Basic Name Reaction
Basic Name Reaction
net/publication/340861231
CITATIONS READS
0 26,259
1 author:
Bhavesh Socha
Sardar Patel University
22 PUBLICATIONS 26 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Bhavesh Socha on 01 May 2020.
!"#$
%
"
!
&
' &
(
)*!
&
"
"*)+,-!
."
/&
001
2
'"
. 3,3, %
"
!
&
' & (
BASIC ORGANIC
NAME
REACTION
ϭ
INDEX
-RQHV2[LGDWLRQ
:LOOLDPVRQ6\QWKHVLV
:XUW]5HDFWLRQ
%LUFKUHGXFWLRQ
'LHOV$OGHUUHDFWLRQ
(QHUHDFWLRQ
2]RQRO\VLVUHDFWLRQ
6LPPRQV±VPLWKUHDFWLRQ
=LHJOHU1DWWD&DWDO\VW
&OHPPHQVHQ5HGXFWLRQ
%DH\HUYLOODJHUR[LGDWLRQ
$OGRO&RQGHQVDWLRQ
%HFNPDQQ5HDFWLRQUHDUUDQJHPHQW
&DQQL]]DURUHDFWLRQ
&XUWLXVUHDUUDQJHPHQW
6FKPLGW5HDUUDQJHPHQW
&ODLVHQ6FKPLGW&RQGHQVDWLRQ
'DNLQ5HDFWLRQ
:ROII.LVKQHU5HGXFWLRQ5HDFWLRQ
)DYRUVNLLUHDUUDQJHPHQW
&RUH\+RXVH6\QWKHVLV
.QRHYHQDJHO&RQGHQWDWLRQUHDFWLRQ
5RVHQPXQGUHDFWLRQ
3HUNLQ&RQGHQVDWLRQUHDFWLRQ
3LQDFRO3LQDFRORQHUHDUUDQJHPHQW
0HHUZHLQ3RQQGRUI9HUOH\039UHGXFWLRQ
/RVVHQ5HDUUDQJHPHQWUHDFWLRQ
6\QWKHVLVRIDOGHK\GHE\6WHSKHQ¶VPHWKRG
(VFKZHLOHU±&ODUNHPHWK\ODWLRQ
&\DQLGHWRDPLGHLQSUHVHQFHRI+&O
*DEULHO3KWKDOLPLGHV\QWKHVLV
0DQQLFKUHDFWLRQ
0HWK\ODWLRQE\'LD]RPHWKDQH
3DWHUQR%XFKL5HDFWLRQ
*LOOPDQUHDJHQW
5LWWHUUHDFWLRQ
*DWWHUPDQ.RFKUHDFWLRQ
5HLPHU7LHPDQQUHDFWLRQ
8OOPDQQV\QWKHVLV
:XUW])LWWLJUHDFWLRQ
6DQGPH\HUUHDFWLRQ
&ODLVHQUHDUUDQJHPHQW
)ULHVUHDUUDQJHPHQW
%LVFKOHU1DSLHUDOVNL5HDFWLRQ
3LFWHW±*DPVUHDFWLRQ
Ϯ
;ϰϲͿ <ŽůďĞʹ^ĐŚŵŝƚƚƌĞĂĐƚŝŽŶ ϲϴ
;ϰϳͿ ůĂŝƐĞŶŽŶĚĞŶƐĂƚŝŽŶ ϲϵ
;ϰϴͿ ^ƚŽďďĞĐŽŶĚĞŶƐĂƚŝŽŶ ϳϬ
;ϰϵͿ ,ĞůůsŽůŚĂƌĚĞůŝŶƐŬLJZĞĂĐƚŝŽŶ;,sZĞĂĐƚŝŽŶͿ ϳϭ
;ϱϬͿ ƚĂƌĚƌĞĂĐƚŝŽŶ ϳϮ
ϯ
;ϭͿ :ŽŶĞƐKdžŝĚĂƚŝŽŶ
The Jones oxidation is an organic reaction for the oxidation of primary and
secondary alcohols to carboxylic acids and ketones, respectively. It is named after its
discoverer, Sir Ewart Jones.
Jones reagent consists of chromium trioxide and sulfuric acid dissolved in a mixture
of acetone and water. As an alternative, potassium dichromate can be used in place of
chromium trioxide. The oxidation is very rapid, quite exothermic, and the yields are
typically high. The reagent rarely oxidizes unsaturated bonds.
Application
Although useful reagent for some applications, due to the carcinogenic nature of
chromium(VI), the Jones oxidation has slowly been replaced by other oxidation
methods. It remains useful in organic synthesis. A variety of spectroscopic techniques,
including IR can be used to monitor the progress of a Jones oxidation reaction and
confirm the presence of the oxidized product. At one time the Jones oxidation was used
in primitive breathalyzers. Aminoindans, which are of pharmalogical interest, are
prepared by the oxidation of the alcohol to ketone which is converted into an amino
group. The alcohol is oxidized to the ketone with the Jones reagent. The reagent was
once used to prepare salicylic acid, a precursor to aspirin. Methcathinone is a
psychoactive stimulant that is sometimes used as an addictive recreational drug. It can
be oxidized from certain alcohols using the Jones reagent.
ϰ
Related processes
Several other chromium compounds are used for the oxidation of alcohols.[3] These
include Collins reagent and pyridinium chlorochromate.The Sarett oxidation is a
similar process.
The Williamson ether synthesis is an organic reaction, forming an ether from
an organohalide and a deprotonated alcohol (alkoxide). This reaction was developed
by Alexander Williamson in 1850.[2] Typically it involves the reaction of an alkoxide
ion with a primary alkyl halide via an SN2 reaction. This reaction is important in the
history of organic chemistry because it helped prove the structure of ethers.
Conditions
Since alkoxide ions are highly reactive, they are usually prepared immediately
prior to the reaction, or are generated in situ. In laboratory chemistry, in situ generation
is most often accomplished by the use of a carbonate base or potassium hydroxide,
while in industrial syntheses phase transfer catalysis is very common. A wide range of
solvents can be used, but protic solvents and apolar solvents tend to slow the reaction
rate strongly, as a result of lowering the availability of the free nucleophile. For this
reason, acetonitrile and N,N-dimethylformamide are particularly commonly used.
A typical Williamson reaction is conducted at 50 to 100 °C and is complete in
1 to 8 h. Often the complete disappearance of the starting material is difficult to achieve,
ϱ
x In liquid ammonia alkali metals dissolve to give a blue solution thought of
simplistically as having "free electrons". The electrons are taken up by the
aromatic ring, one at a time. Once the first electron has been absorbed, a
radical anion has been formed. Next the alcohol molecule donates its
hydroxylic hydrogen to form a new C±H bond; at this point a radical has been
formed. This is followed by the second electron being picked up to give
a carbanion of the cyclohexadienyl type (i.e. with C=C±C±C=C in a six-
membered ring with negative charge). Then this cyclohexadienyl anion
is protonated by the alcohol present. The protonation takes place in the center
of the cyclohexadienyl system. This (regio-)selectivity is characteristic.
x Where the radical anion is initially protonated determines the structure of the
product. With an electron donor such as methoxy (MeO) or with an alkyl
group, protonation has been thought by some investigators as being ortho (i.e.
adjacent or 1,2) to the substituent. Other investigators have thought the
protonation is meta (1,3) to the substituent. Arthur Birch
favored meta protonation. With electron withdrawing substituents, protonation
has been thought to occur at the site of the substituent (ipso), or para (1,4).
Again, there has been varied opinion. A. J. Birch's empirical rules say that for
the donor substituents the final product has the maximum number of
substituents on the final double bonds. For electron withdrawing groups the
double bonds of the product avoid the substituents. The placement preference
of groups in the mechanism and in the final product is termed regioselectivity.
x reaction mechanism provides the details of molecular change as a reaction
proceeds. In the case of donating groups, A. J. Birch's preference
for meta protonation of the radical anion was based on qualitative reasoning,
but this has not been experimentally demonstrated.
x In 1961 a simple computation of the electron densities of the radical anion
revealed that it was the ortho site which was most negative and thus most
likely to protonate. Additionally, the second protonation was determined
computationally to occur in the center of the cyclohexadienyl anion to give an
unconjugated product.
x The uncertainty in the chemical literature is now only of historical
significance. Indeed, some further computational results have been reported,
ϭϭ
which vary from suggesting a preference for meta radical-anion protonation to
suggesting a mixture of ortho and meta protonation.[citation needed]
x In 1990 and 1993 an esoteric test was devised which showed
that ortho protonation of the radical anion was preferred over meta (seven to
one).[citation needed] This was accompanied by more modern computation which
concurred. Both experiment and computations were in agreement with the
early 1961 computations.
x With electron withdrawing groups there are examples in the literature
demonstrating the nature of the carbanion just before final protonation,[citation
needed]
revealing that the initial radical-anion protonation occurs para to the
withdrawing substituent.
x The remaining item for discussion is the final protonation of the
cyclohexadienyl anion. In 1961 it was found that simple Hückel computations
were unable to distinguish between the different protonation sites.[citation
needed]
However, when the computations were modified with somewhat more
realistic assumptions, the Hückel computations revealed the center carbon to
the preferred. The more modern 1990 and 1993 computations were in
agreement.
ϭϮ
(5) Diels-Alder reaction
ϭϯ
involving X groups on both diene and dienophile, a 1,3-substitution pattern may be favored,
an outcome not accounted for by a simplistic resonance structure argument. However,
cases where the resonance argument and the matching of largest orbital coefficients
disagree are rare.
ϭϱ
(6) Ene reaction
The ene reaction (also known as the Alder-ene reaction by its discoverer Kurt Alder in
1943) is a chemical reaction between an alkene with an allylic hydrogen (the ene) and
a compound containing a multiple bond (the enophile), LQRUGHUWRIRUPDQHZı-bond
with migration of the ene double bond and 1,5 hydrogen shift. The product is a
substituted alkene with the double bond shifted to the allylic position.
ϭϲ
(7) Ozonolysis
Ozonolysis of alkenes
ϭϳ
more slowly than with the intended ozonolysis target. The ozonolysis of the indicator,
which causes a noticeable color change, only occurs once the desired target has been
consumed. If the substrate has two alkenes that react with ozone at different rates, one
can choose an indicator whose own oxidation rate is intermediate between them, and
therefore stop the reaction when only the most susceptible alkene in the substrate has
reacted.[6] Otherwise, the presence of unreacted ozone in solution (seeing its blue color)
or in the bubbles (via iodide detection) only indicates when all alkenes have reacted.
After completing the addition, a reagent is then added to convert the
intermediate ozonide to a carbonyl derivative. Reductive work-up conditions are far
more commonly used than oxidative conditions. The use
of triphenylphosphine, thiourea, zinc dust, or dimethyl sulfide produces aldehydes or
ketones while the use of sodium borohydride produces alcohols. The use of hydrogen
peroxide produces carboxylic acids. Recently, the use of amine N-oxides has been
reported to produce aldehydes directly.[7] Other functional groups, such
as benzyl ethers, can also be oxidized by ozone. It has been proposed that small
amounts of acid may be generated during the reaction from oxidation of the solvent,
so pyridine is sometimes used to buffer the reaction. Dichloromethane is often used as
a 1:1 cosolvent to facilitate timely cleavage of the ozonide. Azelaic acid and pelargonic
acids are produced from ozonolysis of oleic acid on an industrial scale.
ϭϴ
Limitations
ϭϵ
(9) Ziegler- Natta Catalyst
A Ziegler±Natta catalyst, named after Karl Ziegler and Giulio Natta, is a catalyst used
in the synthesis of polymers of 1-alkenes (alpha-olefins). Two broad classes of
Ziegler±Natta catalysts are employed, distinguished by their solubility:
x Heterogeneous supported catalysts based on titanium compounds are used in
polymerization reactions in combination with
cocatalysts, organoaluminium compounds such as triethylaluminium,
Al(C2H5)3. This class of catalyst dominates the industry.
x Homogeneous catalysts usually based on complexes of Ti, Zr or Hf. They are
usually used in combination with a different organoaluminium
cocatalyst, methyl aluminoxane (or methyl alumoxane, MAO). These catalysts
traditionally contains metallocene but also feature multidentate oxygen- and
nitrogen-based ligands.
ϮϬ
Ziegler±Natta catalysts are used to polymerize terminal alkenes (ethylene and
alkenes with the vinyl double bond):
;ϭϬͿ Clemmensen Reduction
Clemmensen reduction is a chemical reaction described as a reduction of Ketones
(or aldehydes) to alkanes using zinc amalgam and concentrated hydrochloric acid.
This reaction is named after Erik Christian Clemmensen, a Danish chemist.
The original Clemmensen reduction conditions are particularly effective at
reducing aryl-alkyl ketones, such as those formed in a Friedel-Crafts acylation. The
two-step sequence of Friedel-Crafts acylation followed by Clemmensen reduction
constitutes a classical strategy for the primary alkylation of arenes. With aliphatic or
Ϯϭ
cyclic ketones, modified Clemmensen conditions using activated zinc dust in an
anhydrous solution of hydrogen chloride in diethyl ether or acetic anhydride is much
more effective.
The substrate must be tolerant of the strongly acidic conditions of the Clemmensen
reduction (37% HCl). Several alternatives are available. Acid-sensitive substrates that
are stable to strong base can be reduced using the Wolff-Kishner reduction; a further,
milder method for substrates stable to hydrogenolysis in the presence of Raney
nickel is the two-step Mozingo reduction.
In spite of the antiquity of this reaction, the mechanism of the Clemmensen reduction
remains obscure. Due to the heterogeneous nature of the reaction, mechanistic studies
are difficult, and only a handful of studies have been disclosed.Mechanistic proposals
generally invoke organozinc intermediates, sometimes including zinc carbenoids,
either as discrete species or as organic fragments bound to the zinc metal surface.
However, the corresponding alcohol is believed not to be an intermediate, since
subjection of the alcohol to Clemmensen conditions generally does not afford the
alkane product.
(11) Baeyer villager oxidation
ϮϮ
Mechanism
Condensation types
Ϯϱ
x A Dieckmann condensation involves two ester groups in the same
molecule and yields a cyclic molecule
x A Henry reaction involves an aldehyde and an aliphatic nitro compound.
x A Robinson annulation LQYROYHVDQĮȕ-unsaturated ketone and
a carbonyl group, which first engage in a Michael reaction prior to the aldol
condensation.
x In the Guerbet reaction, an aldehyde, formed in situ from an alcohol, self-
condenses to the dimerized alcohol.
x In the Japp±Maitland condensation water is removed not by an elimination
reaction but by a nucleophilic displacement.
(13) Beckmann Reaction (rearrangement)
Ϯϲ
(16) Schmidt Rearrangement
The Schmidt reaction is an organic reaction in which an azide reacts with a carbonyl group
to give an amine or amide, with expulsion of nitrogen. It is named after Sir Karl Friedrich
Schmidt.
(1)
(2)
ϯϭ
(17) Claisen-Schmidt Condensation
dŚĞ ƌĞĂĐƚŝŽŶ ďĞƚǁĞĞŶ ĂŶ ĂůĚĞŚLJĚĞ Žƌ ŬĞƚŽŶĞ ŚĂǀŝŶŐ ĂŶ ĂůƉŚĂͲŚLJĚƌŽŐĞŶ ǁŝƚŚ ĂŶ ĂƌŽŵĂƚŝĐ
ĐĂƌďŽŶLJůĐŽŵƉŽƵŶĚůĂĐŬŝŶŐĂŶĂůƉŚĂŚLJĚƌŽŐĞŶŝƐĐĂůůĞĚƚŚĞůĂŝƐĞŶʹ^ĐŚŵŝĚƚĐŽŶĚĞŶƐĂƚŝŽŶ͘
/ŶĐĂƐĞƐǁŚĞƌĞƚŚĞƉƌŽĚƵĐƚĨŽƌŵĞĚƐƚŝůůŚĂƐƌĞĂĐƚŝǀĞĂůƉŚĂŚLJĚƌŽŐĞŶĂŶĚĂŚLJĚƌŽdžŝĚĞĂĚũĂĐĞŶƚ
ƚŽĂŶĂƌŽŵĂƚŝĐƌŝŶŐ͕ƚŚĞƌĞĂĐƚŝŽŶǁŝůůƋƵŝĐŬůLJƵŶĚĞƌŐŽĚĞŚLJĚƌĂƚŝŽŶůĞĂĚŝŶŐƚŽƚŚĞĐŽŶĚĞŶƐĂƚŝŽŶ
ƉƌŽĚƵĐƚ͘
ϯϮ
(18) Dakin Reaction
Dakin Reaction is the replacement of the aldehyde group of ortho and para hydroxy
and ortho amino-benzaldehyde (or ketone) by a hydroxyl group on reaction with
alkaline hydrogen peroxide.
ϯϯ
(19) Wolff Kishner Reduction Reaction
The Favorskii rearrangement, named for the Russian chemist Alexei Yevgrafovich
Favorskii, is most principally a rearrangement of cyclopropanones and Į-halo
ketones which leads to carboxylic acid GHULYDWLYHV,QWKHFDVHRIF\FOLFĮ-halo ketones,
the Favorskii rearrangement constitutes a ring contraction. This rearrangement takes
place in the presence of a base, sometimes hydroxide, to yield a carboxylic acid but
most of the time either an alkoxide base or an amine to yield an ester or an amide,
UHVSHFWLYHO\ ĮĮ¶-Dihaloketones eliminate HX under the reaction conditions to give
Įȕ-unsaturated carbonyl compounds.
ϯϰ
(23) Rosenmund reaction
ϯϳ
(25) Pinacol-Pinacolone rearrangement
Mechanism
ϯϵ
In the course of this organic reaction, protonation of one of the ±OH groups occurs and
a carbocation is formed. If both the ±OH groups are not alike, then the one which yields
a more stable carbocation participates in the reaction. Subsequently, an alkyl group
from the adjacent carbon migrates to the carbocation center. The driving force for this
rearrangement step is believed to be the relative stability of the resultant oxonium ion,
which has complete octet configuration at all centers (as opposed to the preceding
carbocation). The migration of alkyl groups in this reaction occurs in accordance with
their usual migratory aptitude, i.e.hydride > phenyl carbanion > tertiary carbanion
(if formed by migration) > secondary carbanion (if formed by migration) > methyl
carbanion . {Why Carbanion? Because every migratory group leaves by taking
electron pair with it.} The conclusion is that the group which stabilizes the carbocation
more effectively is migrated.
ϰϬ
(26) Meerwein-Ponndorf-Verley (MPV) reduction
The MPV reduction was discovered by Meerwein and Schmidt, and separately by
Verley in 1925. They found that a mixture of aluminium ethoxide and ethanol could
reduce aldehydes to their alcohols.[2][3] Ponndorf applied the reaction to ketones and
upgraded the catalyst to aluminium isopropoxide in isopropanol.
ϰϭ
(28) 6\QWKHVLVRIDOGHK\GHE\6WHSKHQ¶VPHWKRG
Mechanism
By addition of hydrogen chloride the used nitrile (1) reacts to its corresponding salt
(2). It is believed that this salt is reduced by a single electron transfer by the tin(II)
chloride (3a and 3b). The resulting salt (4) precipitates after some time as aldimine tin
chloride (5). Hydrolysis of 5 produces a amide (6) from which an aldehyde (7) is
formed.
Substitutes that increase the electron density promote the formation of the aldimin-tin
chloride adduct. By electron withdrawing substituents, the formation of amide
chloride is facilitated. In the past, the reaction was carried out by precipitating the
aldimine-tin chloride, washing it with ether and then hydrolyzing it. However, it has
been found that this step is unnecessary and the aldimine tin chloride can be
hydrolysed directly in the solution.
This reaction is more efficient when aromatic nitriles are used instead of aliphatic
ones. However, even for some aromatic nitriles (e. g. 2-formylbenzoic acid ethyl
ester) the yield can be low.
ϰϰ
alkyl amine. However, aryl amines cannot be prepared via Gabriel synthesis as aryl
halides doQ¶WXQGHUJRVLPSOHQXFOHRSKLOLFVXEVWLWXWLRQ
Mechanism
Step 1
Step 2
The nucleophilic imide ion attacks the electrophilic carbon of the alkyl halide. The
nitrogen atom subsequently replaces the halogen (Fluorine, Chlorine, Bromine or
Iodine) in the alkyl halide and bonds with the carbon itself. This results in the formation
of an N-Alkyl Phthalimide.
ϰϳ
(33) Methylation by Diazomethane
Use
ϱϬ
Safety
Diazomethane is toxic by inhalation or by contact with the skin or eyes (TLV 0.2ppm).
Symptoms include chest discomfort, headache, weakness and, in severe cases,
collapse. Symptoms may be delayed. Deaths from diazomethane poisoning have been
reported. In one instance a laboratory worker consumed a hamburger near a fumehood
where he was generating a large quantity of diazomethane, and died four days later
from fulminating pneumonia.[15] Like any other alkylating agent it is expected to be
carcinogenic, but such concerns are overshadowed by its serious acute toxicity.
CH2N2 may explode in contact with sharp edges, such as ground-glass joints, even
scratches in glassware.[16] Glassware should be inspected before use and preparation
should take place behind a blast shield. Specialized kits to prepare diazomethane with
flame-polished joints are commercially available.
The compound explodes when heated beyond 100 °C, exposed to intense light, alkali
metals, or calcium sulfate. Use of a blast shield is highly recommended while using this
compound.
Proof-of-concept work has been done with microfluidics, in which continuous point-
of-use synthesis from N-methyl-N-nitrosourea and 0.93M potassium hydroxide in
water was followed by point-of-use conversion with benzoic acid, resulting in a 65%
yield of the methyl benzoate ester within seconds at temperatures ranging from 0-50 C.
The yield was better than under capillary conditions; the microfluidics were credited
with "suppression of hot spots, low holdup, isothermal conditions, and intensive
mixing."
The Paternò±Büchi reaction, named after Emanuele Paternò and George Büchi who
established its basic utility and form, is a photochemical reaction that forms four-
membered oxetane rings from a carbonyl and an alkene.
ϱϭ
The Ritter reaction is a chemical reaction that transforms a nitrile into an N-
alkyl amide using various electrophilic alkylating reagents. The original reaction
formed the alkylating agent using an alkene or alcohol in the presence of a strong
acid: The reaction has been the subject of several literature reviews.
Applications
The Ritter reaction is most useful in the formation of amides in which the nitrogen has
a tertiary alkyl group. It is also used in industrial processes as it can be effectively scaled
up from laboratory experiments to large-scale applications while maintaining high
yield. Real world applications include Merck's industrial-scale synthesis of anti-
HIV drug Crixivan (indinavir); the production of the falcipain-2 inhibitor PK-11195;
the synthesis of the alkaloid aristotelone; and synthesis of Amantadine, an antiviral and
antiparkinsonian drug. Other applications of the Ritter reaction include synthesis
of dopamine receptor ligands and licit and illicit production of racemic
amphetamine from allylbenzene and methyl cyanide.
A problem with the Ritter reaction is the necessity of an extremely strong
acid catalyst in order to produce the carbocation. This poses a safety risk when running
the reaction and makes disposal of waste products difficult. However, other methods
have been proposed in order promote carbocation formation, including
[19]
photocatalytic electron transfer or direct photolysis.
Example
ϱϰ
(37) Gatterman-Koch reaction
This reaction doesn't work on unactivated (even the least unactivated). For example,
we absolutely couldn't carry this reaction on a nitro-benzene. The utilization of
cuprous chloride isn't always necessary.
Mechanism
Step 1
Step 2
Step 3
ϱϱ
Step 4
ϱϲ
(38) Reimer-Tiemann reaction
Mechanism
ϱϳ
(39) Ullmann synthesis
The Ullmann reaction or Ullmann coupling is a coupling reaction between aryl halides
Mechanism
ϱϴ
(40) Wurtz- Fittig reaction
The reaction works best for forming asymmetrical products if the halide reactants are
somehow separate in their relative chemical reactivities. One way to accomplish this is
to form the reactants with halogens of different periods. Typically the alkyl halide is
made more reactive than the aryl halide, increasing the probability that the alkyl halide
will form the organosodium bond first and thus act more effectively as
a nucleophile toward the aryl halide. Typically the reaction is used for the alkylation of
aryl halides; however, with the use of ultrasound the reaction can also be made useful
for the production of biphenyl compounds.
Mechanism
ϱϵ
(41) Sandmeyer reaction
The Sandmeyer reaction is a chemical reaction used to synthesize aryl halides from
aryl diazonium salts using copper salts as reagents or catalysts. It is an example of
a radical-nucleophilic aromatic substitution. The Sandmeyer reaction provides a
method through which one can perform unique transformations on benzene, such
as halogenation, cyanation, trifluoromethylation, and hydroxylation.
ϲϬ
(42) Claisen rearrangement
The Claisen rearrangement is an exothermic, concerted (bond cleavage and
recombination) pericyclic reaction. Woodward±Hoffmann rules show a suprafacial,
stereospecific reaction pathway. The kinetics are of the first order and the whole
transformation proceeds through a highly ordered cyclic transition state and is
intramolecular. Crossover experiments eliminate the possibility of the rearrangement
occurring via an intermolecular reaction mechanism and are consistent with an
intramolecular process.
There are substantial solvent effects observed in the Claisen rearrangement, where
polar solvents tend to accelerate the reaction to a greater extent. Hydrogen-bonding
solvents gave the highest rate constants. For example, ethanol/water solvent mixtures
give rate constants 10-fold higher than sulfolane. Trivalent organoaluminium reagents,
such as trimethylaluminium, have been shown to accelerate this reaction.
ϲϰ
substitution reaction is temperature dependent. A low reaction temperature favours para
substitution and with high temperatures the ortho product prevails, this can be
rationalised as exhibiting classic Thermodynamic versus kinetic reaction control as the
ortho product can form a more stable bidentate complex with the
Aluminium. Formation of the ortho product is also favoured in non-polar solvents; as
the solvent polarity increases, the ratio of the para product also increases.
Mechanisms
Two types of mechanisms have appeared in the literature for the Bischler±Napieralski
reaction. Mechanism I ŝŶǀŽůǀĞ a dichlorophosphoryl imine-ester intermediate, while
Mechanism II involves a nitrilium ion intermediate (both shown in brackets). This
mechanistic variance stems from the ambiguity over the timing for the elimination of
the carbonyl oxygen in the starting amide. In Mechanism I, the elimination occurs
with imine formation after cyclization; while in Mechanism II, the elimination yields
the nitrilium intermediate prior to cyclization. Currently, it is believed that different
reaction conditions affect the prevalence of one mechanism over the other
(see reaction conditions).
In certain literature, Mechanism II is augmented with the formation of an imidoyl
chloride intermediate produced by the substitution of chloride for the Lewis
ϲϲ
Limitations
The Étard reaction is most commonly used as a relatively easy method of
converting toluene into benzaldehyde. Obtaining specific aldehyde products from
reagents other than toluene tends to be difficult due to rearrangements. For example, n-
propylbenzene is oxidized to propiophenone, benzyl methyl ketone, and several
chlorinated products, with benzyl methyl ketone being the major product. Another
example arises from the Étard reaction of trans-decalin which results in a mixture of
trans-9-decalol, spiro [4.5]decan-6-one, trans-1-decalone, cis-1-decalone, 9,10-octal-1-
one, and 1-tetralone.
Other oxidation reagents like potassium permanganate or potassium
dichromate oxidize to the more stable carboxylic acids.
Uses
ϳϯ
Reference
(1) Wade, L. G. (2005). Organic Chemistry (6th ed.). Upper Saddle River, New Jersey:
Prentice Hall. pp. 1056±66. ISBN 978-0-13-236731-8.
(2) Smith, M. B.; March, J. (2001). Advanced Organic Chemistry (5th ed.). New York:
Wiley Interscience. pp. 1218±23. ISBN 978-0-471-58589-3.
(3) Mahrwald, R. (2004). Modern Aldol Reactions, Volumes 1 and 2. Weinheim,
Germany: Wiley-VCH Verlag GmbH & Co. KGaA. pp. 1218±23. ISBN 978-3-527-
30714-2.
(4) Wurtz, C. A. (1872). "Sur un aldéhyde-alcool" [On an aldehyde alcohol]. Bulletin de
la Société Chimique de Paris. 2nd series (in French). 17: 436±442.
(5) Wurtz, C. A. (1872). "Ueber einen Aldehyd-Alkohol" [About an aldehyde
alcohol]. Journal für Praktische Chemie (in German). 5 (1): 457±
464. doi:10.1002/prac.18720050148.
(6) Wurtz, C. A. (1872). "Sur un aldéhyde-alcool" [On an aldehyde alcohol]. Comptes
rendus de l'Académie des sciences (in French). 74: 1361.
(7) Heathcock, C. H. (1991). "The Aldol Reaction: Acid and General Base Catalysis".
In Trost, B. M.; Fleming, I. (eds.). Comprehensive Organic Synthesis. 2. Elsevier
Science. pp. 133±179. doi:10.1016/B978-0-08-052349-1.00027-5. ISBN 978-0-08-
052349-1.
(8) Mukaiyama T. (1982). The Directed Aldol Reaction. Org. React. 28. pp. 203±
331. doi:10.1002/0471264180.or028.03. ISBN 978-0471264187.
(9) Paterson, I. (1988). "New Asymmetric Aldol Methodology Using Boron
Enolates". Chem. Ind. 12: 390±394.
(10) Mestres R. (2004). "A green look at the aldol reaction". Green Chemistry. 6 (12): 583±
603. doi:10.1039/b409143b.
(11) M. Braun; R. Devant (1984). "(R) and (S)-2-acetoxy-1,1,2-triphenylethanol ± effective
synthetic equivalents of a chiral acetate enolate". Tetrahedron Letters. 25 (44): 5031±
4. doi:10.1016/S0040-4039(01)91110-4.
(12) Jie Jack Li; et al. (2004). Contemporary Drug Synthesis. Wiley-Interscience. pp. 118±
. ISBN 978-0-471-21480-9.
(13) Wulff W. D.; Andersson B. A (1994). "Stereoselective aldol addition reactions of
Fischer carbene complexes via electronic tuning of the metal center for enolate
ϳϰ
reactivity". Inorganica Chimica Acta. 220 (1±2): 215±231. doi:10.1016/0020-
1693(94)03874-0.
(14) Schetter, B.; Mahrwald, R. (2006). "Modern Aldol Methods for the Total Synthesis of
Polyketides". Angew. Chem. Int. Ed. 45 (45): 7506±
7525. doi:10.1002/anie.200602780. PMID 17103481.
(15) Guthrie, J.P.; Cooper, K.J.; Cossar, J.; Dawson, B.A.; Taylor, K.F. (1984). "The
retroaldol reaction of cinnamaldehyde". Can. J. Chem.62 (8): 1441±
1445. doi:10.1139/v84-243.
(16) Zimmerman, H. E.; Traxler, M. D. (1957). "The Stereochemistry of the Ivanov and
Reformatsky Reactions. I". Journal of the American Chemical Society. 79 (8): 1920±
1923. doi:10.1021/ja01565a041.
(17) Heathcock, C. H.; Buse, C. T.; Kleschnick, W. A.; Pirrung, M. C.; Sohn, J. E.; Lampe,
J. (1980). "Acyclic stereoselection. 7. Stereoselective synthesis of 2-alkyl-3-hydroxy
carbonyl compounds by aldol condensation". Journal of Organic Chemistry. 45 (6):
1066±1081. doi:10.1021/jo01294a030.
(18) Bal, B.; Buse, C. T.; Smith, K.; Heathcock, C. H., (2SR,3RS)-2,4-Dimethyl-3-
Hydroxypentanoic Acid, Org. Synth., Coll. Vol. 7, p.185 (1990); Vol. 63, p.89 (1985).
(19) Jump up to:a b Brown, H. C.; Dhar, R. K.; Bakshi, R. K.; Pandiarajan, P. K.; Singaram,
B. (1989). "Major effect of the leaving group in dialkylboron chlorides and triflates in
controlling the stereospecific conversion of ketones into either E- or Z-enol
borinates". Journal of the American Chemical Society. 111 (9): 3441±
3442. doi:10.1021/ja00191a058.
(20) Ireland, R. E.; Willard, A. K. (1975). "The stereoselective generation of ester
enolates". Tetrahedron Letters. 16 (46): 3975±3978. doi:10.1016/S0040-
4039(00)91213-9.
(21) Narula, A. S. (1981). "An analysis of the diastereomeric transition state interactions for
the kinetic deprotonation of acyclic carbonyl derivatives with lithium
diisopropylamide". Tetrahedron Letters. 22(41): 4119±4122. doi:10.1016/S0040-
4039(01)82081-5.
(22) Ireland, RE; Wipf, P; Armstrong, JD (1991). "Stereochemical control in the ester
enolate Claisen rearrangement. 1. Stereoselectivity in silyl ketene acetal
formation". Journal of Organic Chemistry. 56 (2): 650±657. doi:10.1021/jo00002a030.
ϳϱ
(23) Xie, L; Isenberger, KM; Held, G; Dahl, LM (October 1997). "Highly Stereoselective
Kinetic Enolate Formation: Steric vs Electronic Effects". Journal of Organic
Chemistry. 62 (21): 7516±7519. doi:10.1021/jo971260a. PMID 11671880.
(24) Directed Aldol Synthesis ± Formation of E-enolate and Z-enolate, Cowden, C. J.;
Paterson,
I. Org. React. 1997, 51, 1.
(25) Cowden, C. J.; Paterson, I. (2004). Asymmetric Aldol Reactions Using Boron
Enolates. Organic Reactions. pp. 1±
200. doi:10.1002/0471264180.or051.01. ISBN 978-0471264187.
(26) Evans, D. A.; Nelson J. V.; Vogel E.; Taber T. R. (1981). "Stereoselective aldol
condensations via boron enolates". Journal of the American Chemical
Society. 103 (11): 3099±3111. doi:10.1021/ja00401a031.
(27) Evans, D. A.; Rieger D. L.; Bilodeau M. T.; Urpi F. (1991). "Stereoselective aldol
reactions of chlorotitanium enolates. An efficient method for the assemblage of
polypropionate-related synthons". Journal of the American Chemical Society. 113 (3):
1047±1049. doi:10.1021/ja00003a051.
(28) Evans D. A. et al. Top. Stereochem. 1982, 13, 1±115. (Review)
(29) Roush W. R. (1991). "Concerning the diastereofacial selectivity of the aldol reactions
of .alpha.-methyl chiral aldehydes and lithium and boron propionate enolates". Journal
of Organic Chemistry. 56 (13): 4151±4157. doi:10.1021/jo00013a015.
(30) Masamune S.; Ellingboe J. W.; Choy W. (1982). "Aldol strategy: coordination of the
lithium cation with an alkoxy substituent". Journal of the American Chemical
Society. 104 (20): 1047±1049. doi:10.1021/ja00384a062.
(31) Jump up to:a b Evans, D. A.; Dart M. J.; Duffy J. L.; Rieger D. L. (1995). "Double
Stereodifferentiating Aldol Reactions. The Documentation of "Partially Matched"
Aldol Bond Constructions in the Assemblage of Polypropionate Systems". Journal of
the American Chemical Society. 117 (35): 9073±9074. doi:10.1021/ja00140a027.
(32) Masamune S.; Choy W.; Petersen J. S.; Sita L. R. (1985). "Double Asymmetric
Synthesis and a New Strategy for Stereochemical Control in Organic
Synthesis". Angew. Chem. Int. Ed. Engl. 24: 1±30. doi:10.1002/anie.198500013.
(33) Evans D. A. Aldrichimica Acta 1982, 15, 23. (Review)
(34) Gage J. R.; Evans D. A., Diastereoselective Aldol Condensation Using A Chiral
Oxazolidinone Auxiliary: (2S*,3S*)-3-Hydroxy-3-Phenyl-2-Methylpropanoic
Acid, Organic Syntheses, Coll. Vol. 8, p.339 (1993); Vol. 68, p.83 (1990).
ϳϲ
(35) Evans, D. A.; Bartroli J.; Shih T. L. (1981). "Enantioselective aldol condensations. 2.
Erythro-selective chiral aldol condensations via boron enolates". Journal of the
American Chemical Society. 103 (8): 2127±2129. doi:10.1021/ja00398a058.
(36) Jump up to:a b Evans, D. A.; Bender S. L.; Morris J. (1988). "The total synthesis of the
polyether antibiotic X-206". Journal of the American Chemical Society. 110 (8): 2506±
2526. doi:10.1021/ja00216a026.
(37) Evans, D. A.; Clark J.S.; Metternich R.; Sheppard G.S. (1990). "Diastereoselective
aldol reactions using .beta.-keto imide derived enolates. A versatile approach to the
assemblage of polypropionate systems". Journal of the American Chemical
Society. 112 (2): 866±868. doi:10.1021/ja00158a056.
(38) Evans, D. A.; Ng, H.P.; Clark, J.S.; Rieger, D.L. (1992). "Diastereoselective anti aldol
reactions of chiral ethyl ketones. Enantioselective processes for the synthesis of
polypropionate natural products". Tetrahedron. 48 (11): 2127±
2142. doi:10.1016/S0040-4020(01)88879-7.
(39) Shigehisa, H.; Mizutani, T.; Tosaki, S. Y.; Ohshima, T.; Shibasaki, M, Tetrahedron
2005, 61, 5057-5065.
(40) In this reaction the nucleophile is a boron enolate derived from reaction
with dibutylboron triflate (nBu2BOTf), the base is N,N-diisopropylethylamine. The
thioether is removed in step 2 by Raney Nickel / hydrogen reduction.
(41) S. B. Jennifer Kan; Kenneth K.-H. Ng; Ian Paterson (2013). "The Impact of the
Mukaiyama Aldol Reaction in Total Synthesis". Angewandte Chemie International
Edition. 52 (35): 9097±9108. doi:10.1002/anie.201303914. PMID 23893491.
(42) Teruaki Mukaiyama; Kazuo Banno; Koichi Narasaka (1974). "Reactions of silyl enol
ethers with carbonyl compounds activated by titanium tetrachloride". Journal of the
American Chemical Society. 96 (24): 7503±7509. doi:10.1021/ja00831a019.
(43) 3-Hydroxy-3-Methyl-1-Phenyl-1-Butanone by Crossed Aldol Reaction Teruaki
Mukaiyama and Koichi Narasaka Organic Syntheses, Coll. Vol. 8, p.323 (1993); Vol.
65, p.6 (1987)
(44) Carreira E.M.; Singer R.A.; Lee W.S. (1994). "Catalytic, enantioselective aldol
additions with methyl and ethyl acetate O-silyl enolates ² a chira; tridentate chelate as
a ligand for titanium(IV)"(PDF). Journal of the American Chemical Society. 116 (19):
8837±8. doi:10.1021/ja00098a065.
(45) Kruger J.; Carreira E.M. (1998). "Apparent catalytic generation of chiral metal
enolates: Enantioselective dienolate additions to aldehydes mediated by Tol-BINAP
ϳϳ
center Cu(II) fluoride complexes". Journal of the American Chemical Society. 120 (4):
837±8. doi:10.1021/ja973331t.
(46) Pagenkopf B.L.; Kruger J.; Stojanovic A.; Carreira E.M. (1998). "Mechanistic insights
into Cu-catalyzed asymmetric aldol reactions: Chemical and spectroscopic evidence for
a metalloenolate intermediate". Angew. Chem. Int. Ed. 37 (22): 3124±
6. doi:10.1002/(SICI)1521-3773(19981204)37:22<3124::AID-ANIE3124>3.0.CO;2-
1.
(47) Crimmins M. T.; King B. W.; Tabet A. E. (1997). "Asymmetric Aldol Additions with
Titanium Enolates of Acyloxazolidinethiones: Dependence of Selectivity on Amine
Base and Lewis Acid Stoichiometry". Journal of the American Chemical
Society. 119 (33): 7883±7884. doi:10.1021/ja9716721.
(48) Crimmins M. T.; Chaudhary K. (2000). "Titanium enolates of thiazolidinethione chiral
auxiliaries: Versatile tools for asymmetric aldol additions". Organic Letters. 2 (6): 775±
777. doi:10.1021/ol9913901. PMID 10754681.
(49) Carreira, E. M.; Fettes, A.; Martl, C. (2006). Catalytic Enantioselective Aldol Addition
Reactions. Org. React. 67. pp. 1±216. doi:10.1002/0471264180.or067.01. ISBN 978-
0471264187.
(50) Z. G. Hajos, D. R. Parrish, German Patent DE 2102623 1971
(51) Hajos, Zoltan G.; Parrish, David R. (1974). "Asymmetric synthesis of bicyclic
intermediates of natural product chemistry". Journal of Organic Chemistry. 39 (12):
1615±1621. doi:10.1021/jo00925a003.
(52) Eder, Ulrich; Sauer, Gerhard; Wiechert, Rudolf (1971). "New Type of Asymmetric
Cyclization to Optically Active Steroid CD Partial Structures". Angewandte Chemie
International Edition in English. 10(7): 1615±1621. doi:10.1002/anie.197104961.
(53) List, Benjamin (2006). "The ying and yang of asymmetric aminocatalysis". Chemical
Communications (8): 819±824. doi:10.1039/b514296m. PMID 16479280.
(54) Northrup, Alan B.; MacMillan David W. C. (2002). "The First Direct and
Enantioselective Cross-Aldol Reaction of Aldehydes" (PDF). Journal of the American
Chemical Society. 124 (24): 6798±6799. doi:10.1021/ja0262378. PMID 12059180.
(55) Northrup A. B.; Mangion I. K.; Hettche F.; MacMillan D. W. C. (2004).
"Enantioselective Organocatalytic Direct Aldol Reactions of -Oxyaldehydes: Step One
in a Two-Step Synthesis of Carbohydrates". Angewandte Chemie International Edition
in English. 43 (16): 2152±2154. doi:10.1002/anie.200453716. PMID 15083470.
ϳϴ
(56) Evans, D. A.; Tedrow, J. S.; Shaw, J. T.; Downey, C. W. (2002). "Diastereoselective
Magnesium Halide-Catalyzed anti-Aldol Reactions of Chiral N-
Acyloxazolidinones". Journal of the American Chemical Society. 124 (3): 392±
393. doi:10.1021/ja0119548. PMID 11792206.
(57) Evans, David A.; Downey, C. Wade; Shaw, Jared T.; Tedrow, Jason S. (2002).
"Magnesium Halide-Catalyzed Anti-Aldol Reactions of Chiral N-
Acylthiazolidinethiones". Organic Letters. 4 (7): 1127±
1130. doi:10.1021/ol025553o. PMID 11922799.
(58) Magdziak, D.; Lalic, G.; Lee, H. M.; Fortner, K. C.; Aloise, A. D.; Shair, M. D. (2005).
"Catalytic Enantioselective Thioester Aldol Reactions That Are Compatible with Protic
Functional Groups". Journal of the American Chemical Society. 127 (20): 7284±
7285. doi:10.1021/ja051759j. PMID 15898756.
ϳϵ
Buy your books fast and straightforward online - at one of world’s
fastest growing online book stores! Environmentally sound due to
Print-on-Demand technologies.
Buy your books online at
www.morebooks.shop
Kaufen Sie Ihre Bücher schnell und unkompliziert online – auf einer
der am schnellsten wachsenden Buchhandelsplattformen weltweit!
Dank Print-On-Demand umwelt- und ressourcenschonend produzi
ert.
Bücher schneller online kaufen
www.morebooks.shop
KS OmniScriptum Publishing
Brivibas gatve 197
LV-1039 Riga, Latvia info@omniscriptum.com
Telefax: +371 686 204 55 www.omniscriptum.com
View publication stats