0% found this document useful (0 votes)
58 views36 pages

Reactivity of Alcohols

Uploaded by

jannatulabc10
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
58 views36 pages

Reactivity of Alcohols

Uploaded by

jannatulabc10
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 36

REACTIVITY OF

ALCOHOLS
CHAPTER OVERVIEW

Reactivity 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.

Topic hierarchy
Dehydrating Alcohols to Make Alkenes
Electrophilic Substitution at Oxygen
Elimination Reactions of Alcohols
Hydroxyl Group Substitution
Additional Methods of Hydroxyl Substitution
Reactions of alcohols with hydrohalic acids (HX)
Reduction of Alcohols
Replacing the OH Group by Halogen Atoms
The Oxidation of Alcohols
Oxidation by Chromic Acid
Oxidation by PCC (pyridinium chlorochromate)
The Reaction Between Alcohols and Sodium
The Triiodomethane (Iodoform) Reaction
Thionyl Chloride

Reactivity of Alcohols is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
Dehydrating Alcohols to Make Alkenes
This page discusses the dehydration of alcohols in the lab to make alkenes—for example, dehydrating ethanol to produce ethene.

The dehydration of ethanol to give ethene


This is a simple method of making gaseous alkenes such as ethene. If ethanol vapor is passed over heated aluminum oxide powder,
the ethanol is essentially cracked to yield ethene and water vapor.

To produce a few test tubes of ethene, the following apparatus can be used:

This system can be scaled up by boiling ethanol in a flask and passing the vapor over aluminum oxide that is heated in a long tube.

Dehydration of alcohols using an acid catalyst


The acid catalysts normally used in alcohol dehydration are either concentrated sulfuric acid or concentrated phosphoric(V) acid,
H3PO4. Concentrated sulfuric acid produces messy results. Because sulfuric acid is also a strong oxidizing agent, it oxidizes some
of the alcohol to carbon dioxide and is simultaneously reduced itself to sulfur dioxide. Both of these gases must be removed from
the alkene. Sulfuric acid also reacts with the alcohol to produce a mass of carbon. There are other side reactions as well (not
discussed here).

The dehydration of ethanol to yield ethene


In this process, ethanol is heated with an excess of concentrated sulfuric acid at a temperature of 170°C. The gases produced are
passed through a sodium hydroxide solution to remove the carbon dioxide and sulfur dioxide produced from side reactions. The
ethene is collected over water.

The concentrated sulfuric acid is a catalyst. Therefore, it is written over the reaction arrow rather than in the equation.

The dehydration of cyclohexanol to yield cyclohexene


This is a preparation commonly used to illustrate the formation and purification of a liquid product. The fact that the carbon atoms
are joined in a ring has no bearing on the chemistry of the reaction. Cyclohexanol is heated with concentrated phosphoric(V) acid,
and the liquid cyclohexene distils off and can be collected and purified. Phosphoric(V) acid tends to be used instead of sulfuric acid
because it is safer and facilitates a less complex reaction.

The dehydration of more complicated alcohols


With more complicated alcohols, the formation of more than one alkene is possible. Butan-2-ol is a good example of this, with
three different alkenes formed when it is dehydrated.

1 https://chem.libretexts.org/@go/page/3901
When an alcohol is dehydrated, the -OH group and a hydrogen atom from the next carbon atom in the chain are removed. With
molecules like butan-2-ol, there are two possibilities for this.

The following products are formed:

The products are but-1-ene, CH2=CHCH2CH3, and but-2-ene, CH3CH=CHCH3.


This situation is further complicated by the fact that but-2-ene exhibits geometric isomerism; thus, a mixture of two isomers is
formed: cis-but-2-ene and trans-but-2-ene.

The compoundcis-but-2-ene is also known as (Z)-but-2-ene; trans-but-2-ene is also known as (E)-but-2-ene. Which isomer is
formed is a matter of chance.
Hence, the dehydration of butan-2-ol leads to a mixture containing the following compounds:
but-1-ene
cis-but-2-ene (also known as (Z)-but-2-ene)
trans-but-2-ene (also known as (E)-but-2-ene)

Contributors
Jim Clark (Chemguide.co.uk)

This page titled Dehydrating Alcohols to Make Alkenes is shared under a CC BY-NC 4.0 license and was authored, remixed, and/or curated by
Jim Clark.

2 https://chem.libretexts.org/@go/page/3901
Electrophilic Substitution at Oxygen
Because of its enhanced acidity, the hydrogen atom of a hydroxyl group is 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 when mixing an alcohol with deuterium oxide (heavy water). This
exchange, which is catalyzed by acid or base, is rapid under normal conditions because it is difficult to avoid traces of these
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 these substitution reactions proceed is straightforward. The oxygen atom of an alcohol is nucleophilic;
therefore, it is prone to attack by electrophiles. The resulting "onium" intermediate then loses a proton to a base, forming the
substitution product. If a strong electrophile is not present, then 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; 1º and 2º hypochlorites lose HCl, forming 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 creates 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.

Ester Formation
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; the R'–O– group represents the alcohol moiety. The electrophilic atoms in the acid chlorides and
anhydrides are colored red. Examples of specific esterification reactions are shown below.

1 https://chem.libretexts.org/@go/page/739
Acyl Chloride Esterification

Anhydride Esterification

Methanesulfonate Formation

Tosylate Formation

Contributors
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry

This page titled Electrophilic Substitution at Oxygen is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
William Reusch.

2 https://chem.libretexts.org/@go/page/739
Elimination Reactions of Alcohols
The discussion of alkyl halide reactions noted that 2º and 3º-alkyl halides experience 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 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 in 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 create substitution products. The conjugate bases of sulfuric and phosphoric acids are not good nucleophiles, and do not
participate in substitution under typical conditions.

The first two examples (in the top row) are typical, and the more facile elimination of the 3º-alcohol suggests the predominant E1
character for the reaction. This agrees with the tendency of branched 1º and 2º-alcohols to yield 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 to alkyl
halide eliminations. Therefore, 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 these types of reversible reactions, 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.

1 https://chem.libretexts.org/@go/page/1166
Energy Profile

Base-induced E2 eliminations of alcohols may be achieved from sulfonate ester derivatives. This has the advantage of avoiding
strong acids, which may cause molecular rearrangement or double bond migration in some cases. Because 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 the formation of 2-butene (cis + trans) over 1-butene.

CH3SO2Cl C2H5O(–)Na(+)
CH3CH2CH(CH3)–OH CH3CH2CH(CH3)–OSO2CH3

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

The E2 elimination of 3º-alcohols under relatively 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. 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 hydrogen atoms, which
interferes with the approach of the 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 impeded 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.

2 https://chem.libretexts.org/@go/page/1166
Pyrolytic syn-Eliminations
Ester derivatives of alcohols may undergo unimolecular syn-elimination upon heating. To see examples of these, Click Here.

Contributors
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry

This page titled Elimination Reactions of Alcohols is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
William Reusch.

3 https://chem.libretexts.org/@go/page/1166
Hydroxyl Group Substitution
Nucleophilic Substitution of the Hydroxyl Group
The chemical behavior of alkyl halides can be used as a reference in discovering analogous substitution and elimination reactions of
alcohols. The chief difference, of course, is a change in the leaving anion from halide to hydroxide. Because oxygen is slightly more
electronegative than chlorine (3.5 vs. 2.8 on the Pauling scale), the C-O bond is expected to be more polar than a C-Cl bond. Furthermore,
an independent measure of the electrophilic characteristics of carbon atoms from their NMR chemical shifts (both 13C and 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). HBr is a much stronger acid than water (by more than 18
orders of magnitude), and this difference is reflected in reactions that generate their respective conjugate bases. The weaker base, bromide,
is more stable, and its release in a substitution or elimination reaction is much more favorable than that of hydroxide ion, a stronger and
less stable base.
A clear 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 a strong acid, converting –OH to –
OH2(+). Because 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
acids, effectively removing the nucleophilic co-reactant required 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 affected by these acids. As with alkyl halides, the nucleophilic substitution of
1º-alcohols proceeds by an SN2 mechanism, whereas 3º-alcohols react by an 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(–) CH3CH2CH2CH2–Br + 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 type shown above. Alcohols with acid-sensitive groups do not, of course, 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 the 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); thus, the leaving anion is
correspondingly more stable than the hydroxide ion. The mesylate and tosylate compounds are particularly useful because 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 to produce alkyl bromide and chloride
products.

1 https://chem.libretexts.org/@go/page/1167
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. The phosphorus
and thionyl halides, on the other hand, only act to convert alcohols to the corresponding alkyl halides.

pyridine Na(+) CN(–)


CH3CH2CH2CH2–OH + CH3SO2Cl CH3CH2CH2CH2–OSO2CH3 CH3CH2CH2CH2–CN + CH3SO2O(

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. Phosphorus tribromide is best used
with 1º-alcohols because 2º-alcohols often yield 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
because 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 iodine 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 required for the final SN2 reaction of the chlorosufite intermediate. In the absence of a base, chlorosufites decompose upon heating to
yield the expected alkyl chloride with retention of configuration
Tertiary alcohols are not commonly used for substitution reactions of the type discussed here because SN1 and E1 reaction paths are
dominant and are difficult to control.

2 https://chem.libretexts.org/@go/page/1167
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 is
displayed below. Abbreviations for the more commonly used sulfonyl derivatives are given in the following table.

Sulfonyl Group CH3SO2– CH3C6H4SO2– BrC6H4SO2– CF3SO2–

Name & Abbrev. Mesyl or Ms Tosyl or Ts Brosyl or Bs Trifyl or Tf

Inversion Proof

For a more complete discussion of hydroxyl substitution reactions and a description of other selective methods for this transformation,
Click Here.

Contributors
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry

This page titled Hydroxyl Group Substitution is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by William Reusch.

3 https://chem.libretexts.org/@go/page/1167
Additional Methods of Hydroxyl Substitution
The most common methods for converting 1º- and 2º-alcohols to the corresponding chloro and bromo alkanes (i.e. replacement of
the hydroxyl group) are treatments with thionyl chloride and phosphorus tribromide, respectively. These reagents are generally
preferred over the use of concentrated HX due to the harsh acidity of these hydrohalic acids and the carbocation rearrangements
associated with their use.
Of course, it is possible to avoid such problems by first preparing a mesylate or tosylate derivative, followed by nucleophilic
substitution of the sulfonate ester by the appropriate halide anion. In this two-step approach, a clean configurational inversion
occurs in the first SN2 reaction; however, the resulting alkyl halide may then undergo repeated SN2 halogen exchange reactions,
thus destroying any stereoisomeric identity held by the initial carbinol carbon. For these and other reasons, alternative mild and
selective methods for transforming such alcohols by nucleophilic substitution of the hydroxyl group have been devised. It should
be noted that 3º-alcohols are not good substrates for the new procedures.

Disadvantages to using P Br and SOC l 3 2

Despite their general usefulness, phosphorous tribromide and thionyl chloride have shortcomings. Hindered 1º- and 2º-alcohols
react sluggishly with the former and may form rearrangement products, as noted in the following equation.

Below, an abbreviated mechanism for the reaction is displayed. The initially formed trialkylphosphite ester may be isolated if the
HBr byproduct is scavenged by a base. In the presence of HBr, a series of acid-base and SN2 reactions occur, along with the
transient formation of carbocation intermediates. Rearrangement (pink arrows) of the carbocations leads to isomeric products.

The reaction of thionyl chloride with chiral 2º-alcohols has been observed to proceed with either inversion or retention. In the
presence of a base such as pyridine, the intermediate chlorosulfite ester reacts to form a "pyridinium" salt, which undergoes a
relatively clean SN2 reaction to produce the inverted chloride. In ether and similar solvents, the chlorosulfite reacts with retention
of configuration, presumably by way of a tight or intimate ion pair. This is classified as an SNi reaction (nucleophilic substitution
internal). The carbocation partner in the ion pair may also rearrange. These reactions are illustrated by the following equations. An
alternative explanation for the retention of configuration, involving an initial solvent molecule displacement of the chlorosulfite
group (as SO2 and chloride anion), followed by chloride ion displacement of the solvent moiety, has been suggested. In this case,
two inversions lead to retention.

1 https://chem.libretexts.org/@go/page/1168
Another characteristic of thionyl chloride reactions is their tendency to yield allylic rearrangement products with allylic alcohols.
This fact is demonstrated by the following equations. Reactions of this type have been classified as SNi', where the prime mark
indicates an allylic characteristic to the internal substitution. They may also be considered retro-ene reactions, a special class of
pericyclic reactions.

A similar substitutive rearrangement also occurs with propargyl alcohols, as shown below.

Hypervalent Phosphorous Reagents


The ability of phosphorous to assume many different valencies or oxidation states was noted elsewhere. The nucleophilicity of
trialkyl phosphines allows them to bond readily to electrophiles, and the resulting phosphonium ions may then bond reversibly to
other nucleophiles, especially oxygen nucleophiles. The use of phosphorous ylides in the Wittig reaction is an example of this
reactivity.
The reaction of triphenylphosphine with halogens further illustrates this hypervalency. As shown in the following diagram,
triphenylphosphine (yellow box on the left) reacts to form a pentavalent dihalide, which is in equilibrium with its ionic components
in solution.

Chemists have used these and similar reagents to achieve the mild conversion of alcohols to alkyl halides with clean inversion of
configuration. As with other OH substitution reactions, an inherently poor leaving group (hydroxide anion) is modified to provide a
better leaving group, the stable compound triphenylphosphine oxide. Two such reactions are shown in the following diagram.

2 https://chem.libretexts.org/@go/page/1168
Using this approach, even sluggish alcohols that are prone to rearrangement (e.g. neopentyl alcohol) are converted to their
corresponding halides.

The instability of vicinal diiodides relative to their double bond analogs is the driving force for a novel transformation of vic-
glycols to their corresponding alkenes. An example is displayed below.

The allylic rearrangement observed in thionyl chloride is similarly avoided by using triphenylphosphine dichloride, or alternatively,
by a two-step procedure via a sulfonate ester. This is displayed below.

The Mitsunobu Reaction


The Japanese chemist, O.Mitsunobu, devised a general and exceptionally versatile variant of hypervalent phosphorous chemistry
that has been applied to a wide selection of alcohols. This method, which now carries his name, uses a reagent mixture consisting
of triphenylphosphine, diethyl azodicarboxylate (DEAD) and a moderate to strong acid. The steps leading to hydroxyl substitution
are outlined in the following diagram. It should be noted that the nucleophile involved in the final SN2 substitution may be the
conjugate base of the acid component or a separate species.

A common use of the Mitsunobu reaction is to invert the configuration of a 2º-alcohol. This application usually employs benzoic
acid or a benzoate salt, and the resulting configurationally inverted ester is then hydrolyzed to the epimeric alcohol. An example of
this procedure is displayed below.

3 https://chem.libretexts.org/@go/page/1168
In addition to achieving configurational inversion of carbinol sites, the Mitsunobu reaction has also been used to introduce the
azide precursor of amines and for the intramolecular preparation of cyclic ethers. Examples are shown below.

Contributors
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry

This page titled Additional Methods of Hydroxyl Substitution is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by William Reusch.

4 https://chem.libretexts.org/@go/page/1168
Reactions of alcohols with hydrohalic acids (HX)
When alcohols react with a hydrogen halide, a substitution occurs, producing an alkyl halide and water:

Scope of Reaction
The order of reactivity of alcohols is 3° > 2° > 1° methyl.
The order of reactivity of the hydrogen halides is HI > HBr > HCl (HF is generally unreactive).
The reaction is acid catalyzed. Alcohols react with the strongly acidic hydrogen halides HCl, HBr, and HI, but they do not react
with nonacidic NaCl, NaBr, or NaI. Primary and secondary alcohols can be converted to alkyl chlorides and bromides by allowing
them to react with a mixture of a sodium halide and sulfuric acid:

Mechanisms of the Reactions of Alcohols with HX


Secondary, tertiary, allylic, and benzylic alcohols appear to react by a mechanism that involves the formation of a carbocation in an
S 1 reaction with the protonated alcohol acting as the substrate.
N

The S 1 mechanism is illustrated by the reaction tert-butyl alcohol and aqueous hydrochloric acid (H O , C l ). The first two
N 3
+ −

steps in this S 1 substitution mechanism are protonation of the alcohol to form an oxonium ion. Although the oxonium ion is
n

formed by protonation of the alcohol, it can also be viewed as a Lewis acid-base complex between the cation (R ) and H O.
+
2

Protonation of the alcohol converts a poor leaving group (OH-) to a good leaving group (\)H_2O\_), which makes the dissociation
step of the S 1 mechanism more favorable.
N

In step 3, the carbocation reacts with a nucleophile (a halide ion) to complete the substitution.

When we convert an alcohol to an alkyl halide, we perform the reaction in the presence of acid and in the presence of halide ions
and not at elevated temperatures. Halide ions are good nucleophiles (they are much stronger nucleophiles than water), and because
halide ions are present in a high concentration, most of the carbocations react with an electron pair of a halide ion to form a more
stable species, the alkyl halide product. The overall result is an S 1 reaction.
n

Primary Alcohols
Not all acid-catalyzed conversions of alcohols to alkyl halides proceed through the formation of carbocations. Primary alcohols and
methanol react to form alkyl halides under acidic conditions by an SN2 mechanism.
In these reactions, the function of the acid is to produce a protonated alcohol. The halide ion then displaces a molecule of water (a
good leaving group) from carbon; this produces an alkyl halide:

1 https://chem.libretexts.org/@go/page/15409
Again, acid is required. Although halide ions (particularly iodide and bromide ions) are strong nucleophiles, they are not strong
enough to carry out substitution reactions with alcohols themselves. Direct displacement of the hydroxyl group does not occur
because the leaving group would have to be a strongly basic hydroxide ion:

We can see now why the reactions of alcohols with hydrogen halides are acid-promoted.

The role of acid catalysis


Acid protonates the alcohol hydroxyl group, making it a good leaving group. However, other strong Lewis acids can be used
instead of hydrohalic acids. Because the chloride ion is a weaker nucleophile than bromide or iodide ions, hydrogen chloride does
not react with primary or secondary alcohols unless zinc chloride or a similar Lewis acid is added to the reaction mixture as well.
Zinc chloride, a good Lewis acid, forms a complex with the alcohol through association with an unshared pair of electrons on the
oxygen atom. This enhances the hydroxyl’s leaving group potential sufficiently so that chloride can displace it.

Rearrangement
As we might expect, many reactions of alcohols with hydrogen halides, particularly those in which carbocations are formed, are
accompanied by rearrangements. The general rule is that if rearrangement CAN OCCUR (to form more stable or equally stable
cations), it will! In these reactions, mixtures of products can be formed.

Contributors
Prof. Paul G. Wenthold (Purdue University)

Reactions of alcohols with hydrohalic acids (HX) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2 https://chem.libretexts.org/@go/page/15409
Reduction of Alcohols
Because the most electrophilic site of an alcohol is the hydroxyl proton and because OH- is a poor leaving group, alcohols do not
undergo substitution reactions with nucleophiles. However, they can be converted into tosylates or other sulfate esters, which have
very good (sulfate) leaving groups that are readily displaced in substitution reactions.
Nucleophiles that displace tosylate groups include hydride ions, H-. Therefore, tosylates formed from alcohols undergo
nucleophilic substitution reacctions with hydride sources, such as lithium aluminum hydride (LiAlH , aka LAH ). The net result
4

of the process is the reduction of alcohols to alkanes.


ROH + T sC l → ROT s (1)

ROT s + LiAlH4 → RH (2)

Reduction of Alcohols is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1 https://chem.libretexts.org/@go/page/15731
Replacing the OH Group by Halogen Atoms
This page looks at reactions in which the -OH group in an alcohol is replaced by a halogen such as chlorine or bromine. It includes
a simple test for an -OH group using phosphorus(V) chloride. The general reaction looks like this:
ROH + H X → RX + H2 O (1)

Reaction with hydrogen chloride


Tertiary alcohols react reasonably rapidly with concentrated hydrochloric acid, but for primary or secondary alcohols the reaction
rates are too slow for the reaction to be of much importance. A tertiary alcohol reacts if it is shaken with with concentrated
hydrochloric acid at room temperature. A tertiary halogenoalkane (haloalkane or alkyl halide) is formed.

Replacing -OH by bromine


Rather than using hydrobromic acid,the alcohol is typically treated with a mixture of sodium or potassium bromide and
concentrated sulfuric acid. This produces hydrogen bromide, which reacts with the alcohol. The mixture is warmed to distil off the
bromoalkane.
C H3 C H2 OH + H Br → C H3 C H2 Br + H2 O (2)

Replacing -OH by iodine


In this case, the alcohol is reacted with a mixture of sodium or potassium iodide and concentrated phosphoric(V) acid, H3PO4, and
the iodoalkane is distilled off. The mixture of the iodide and phosphoric(V) acid produces hydrogen iodide, which reacts with the
alcohol.

C H3 C H2 OH + H I → C H3 C H2 I + H2 O (3)

Phosphoric(V) acid is used instead of concentrated sulfuric acid because sulfuric acid oxidizes iodide ions to iodine and produces
hardly any hydrogen iodide. A similar phenomenon occurs to some extent with bromide ions in the preparation of bromoalkanes
but not enough to interfere with the main reaction. There is no reason why you could not use phosphoric(V) acid in the bromide
case instead of sulfuric acid if desired.

Reacting Alcohols with Phosphorus Halides


Alcohols react with liquid phosphorus(III) chloride (also called phosphorus trichloride) to yield chloroalkanes.
3C H3 C H2 C H2 OH + P C l3 → 3C H3 C H2 C H2 C l + H3 P O3 (4)

Alcohols also violently react with solid phosphorus(V) chloride (phosphorus pentachloride) at room temperature, producing clouds
of hydrogen chloride gas. While it is not a good approach to make chloroalkanes, it is a good test for the presence of -OH groups.
To show that a substance was an alcohol, you would first have to eliminate all the other groups that also react with phosphorus(V)
chloride. For example, carboxylic acids (containing the -COOH group) also react with it (because of the -OH in -COOH) as does
water (H-OH).
If you have a neutral liquid not contaminated with water, and clouds of hydrogen chloride are produced when you add
phosphorus(V) chloride, then you have an alcohol group present.
C H3 C H2 C H2 OH + P C l5 → C H3 C H2 C H2 C l + P OC l3 + H C l (5)

There are also side reactions involving the P OC l reacting with the alcohol.
3

Other reactions involving phosphorus halides


Instead of using phosphorus(III) bromide or iodide, the alcohol is usually heated under reflux with a mixture of red phosphorus and
either bromine or iodine. The phosphorus first reacts with the bromine or iodine to give the phosphorus(III) halide.
2 P(s) + 3Br2 → 2P Br3 (6)

1 https://chem.libretexts.org/@go/page/3903
2 P(s) + 3 I2 → 2P I3 (7)

These then react with the alcohol to give the corresponding halogenoalkane, which can be distilled off.
3C H3 C H2 C H2 OH + P Br3 → 3C H3 C H2 C H2 Br + H3 P O3 (8)

3C H3 C H2 C H2 OH + P I3 → 3C H3 C H2 C H2 I + H3 P O3 (9)

Reacting alcohols with Thionyl Chloride


Sulfur dichloride oxide (thionyl chloride) has the formula SOCl2. Traditionally, the formula is written as shown, despite the fact
that the modern name writes the chlorine before the oxygen (alphabetical order). The sulfur dichloride oxide reacts with alcohols at
room temperature to produce a chloroalkane. Sulfur dioxide and hydrogen chloride are given off. Care would have to be taken
because both of these are poisonous.
C H3 C H2 C H2 OH + SOC l2 → C H3 C H2 C H2 C l + S O2 + H C l (10)

The advantage that this reaction has over the use of either of the phosphorus chlorides is that the two other products of the reaction
(sulfur dioxide and HCl) are both gases. That means that they separate themselves from the reaction mixture.

Contributors
Jim Clark (Chemguide.co.uk)

This page titled Replacing the OH Group by Halogen Atoms is shared under a CC BY-NC 4.0 license and was authored, remixed, and/or curated
by Jim Clark.

2 https://chem.libretexts.org/@go/page/3903
The Oxidation of Alcohols
This page looks at the oxidation of alcohols using acidified sodium or potassium dichromate(VI) solution. This reaction is used to
make aldehydes, ketones and carboxylic acids, and as a way of distinguishing between primary, secondary and tertiary alcohols.

Oxidizing the different types of alcohols


The oxidizing agent used in these reactions is normally a solution of sodium or potassium dichromate(VI) acidified with dilute
sulfuric acid. If oxidation occurs, then the orange solution containing the dichromate(VI) ions is reduced to a green solution
containing chromium(III) ions. The electron-half-equation for this reaction is as follows:
2− + − 3+
C r2 O + 14 H + 6e → 2C r + 7 H2 O (1)
7

Primary alcohols
Primary alcohols can be oxidized to either aldehydes or carboxylic acids, depending on the reaction conditions. In the case of the
formation of carboxylic acids, the alcohol is first oxidized to an aldehyde, which is then oxidized further to the acid.
An aldehyde is obtained if an excess amount of the alcohol is used, and the aldehyde is distilled off as soon as it forms. An excess
of the alcohol means that there is not enough oxidizing agent present to carry out the second stage, and removing the aldehyde as
soon as it is formed means that it is not present to be oxidized anyway!
If you used ethanol as a typical primary alcohol, you would produce the aldehyde ethanal, C H C H O. The full equation for this
3

reaction is fairly complicated, and you need to understand the electron-half-equations in order to work it out.
2− + 3+
3C H3 C H2 OH + C r2 O + 8H → 3C H3 C H O + 2C r + 7 H2 O (2)
7

In organic chemistry, simplified versions are often used that concentrate on what is happening to the organic substances. To do that,
oxygen from an oxidizing agent is represented as [O]. That would produce the much simpler equation:

It also helps in remembering what happens. You can draw simple structures to show the relationship between the primary alcohol
and the aldehyde formed.

Full oxidation to carboxylic acids


An excess of the oxidizing agent must be used, and the aldehyde formed as the half-way product should remain in the mixture. The
alcohol is heated under reflux with an excess of the oxidizing agent. When the reaction is complete, the carboxylic acid is distilled
off. The full equation for the oxidation of ethanol to ethanoic acid is as follows:
2− 3+
3C H3 C H2 OH + 2C r2 O + 16H + → 3C H3 C OOH + 4C r + 11 H2 O (3)
7

The more typical simplified version looks like this:


C H3 C H2 OH + 2[O] → C H3 C OOH + H2 O (4)

Alternatively, you could write separate equations for the two stages of the reaction - the formation of ethanal and then its
subsequent oxidation.
C H3 C H2 OH + [O] → C H3 C H O + H2 O (5)

C H3 C H O + [O] → C H3 C OOH (6)

This is what is happening in the second stage:

1 https://chem.libretexts.org/@go/page/3904
Secondary alcohols
Secondary alcohols are oxidized to ketones - and that's it. For example, if you heat the secondary alcohol propan-2-ol with sodium
or potassium dichromate(VI) solution acidified with dilute sulfuric acid, propanone is formed. Changing the reaction conditions
makes no difference to the product. Folloiwng is the simple version of the equation, showing the relationship between the
structures:

If you look back at the second stage of the primary alcohol reaction, you will see that an oxygen inserted between the carbon and
the hydrogen in the aldehyde group to produce the carboxylic acid. In this case, there is no such hydrogen - and the reaction has
nowhere further to go.

Tertiary alcohols
Tertiary alcohols are not oxidized by acidified sodium or potassium dichromate(VI) solution - there is no reaction whatsoever. If
you look at what is happening with primary and secondary alcohols, you will see that the oxidizing agent is removing the hydrogen
from the -OH group, and a hydrogen from the carbon atom is attached to the -OH. Tertiary alcohols don't have a hydrogen atom
attached to that carbon.
You need to be able to remove those two particular hydrogen atoms in order to set up the carbon-oxygen double bond.

Using these reactions as a test for the different types of alcohols


First, the presence of an alcohol must be confirmed by testing for the -OH group. The liquid would need to be verified as neutral,
free of water and that it reacted with solid phosphorus(V) chloride to produce a burst of acidic steamy hydrogen chloride fumes. A
few drops of the alcohol would be added to a test tube containing potassium dichromate(VI) solution acidified with dilute sulfuric
acid. The tube would be warmed in a hot water bath.

Determining the tertiary alcohol


In the case of a primary or secondary alcohol, the orange solution turns green. With a tertiary alcohol, there is no color change.
After heating, the following colors are observed:

2 https://chem.libretexts.org/@go/page/3904
Distinguishing between the primary and secondary alcohols
A sufficient amount of the aldehyde (from oxidation of a primary alcohol) or ketone (from a secondary alcohol) must be produced
to be able to test them. There are various reactions that aldehydes undergo that ketones do not. These include the reactions with
Tollens' reagent, Fehling's solution and Benedict's solution, and these reactions are covered on a separate page.
These tests can be difficult to carry out, and the results are not always as clear-cut as the books say. A much simpler but fairly
reliable test is to use Schiff's reagent. Schiff's reagent is a fuchsin dye decolorized by passing sulfur dioxide through it. In the
presence of even small amounts of an aldehyde, it turns bright magenta.
It must, however, be used absolutely cold, because ketones react with it very slowly to give the same color. If you heat it, obviously
the change is faster - and potentially confusing. While you are warming the reaction mixture in the hot water bath, you can pass any
vapors produced through some Schiff's reagent.

If the Schiff's reagent quickly becomes magenta, then you are producing an aldehyde from a primary alcohol.
If there is no color change in the Schiff's reagent, or only a trace of pink color within a minute or so, then you are not producing
an aldehyde; therefore, no primary alcohol is present.
Because of the color change to the acidified potassium dichromate(VI) solution, you must, therefore, have a secondary alcohol.
You should check the result as soon as the potassium dichromate(VI) solution turns green - if you leave it too long, the Schiff's
reagent might start to change color in the secondary alcohol case as well.

Contributors
Jim Clark (Chemguide.co.uk)

This page titled The Oxidation of Alcohols is shared under a CC BY-NC 4.0 license and was authored, remixed, and/or curated by Jim Clark.

3 https://chem.libretexts.org/@go/page/3904
Oxidation by Chromic Acid
One of the reagents that is commonly used for oxidation in organic chemistry is chromic acid. This reagent is straightforward to use
once deciphered. However, there are a vast number of different ways that textbooks (and instructors) show it being used in
reactions.

Chromic acid, H2CrO4, is a strong acid and a reagent for oxidizing alcohols to ketones and carboxylic acids. For fairly mundane
reasons owing primarily to safety and convenience, chromic acid tends to be made in the reaction vessel as needed (through
addition of acid to a source of chromium), rather than being dispensed from a bottle.
And that’s where the trouble begins. Choosing a source of chromium to make H2CrO4 from is a lot like choosing a favorite brand
of bottled water. Beyond the packaging, they’re pretty much all the same. Depending on which textbook or instructor you have,
however, you might see several different ways to do this, and it can be very confusing.
The key point is that Na2CrO4 (sodium chromate), Na2Cr2O7 (sodium dichromate), K2CrO4(potassium chromate), K2Cr2O7
(potassium dichromate), and CrO3 (chromium trioxide) are all alike in one crucial manner: when they are combined with aqueous
acid, each of them forms H2CrO4, and ultimately it’s H2CrO4 which does the important chemistry. Unfortunately I rarely see this
point explained in textbooks. I remember this causing some confusion for me when I took the course. The K or Na ions present are
just spectators.
Once H2CrO4 is formed, its reactions are pretty straightforward: it converts primary alcohols (and aldehydes) to carboxylic acids
and secondary alcohols to ketones.

1 https://chem.libretexts.org/@go/page/15384
It does this through addition of the alcohol oxygen to chromium, which makes it a good leaving group; a base (water being the
most likely culprit) can then remove a proton from the carbon, forming a new π bond and breaking the O-Cr bond.

Due to its high toxicity, chromic acid tends to find very little use in the organic chemistry laboratory outside of undergrad labs.
There are far more useful reagents out there for performing these transformations.

2 https://chem.libretexts.org/@go/page/15384
Contributors
James Ashenhurst (MasterOrganicChemistry.com)

Oxidation by Chromic Acid is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3 https://chem.libretexts.org/@go/page/15384
Oxidation by PCC (pyridinium chlorochromate)
Pyridinium chlorochromate (PCC) is a milder version of chromic acid.

PCC oxidizes alcohols one rung up the oxidation ladder, from primary alcohols to aldehydes and from secondary alcohols to
ketones. In contrast to chromic acid, PCC will not oxidize aldehydes to carboxylic acids. Similar to or the same as: C rO and 3

pyridine (the Collins reagent) will also oxidize primary alcohols to aldehydes. Here are two examples of PCC in action.
If you add one equivalent of PCC to either of these alcohols, the oxidized version will be produced. The byproducts (featured in
grey) are Cr(IV) as well as pyridinium hydrochloride.
One has to be careful with the amount of water present in the reaction. If water is present, it can add to the aldehyde to create
the hydrate, which could be further oxidized by a second equivalent of PCC if it is present. This is not a concern with ketones
because there is no H directly bonded to C.

How does it work? Oxidation reactions of this sort are actually a type of elimination reaction. The reaction starts with a carbon-
oxygen single bond and results in a carbon-oxygen double bond. The elimination reaction can occur because of the good leaving
group on the oxygen, namely the chromium, which will be displaced when the neighboring C-H bond is broken with a base.

1 https://chem.libretexts.org/@go/page/15407
The first step is attack of oxygen on the chromium to form the Cr-O bond. Secondly, a proton on the (now positive) OH is
transferred to one of the oxygens of the chromium, possibly through the intermediacy of the pyridinium salt. A chloride ion is then
displaced in a reaction reminiscent of a 1,2 elimination reaction to form what is known as a chromate ester.
The C-O double bond is formed when a base removes the proton on the carbon adjacent to the oxygen. [aside: I've drawn the base
as Cl(-) although there are certainly other species which could also act as bases here (such as an alcohol). It is also possible for
pyridine to be used as the base here, although only very low concentrations of the deprotonated form will be present under these
acidic conditions.] The electrons from the C-H bond move to form the C-O bond, and in the process, the O-Cr bond is broken, and
Cr(VI) becomes Cr(IV) (drawn here as O=Cr(OH)2 ).
Real life notes: If you end up using PCC in the lab, do not forget to add molecular sieves or Celite or some other solid to the bottom
of the flask because otherwise you get a nasty brown tar that is difficult to clean up. The toxicity and mess associated with
chromium has spurred the development of other alternatives, such as TPAP, IBX, DMP, and a host of other reagents you generally
do not typically learn about until grad school.

Contributors
James Ashenhurst (MasterOrganicChemistry.com)

Oxidation by PCC (pyridinium chlorochromate) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

2 https://chem.libretexts.org/@go/page/15407
The Reaction Between Alcohols and Sodium
This page describes the reaction between alcohols and metallic sodium,and introduces the properties of the alkoxide that is formed.
We will look at the reaction between sodium and ethanol as being typical, but you could substitute any other alcohol and the
reaction would be the same.

The Reaction between Sodium Metal and Ethanol


If a small piece of sodium is dropped into ethanol, it reacts steadily to give off bubbles of hydrogen gas and leaves a colorless
solution of sodium ethoxide: C H C H ON a. The anion component is an alkoxide.
3 2

− +
2C H3 C H2 OH(l) + 2N a(s) → 2C H3 C H2 O + 2N a + H2(g) (1)
(aq) (aq)

If the solution is evaporated carefully to dryness, then sodium ethoxide (C H C H ON a) is left behind as a white solid. Although
3 2

initially this appears as something new and complicated, in fact, it is exactly the same (apart from being a more gentle reaction) as
the reaction between sodium and water - something you have probably known about for years.
− +
2 H2 O(l) + 2N a(s) → 2OH + 2N a + H2(g) (2)
(aq) (aq)

If the solution is evaporated carefully to dryness, then the sodium hydroxide (N aOH ) is left behind as a white solid.
We normally, of course, write the sodium hydroxide formed as N aOH rather than H ON a - but that's the only difference. Sodium
ethoxide is just like sodium hydroxide, except that the hydrogen has been replaced by an ethyl group. Sodium hydroxide contains
OH

ions; sodium ethoxide contains C H C H O ions.
3 2

 Note
The reason that the ethoxide formula is written with the oxygen on the right unlike the hydroxide ion is simply a matter of
clarity. If you write it the other way around, it doesn't immediately look as if it comes from ethanol. You will find the same
thing happens when you write formulae for organic salts like sodium ethanoate, for example.

There are two simple uses for this reaction:

To safely dispose of small amounts of sodium: If you spill some sodium on the bench or have a small
amount left over from a reaction you cannot simply dispose of it in the sink. It tends to react explosively with the water - and
comes flying back out at you again! It reacts much more gently with ethanol. Ethanol is, therefore, used to dissolve small
quantities of waste sodium. The solution formed can be washed away without problems (provided you remember that sodium
ethoxide is strongly alkaline - see below).
To test for the -OH group in alcohols: Because of the dangers involved in handling sodium, this is not the
best test for an alcohol at this level. Because sodium reacts violently with acids to produce a salt and hydrogen, you would first
have to be sure that the liquid you were testing was neutral. You would also have to be confident that there was no trace of
water present because sodium reacts with the -OH group in water even better than with the one in an alcohol. With those
provisos, if you add a tiny piece of sodium to a neutral liquid free of water and get bubbles of hydrogen produced, then the
liquid is an alcohol.

Ethoxide Ions are Strongly Basic


If you add water to sodium ethoxide, it dissolves to give a colorless solution with a high pH. The solution is strongly alkaline
because ethoxide ions are Brønsted-Lowry bases and remove hydrogen ions from water molecules to produce hydroxide ions,
which increase the pH.
− −
C H3 C H2 O + H2 O → C H3 C H2 OH + OH (3)

Ethoxide Ions are Good Nucleophiles


A nucleophile is a chemical species that carries a negative or partial negative charge that it uses to attack positive centers in other
molecules or ions. Hydroxide ions are good nucleophiles, and you may have come across the reaction between a halogenoalkane
(also called a haloalkane or alkyl halide) and sodium hydroxide solution. The hydroxide ions replace the halogen atom.

1 https://chem.libretexts.org/@go/page/3902
− −
C H3 C H2 C H2 Br + OH → C H3 C H2 C H2 OH + Br (4)

In this case, an alcohol is formed. The ethoxide ion behaves in exactly the same way. If you knew the mechanism for the hydroxide
ion reaction, you could work out exactly what happens in the reaction between a halogenoalkane and ethoxide ion.
Compare this equation with the last one.

C H3 C H2 C H2 OH + C H3 C H2 Br → C H3 C H2 C H2 OC H2 C H3 + Br (5)

The only difference is that where there was a hydrogen atom at the right-hand end of the product molecule, an alkyl group is now
present. Two alkyl (or other hydrocarbon) groups bridged by an oxygen atom is called an ether. This particular one is 1-
ethoxypropane or ethyl propyl ether. This reaction is known as the Williamson Ether Synthesis and is a good method of
synthesizing ethers in the lab.

Contributors

UndefinedNameError: reference to undefined name 'ContribClark' (click for details)

This page titled The Reaction Between Alcohols and Sodium is shared under a CC BY-NC 4.0 license and was authored, remixed, and/or curated
by Jim Clark.

2 https://chem.libretexts.org/@go/page/3902
The Triiodomethane (Iodoform) Reaction
The triiodomethane (iodoform) reaction can be used to identify the presence of a CH3CH(OH) group in alcohols. There are two
apparently quite different mixtures of reagents that can be used to do this reaction, but are chemically equivalent.
Iodine and sodium hydroxide solution
This is chemically the more obvious method. Iodine solution is added to a small amount of an alcohol, followed by just
enough sodium hydroxide solution to remove the color of the iodine. If nothing happens in the cold, it may be necessary to
warm the mixture very gently. A positive result is the appearance of a very pale yellow precipitate of triiodomethane
(previously known as iodoform): CHI3, which apart from its color, can also be recognized by its faintly "medical" smell. It is
used as an antiseptic on the sort of sticky plasters you put on minor cuts, for example.
Using potassium iodide and sodium chlorate(I) solutions
Sodium chlorate(I) is also known as sodium hypochlorite. Potassium iodide solution is added to a small amount of an
alcohol, followed by sodium chlorate(I) solution. Again, if no precipitate is formed in the cold, it may be necessary to warm
the mixture very gently. The positive result is the same pale yellow precipitate as before.

What the Iodoform Reaction Shows


A positive result is the formation of a pale yellow precipitate of triiodomethane (iodoform) - is given by an alcohol containing the
grouping:

"R" can be a hydrogen atom or a hydrocarbon group (for example, an alkyl group). If "R" is hydrogen, then you have the primary
alcohol ethanol, CH3CH2OH.
1. Ethanol is the only primary alcohol to give the triiodomethane (iodoform) reaction.
2. If "R" is a hydrocarbon group, then you have a secondary alcohol. Lots of secondary alcohols give this reaction, but those that
do all have a methyl group attached to the carbon with the -OH group.
3. No tertiary alcohols can contain this group because no tertiary alcohols can have a hydrogen atom attached to the carbon with
the -OH group. No tertiary alcohols give the triiodomethane (iodoform) reaction.

Mechanism
We will assume the iodine/sodium hydroxide solution for the reaction:

1 https://chem.libretexts.org/@go/page/3906
This is being given as a flow scheme rather than full equations. The equations for the other two steps are given elsewhere.

Contributors
Jim Clark (Chemguide.co.uk)

This page titled The Triiodomethane (Iodoform) Reaction is shared under a CC BY-NC 4.0 license and was authored, remixed, and/or curated by
Jim Clark.

2 https://chem.libretexts.org/@go/page/3906
Thionyl Chloride
If There is one thing you learn how to do well in Org 1, it’s make alcohols. Let’s count the ways: hydroboration, acid-catalyzed
hydration, oxymercuration for starters, and then substitution of alkyl halides with water or H O . If you want to extend it even

further, There is dihydroxylation (to make diols) using OsO or cold KM nO , and even opening of epoxides under acidic or
4 4

basic conditions to give alcohols.


There is just one issue here and it comes up once you try to use alcohols in synthesis. Let’s say you want to use that alcohol in a
subsequent substitution step, getting rid of the \(HO^-)\ and replacing it with something else. See any problems with that?
Remember that good leaving groups are weak bases – and the hydroxide ion, being a strong base, tends to be a pretty bad leaving
group.
So what can we do?
What you want to do is convert the alcohol into a better leaving group. One way is to convert the alcohol into a sulfonate ester – we
talked about that with T sC l and M sC l. Alternatively, alcohols can be converted into alkyl chlorides with thionyl chloride (
SOC l ). This is a useful reaction, because the resulting alkyl halides are versatile compounds that can be converted into many
2

compounds that are not directly accessible from the alcohol itself.

If you take an alcohol and add thionyl chloride, it will be converted into an alkyl chloride. The byproducts here are hydrochloric
acid (H C l) and sulfur dioxide (SO ). Note: there are significant differences in how this reaction is taught at different schools.
2

Consult your instructor to be 100% sure that this applies to your course). See post here

 Example 1 : Conversion of Alcohols to Alkyl Chlorides

There is one important thing to note here: see the stereochemistry? It’s been inverted.*(white lie alert – see below) That’s an
important difference between SOC l and TsCl, which leaves the stereochemistry alone. We’ll get to the root cause of that in a
2

moment, but in the meantime, can you think of a mechanism which results in inversion of configuration at carbon?

As an extra bonus, thionyl chloride will also convert carboxylic acids into acid chlorides (“acyl chlorides”). Like alcohols,
carboxylic acids have their limitations as reactants: the hydroxyl group interferes with many of the reactions we learn for
nucleophilic acyl substitution (among others). Conversion of the OH into Cl solves this problem.

1 https://chem.libretexts.org/@go/page/15405
 Example 2 : Conversation of Carboxylic Acids to Acid Chlorides

Mechanism
As you might have guessed, conversion of alcohols to alkyl halides proceeds through a substitution reaction – specifically, an S 2 N

mechanism. The first step is attack of the oxygen upon the sulfur of SOC l , which results in displacement of chloride ion. This has
2

the side benefit of converting the alcohol into a good leaving group: in the next step, chloride ion attacks the carbon in S 2 N

fashion, resulting in cleavage of the C–O bond with inversion of configuration. The H OSC l breaks down into H C l and sulfur
dioxide gas, which bubbles away.

Formation of Alkyl Chlorides


Since the reaction proceeds through a backside attack (SN 2 ), there is inversion of configuration at the carbon

The mechanism for formation of acid chlorides from carboxylic acids is similar. The conversion of caboxylic acids to acid
chlorides is similar, but proceeds through a [1,2]-addition of chloride ion to the carbonyl carbon followed by [1,2]-elimination to
give the acid chloride, SO and H C l
2

2 https://chem.libretexts.org/@go/page/15405
In the laboratory
Like many sulfur-containing compounds, thionyl chloride is noseworthy for its pungent smell. Thionyl chloride has a nauseating
sickly-sweet odor to it that imprints itself forever upon your memory. One accident that occurred during my time as a TA involved
a student dropping a flask with 5 mL of thionyl chloride into a rotovap bath outside the fume hood. The cloud of SO and H C l 2

that formed cleared the teaching lab for half an hour, so you can imagine what thionyl chloride would do if exposed to the moisture
in your lungs. Treat with caution, just as you would if you were working with phosgene.
*Here’s the white lie. Although it’s often taught that SOC l leads to 100% inversion of configuration, in reality it’s not always that
2

simple. Inversion of configuration with SOC l is very solvent dependent. Depending on the choice of solvent, one can get either
2

straight inversion, or a mixture of retention and inversion. For the purposes of beginning organic classes, most students can ignore
this message.

Contributors
James Ashenhurst (MasterOrganicChemistry.com)

Thionyl Chloride is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3 https://chem.libretexts.org/@go/page/15405

You might also like