CH2CH2CH2–CN +  Na(+) Br(–)
CH3CH2CH2CH2–OH +  Na(+) CN(–)     No Reaction

The key factor here is the stability of the leaving anion (bromide vs. hydroxide). We know that HBr is a much stronger acid than water (by more than 18 powers of ten), and this difference will be reflected in reactions that generate their conjugate bases. The weaker base, bromide anion, is more stable and its release in a substitution or elimination reaction will be much more favorable than that of hydroxide ion, a stronger and less stable base.
Clearly, an obvious step toward improving the reactivity of alcohols in SN2 reactions would be to modify the –OH functional group in a way that improves its stability as a leaving anion. One such modification is to conduct the substitution reaction in strong acid so that –OH is converted to –OH2(+). Since the hydronium ion (H3O(+)) is a much stronger acid than water, its conjugate base (H2O) is a better leaving group than hydroxide ion. The only problem with this strategy is that many nucleophiles, including cyanide, are deactivated by protonation in strong acid, effectively removing the nucleophilic co-reactant needed for the substitution. The strong acids HCl, HBr and HI are not subject to this difficulty because their conjugate bases are good nucleophiles and are even weaker bases than alcohols. The following equations illustrate some substitution reactions of alcohols that may be effected by these acids. As was true for alkyl halides, nucleophilic substitution of 1º-alcohols proceeds by a SN2 mechanism, whereas 3º-alcohols react by a SN1 mechanism. Reactions of 2º-alcohols may occur by both mechanisms and often produce some rearranged products. The numbers in parentheses next to the mineral acid formulas represent the weight percentage of a concentrated aqueous solution, the form in which these acids are normally used.

CH3CH2CH2CH2–OH +  HBr (48%)     CH3CH2CH2CH2–OH2(+) Br(–)     CH3CH2CH2CH2Br +  H2O       SN2
(CH3)3C–OH +  HCl (37%)     (CH3)3C–OH2(+) Cl(–)     (CH3)3C(+) Cl(–) +  H2O    (CH3)3C–Cl +  H2O       SN1

Although these reactions are sometimes referred to as "acid-catalyzed" this is not strictly correct. In the overall transformation a strong HX acid is converted to water, a very weak acid, so at least a stoichiometric quantity of HX is required for a complete conversion of alcohol to alkyl halide. The necessity of using equivalent quantities of very strong acids in this reaction limits its usefulness to simple alcohols of the kind shown above. Alcohols having acid sensitive groups would, of course, not tolerate such treatment. Nevertheless, the idea of modifying the -OH functional group to improve its stability as a leaving anion can be pursued in other directions. The following diagram shows some modifications that have proven effective. In each case the hydroxyl group is converted to an ester of a strong acid. The first two examples show the sulfonate esters described earlier. The third and fourth examples show the formation of a phosphite ester (X represents remaining bromines or additional alcohol substituents) and a chlorosulfite ester respectively. All of these leaving groups (colored blue) have conjugate acids that are much stronger than water (by 13 to 16 powers of ten) so the leaving anion is correspondingly more stable than hydroxide ion. The mesylate and tosylate compounds are particularly useful in that they may be used in substitution reactions with a wide variety of nucleophiles. The intermediates produced in reactions of alcohols with phosphorus tribromide and thionyl chloride (last two examples) are seldom isolated, and these reactions continue on to alkyl bromide and chloride products.

The importance of sulfonate ester intermediates in general nucleophilic substitution reactions of alcohols may be illustrated by the following conversion of 1-butanol to pentanenitrile (butyl cyanide), a reaction that does not occur with the alcohol alone (see above). The phosphorus and thionyl halides, on the other hand, only act to convert alcohols to the corresponding alkyl halides.

CH3CH2CH2CH2–OH + CH3SO2Cl pyridine

 
CH3CH2CH2CH2–OSO2CH3 Na(+) CN(–)

 
CH3CH2CH2CH2CN + CH3SO2O(–) Na(+)

Some examples of alcohol substitution reactions using this approach to activating the hydroxyl group are shown in the following diagram. The first two cases serve to reinforce the fact that sulfonate ester derivatives of alcohols may replace alkyl halides in a variety of SN2 reactions. The next two cases demonstrate the use of phosphorus tribromide in converting alcohols to bromides. This reagent may be used without added base (e.g. pyridine), because the phosphorous acid product is a weaker acid than HBr. Phosphorous tribromide is best used with 1º-alcohols, since 2º-alcohols often give rearrangement by-products resulting from competing SN1 reactions. Note that the ether oxygen in reaction 4 is not affected by this reagent; whereas, the alternative synthesis using concentrated HBr cleaves ethers. Phosphorus trichloride (PCl3) converts alcohols to alkyl chlorides in a similar manner, but thionyl chloride is usually preferred for this transformation since the inorganic products are gases (SO2 & HCl). Phosphorus triiodide is not stable, but may be generated in situ from a mixture of red phosphorus and bromine, and acts to convert alcohols to alkyl iodides. The last example shows the reaction of thionyl chloride with a chiral 2º-alcohol. The presence of an organic base such as pyridine is important, because it provides a substantial concentration of chloride ion needed for the final SN2 reaction of the chlorosufite intermediate. In the absence of base chlorosufites decompose on heating to give the expected alkyl chloride with retention of configuration
Tertiary alcohols are not commonly used for substitution reactions of the kind discussed here, because SN1 and E1 reaction paths are dominant and are difficult to control. This aspect of alcohol chemistry will be touched upon in the next section.

The importance of sulfonate esters as intermediates in many substitution reactions cannot be overstated. A rigorous proof of the configurational inversion that occurs at the substitution site in SN2 reactions makes use of such reactions. An example of such a proof will display above when the An Inversion Proof button beneath the diagram is pressed. Abbreviations for the more commonly used sulfonyl derivatives are given in the following table.

Sulfonyl GroupCH3SO2CH3C6H4SO2BrC6H4SO2CF3SO2
Name & Abbrev.Mesyl or MsTosyl or TsBrosyl or BsTrifyl or Tf

Many other selective methods for hydroxyl substitution have been developed. To examine some of these Click Here

      3. Elimination Reactions of Alcohols
In the discussion of alkyl halide reactions we noted that 2º and 3º-alkyl halides experienced rapid E2 elimination when treated with strong bases, such as hydroxide and alkoxides. Alcohols do not undergo such base-induced elimination reactions and are, in fact, often used as solvents for such reactions. This is yet another example of how leaving group stability often influences the rate of a reaction.
When an alcohol is treated with sodium hydroxide, the following acid-base equilibrium occurs. Most alcohols are slightly weaker acids than water so the left side is favored.

R–O–H   +   Na(+) OH(–)     R–O(–) Na(+)   +   H–OH

The elimination of water from an alcohol is called dehydration. Recalling that water is a much better leaving group than hydroxide ion, it is sensible to use acid-catalysis rather than base-catalysis to achieve such reactions. Four examples of this useful technique are shown below. Note that hydrohalic acids (HX) are not normally used as catalysts because their conjugate bases are good nucleophiles and may give substitution products. The conjugate bases of sulfuric and phosphoric acids are not good nucleophiles and do not give substitution under the usual conditions of their use.

The first two examples (top row) are typical, and the more facile elimination of the 3º-alcohol suggests predominant E1 character for the reaction. This agrees with the tendency of branched 1º and 2º-alcohols to give rearrangement products, as shown in the last example. The last two reactions also demonstrate that the Zaitsev Rule applies to alcohol dehydrations as well as alkyl halide eliminations. Thus the more highly-substituted double bond isomer is favored among the products.
It should be noted that the acid-catalyzed dehydrations discussed here are the reverse of the acid-catalyzed hydration reactions of alkenes. Indeed, for reversable reactions such as this the laws of thermodynamics require that the mechanism in both directions proceed by the same reaction path. This is known as the principle of microscopic reversibility. To illustrate, the following diagram lists the three steps in each transformation. The dehydration reaction is shown by the blue arrows; the hydration reaction by magenta arrows. The intermediates in these reactions are common to both, and common transition states are involved. This can be seen clearly in the energy diagrams depicted by clicking the button beneath the equations.

Base induced E2 eliminations of alcohols may be achieved if their sulfonate ester derivatives are used. This has the advantage of avoiding strong acids, which may cause molecular rearrangement and / or double bond migration in some cases. Since 3º-sulfonate derivatives are sometimes unstable, this procedure is best used with 1º and 2º-mesylates or tosylates. Application of this reaction sequence is shown here for 2-butanol. The Zaitsev Rule favors formation of 2-butene (cis + trans) over 1-butene.

CH3CH2CH(CH3)–OH CH3SO2Cl

 
CH3CH2CH(CH3)–OSO2CH3 C2H5O(–)Na(+)

 
CH3CH=CHCH3 + CH3CH2CH=CH2 + CH3SO2O(–) Na(+) + C2H5OH

The E2 elimination of 3º-alcohols under relatvely non-acidic conditions may be accomplished by treatment with phosphorous oxychloride (POCl3) in pyridine. This procedure is also effective with hindered 2º-alcohols, but for unhindered and 1º-alcohols an SN2 chloride ion substitution of the chlorophosphate intermediate competes with elimination. Some examples of these and related reactions are given in the following figure. The first equation shows the dehydration of a 3º-alcohol. The predominance of the non-Zaitsev product (less substituted double bond) is presumed due to steric hindrance of the methylene group hydrogens, which interferes with the approach of base at that site. The second example shows two elimination procedures applied to the same 2º-alcohol. The first uses the single step POCl3 method, which works well in this case because SN2 substitution is retarded by steric hindrance. The second method is another example in which an intermediate sulfonate ester confers halogen-like reactivity on an alcohol. In every case the anionic leaving group is the conjugate base of a strong acid.

Pyrolytic syn-Eliminations
To see some examples of thermal, unimolecular syn-eliminations Click Here


      4. Oxidation Reactions of Alcohols
Simple 1º and 2º-alcohols in the gaseous state lose hydrogen when exposed to a hot copper surface. This catalytic dehydrogenation reaction produces aldehydes (as shown below) and ketones, and since the carbon atom bonded to the oxygen is oxidized, such alcohol to carbonyl conversions are generally referred to as oxidation reactions. Gas phase dehydrogenations of this kind are important in chemical manufacturing, but see little use in the research laboratory. Instead, alcohol oxidations are carried out in solution, using reactions in which the hydroxyl hydrogen is replaced by an atom or group that is readily eliminated together with the alpha-hydrogen. The decomposition of 1º and 2º-alkyl hypochlorites, referred to earlier, is an example of such a reaction.

RCH2–OH   +   hot Cu    RCH=O   +   H2

RCH2–O–Cl   +   base    RCH=O   +   H–Cl

The most generally useful reagents for oxidizing 1º and 2º-alcohols are chromic acid derivatives. Two such oxidants are Jones reagent (a solution of sodium dichromate in aqueous sulfuric acid) and pyridinium chlorochromate, C5H5NH(+)CrO3Cl(–), commonly named by the acronym PCC and used in methylene chloride solution. In each case a chromate ester of the alcohol substrate is believed to be an intermediate, which undergoes an E2-like elimination to the carbonyl product. The oxidation state of carbon increases by 2, while the chromium decreases by 3 (it is reduced). Since chromate reagents are a dark orange-red color (VI oxidation state) and chromium III compounds are normally green, the progress of these oxidations is easily observed. Indeed, this is the chemical transformation on which the Breathalizer test is based. The following equations illustrate some oxidations of alcohols, using the two reagents defined here. Both reagents effect the oxidation of 2º-alcohols to ketones, but the outcome of 1º-alcohol oxidations is different. Oxidation with the PCC reagent converts 1º-alcohols to aldehydes; whereas Jones reagent continues the oxidation to the carboxylic acid product, as shown in the second reaction. Reaction mechanisms for these transformations are displayed on clicking the "Show Mechanism" button. For the first two reactions the mechanism diagram also shows the oxidation states of carbon (blue Arabic numbers) and chromium (Roman numbers). The general base (B:) used in these mechanisms may be anything from water to pyridine, depending on the specific reaction.


Two structural requirements for the oxidation to carbonyl products should now be obvious:
        1. The carbon atom bonded to oxygen must also bear a hydrogen atom.
            Tertiary alcohols (R3C–OH) cannot be oxidized in this fashion.
        2. The oxygen atom must be bonded to a hydrogen atom so that a chromate ester intermediate (or other suitable leaving group) may be formed.
            Ethers (R–O–R) cannot be oxidized in this fashion.

The fourth reaction above illustrates the failure of 3º-alcohols to undergo oxidation. The second reaction mechanism explains why 1º-alcohols undergo further oxidation by Jones reagent. The aqueous solvent system used with this reagent permits hydration (addition of water) to the aldehyde carbonyl group. The resulting hydrate (structure shown below the aldehyde) meets both the requirements stated above, and is further oxidized by the same chromate ester mechanism. Water is not present when the PCC reagent is used, so the oxidation stops at the aldehyde stage.
Another chromate oxidizing agent, similar to PCC, is pyridinium dichromate, (C5H5NH(+) )2 Cr2O7(–2), known by the acronym PDC. Both PCC and PDC are orange crystalline solids that are soluble in many organic solvents. Since PDC is less acidic than PCC it is often used to oxidize alcohols that may be sensitive to acids. In methylene chloride solution, PDC oxidizes 1º- and 2º-alcohols in roughly the same fashion as PCC, but much more slowly. However, in DMF (dimethylformamide) solution saturated 1º-alcohols are oxidized to carboxylic acids. In both solvents allylic alcohols are oxidized efficiently to conjugated enals and enones respectively.
Many mild and selective alternatives to PCC and Jones' reagent exist for the oxidation of 1º- and 2º-alcohols. To examine some of these Click Here


Practice Problems

The following problems explore several aspects of alcohol nomenclature and chemistry. In a few cases reactions of phenols may be selected, so these may be deferred to the next chapter.

Ethers

      1. Nomenclature
Ethers are compounds having two alkyl or aryl groups bonded to an oxygen atom, as in the formula R1–O–R2. The ether functional group does not have a characteristic IUPAC nomenclature suffix, so it is necessary to designate it as a substituent. To do so the common alkoxy substituents are given names derived from their alkyl component, as shown in the table on the right below. Examples of ether nomenclature are provided on the left. Simple ethers are given common names in which the alkyl groups bonded to the oxygen are named in alphabetical order followed by the word "ether". The top left example shows the common name in blue under the IUPAC name. Many simple ethers are symmetrical, in that the two alkyl substituents are the same. These are named as "dialkyl ethers". Examples are: CH3CH2OCH2CH3, diethyl ether (sometimes referred to as ether), and CH3OCH2CH2OCH3, ethylene glycol dimethyl ether (glyme).

Alkyl GroupName   Alkoxy GroupName
CH3MethylCH3O–Methoxy
CH3CH2Ethyl CH3CH2O– Ethoxy
(CH3)2CH–Isopropyl(CH3)2CHO– Isopropoxy
(CH3)3C–tert-Butyl (CH3)3CO– tert-Butoxy
C6H5Phenyl C6H5O– Phenoxy

Sulfur analogs of ethers (R–S–R') are called sulfides. For example, (CH3)3C–S–CH3 is tert-butyl methyl sulfide. Sulfides are chemically more reactive than ethers, reflecting the greater nucleophilicity of sulfur relative to oxygen.


      2. Preparation of Ethers
Ethers are usually prepared from alcohols or their conjugate bases. One important procedure, known as the Williamson Ether Synthesis, proceeds by an SN2 reaction of an alkoxide nucleophile with an alkyl halide. Reactions #1 and #2 below are two examples of this procedure. When applied to an unsymmetrical ether, as in this case, there are two different combinations of reactants are possible. Of these one is usually better than the other. Since alkoxide anions are strong bases, the possibility of a competing E2 elimination must always be considered. Bearing in mind the factors that favor substitution over elimination, a 1º-alkyl halide should be selected as a preferred reactant whenever possible. Thus, reaction #1 gives a better and cleaner yield of benzyl isopropyl ether than does reaction #2, which generates considerable elimination product.

A second general ether synthesis, alkoxymercuration, is patterned after the oxymercuration reaction. Reactions #3 and #4 are examples of this two-step procedure. Note that the alcohol reactant is used as the solvent, and a trifluoroacetate mercury (II) salt is used in preference to the acetate (trifluoroacetate anion is a poorer nucleophile than acetate). The mechanism of alkoxymercuration is similar to that of oxymercuration, with an initial anti-addition of the mercuric species and alcohol being followed by reductive demercuration.
Acid-catalyzed dehydration of small 1º-alcohols constitutes a specialized method of preparing symmetrical ethers. As shown in the following two equations, the success of this procedure depends on the temperature. At 110º to 130 ºC an SN2 reaction of the alcohol conjugate acid leads to an ether product. At higher temperatures (over 150 ºC) an E2 elimination takes place.

2 CH3CH2-OH   +   H2SO4130 ºC

CH3CH2-O-CH2CH3   +   H2O
CH3CH2-OH   +   H2SO4150 ºC

CH2=CH2   +   H2O


      3. Reactions of Ethers
Ethers are widely used as solvents for a variety of organic compounds and reactions, suggesting that they are relatively unreactive themselves. Indeed, with the exception of the alkanes, cycloalkanes and saturated fluorocarbons, ethers are probably the least reactive, common class of organic compounds. The inert nature of the ethers relative to the alcohols is undoubtedly due to the absence of the reactive O–H bond.
The most common reaction of ethers is cleavage of the C–O bond by strong acids. This may occur by SN1 or E1 mechanisms for 3º-alkyl groups or by an SN2 mechanism for 1º-alkyl groups. Some examples are shown in the following diagram. The conjugate acid of the ether is an intermediate in all these reactions, just as conjugate acids were intermediates in certain alcohol reactions.

The first two reactions proceed by a sequence of SN2 steps in which the iodide or bromide anion displaces an alcohol in the first step, and then converts the conjugate acid of that alcohol to an alkyl halide in the second. Since SN2 reactions are favored at least hindered sites, the methyl group in example #1 is cleaved first. The 2º-alkyl group in example #3 is probably cleaved by an SN2 mechanism, but the SN1 alternative cannot be ruled out. The phenol formed in this reaction does not react further, since SN2, SN1 and E1 reactions do not take place on aromatic rings. The last example shows the cleavage of a 3º-alkyl group by a strong acid. Acids having poorly nucleophilic conjugate bases are often chosen for this purpose so that E1 products are favored. The reaction shown here (#4) is the reverse of the tert-butyl ether preparation described earlier.
Ethers in which oxygen is bonded to 1º- and 2º-alkyl groups are subject to peroxide formation in the presence of air (gaseous oxygen). This reaction presents an additional hazard to the use of these flammable solvents, since peroxides decompose explosively when heated or struck. The mechanism of peroxide formation is believed to be free radical in nature (note that molecular oxygen has two unpaired electrons).

R–O–CH(CH3)2   +   O2R–O–C(CH3)2–O–O–H   a peroxide

Because of their chemical stabilty, ethers may be used to protect hydroxyl functions from undergoing unwanted reactions.
To learn more about this application Click Here


      4. Reactions of Epoxides
Epoxides (oxiranes) are three-membered cyclic ethers that are easily prepared from alkenes by reaction with peracids. Because of the large angle strain in this small ring, epoxides undergo acid and base-catalyzed C–O bond cleavage more easily than do larger ring ethers. Among the following examples, the first is unexceptional except for the fact that it occurs under milder conditions and more rapidly than other ether cleavages. The second and third examples clearly show the exceptional reactivity of epoxides, since unstrained ethers present in the same reactant or as solvent do not react. The aqueous acid used to work up the third reaction, following the Grignard reagent cleavage of the ethylene oxide, simply neutralizes the magnesium salt of the alcohol product.




Practice Problems

Two problems involving the reactions of ethers are given here, together with a large collection of multiple choice questions covering halides, alcohols and ethers.


Sulfur is below oxygen in the periodic table. To see examples of organosulfur compounds and their chemistry Click Here


Return to Table of Contents