Вы находитесь на странице: 1из 40

http://www.internetchemistry.com/chemistry/rearrangement-reactions.

htm
Rearrangement reactions
In chemistry a rearrangement is a chemical reaction in which the carbon skeleton of a molecule
is rearranged to give a structural isomer of the original molecule.
Online available information resources about rearrangement reactions and mechanisms in
chemistry.
Further information categories about related topics are listed in the navigation menu on the left
side of these page.
Anionic
Rearrangements Induced by Bases or Electron Rich Sites. Virtual Textbook of Organic Chemistry - [e]

Rearrangements

Cationic
Rearrangements
Rearrangements Induced by Cationic or Electron Deficient Sites. Virtual Textbook of Organic Chemistry - [e]

[Lec] [Part] [Special]


|

Partial Information

Amadori
Animated slide show. Pudue University, USA - Format: PPT - [e]

rearrangement

Amadori
rearrangement
Development of a new active based on Amadori rearrangement taking place in protein based tissues. University of Mnster,
Germany - Format: PDF - [e]
Baeyer-Villiger
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Baker-Venkataraman
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Beckmann
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Benzilic
Acid
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Brook
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Claisen
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Claisen-Eschenmoser
Rearrangement
A Practical Variant of the Claisen-Eschenmoser Rearrangement: Synthesis of Unsaturated Morpholine Amides. Berkeley
University, USA - Format: PDF - [e]
Cope
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Curtius
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Dimroth
Rearrangement
Ring-fission of a heterocyclic imine with subsequent recyclization to a more stable amino entity - Format: PDF - [e]
Favorskii
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Reaction

Fries
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Heyns
Rearrangement
Development of Polymer-supported synthetic procedure for Heyns Rearrangement Products. Dissertation, 1999 - Format: PDF [e]
Ireland-Claisen
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Lobry-de
Bruyn-van
Description of the reaction and detail at the mechanism. Wiki - [e]

transformation

Ekenstein

Newman-Kwart
Rearrangement
In the NKR a intramolecular aryl migration of O-thiocarbamates at high temperatures leads to S-thiocarbamates. Organic
Chemistry Portal - [d, e]
Overman
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Pinacol
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Pummerer
The Sulfinate-Sulfone Pummerer Rearrangement - Format: PDF - [e]

Rearrangement

Schmidt
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Stevens
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Wittig
[1,2]- and [2,3]-Wittig Rearrangement - Format: PDF - [e]

Rearrangement

Wittig:
[1,2]-Wittig
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Wittig:
[2,3]-Wittig
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

Wolff
Description of the reaction and detail at the mechanism. Organic Chemistry Portal - [d, e]

Rearrangement

[Lec] [Part] [Special]

A rearrangement reaction is a broad class of organic reactions where the carbon


skeleton of a molecule is rearranged to give a structural isomer of the original
molecule.[1] Often a substituent moves from one atom to another atom in the same
molecule. In the example below the substituent R moves from carbon atom 1 to carbon
atom 2:

Intermolecular rearrangements also take place.


A rearrangement is not well represented by simple and discrete electron transfers
(represented by curly arrows in organic chemistry texts). The actual mechanism of alkyl
groups moving, as in Wagner-Meerwein rearrangement, probably involves transfer of
the moving alkyl group fluidly along a bond, not ionic bond-breaking and forming. In
pericyclic reactions, explanation by orbital interactions give a better picture than simple
discrete electron transfers. It is, nevertheless, possible to draw the curved arrows for a
sequence of discrete electron transfers that give the same result as a rearrangement
reaction, although these are not necessarily realistic. In allylic rearrangement, the
reaction is indeed ionic.
Three key rearrangement reactions are 1,2-rearrangements, pericyclic
reactions and olefin metathesis.
Contents
[hide]

1 1,2-rearrangements
2 Pericyclic reactions
3 Olefin metathesis
4 See also
5 References

1,2-rearrangements[edit]
Main article: 1,2-rearrangement
A 1,2-rearrangement is an organic reaction where a substituent moves from one atom
to another atom in a chemical compound. In a 1,2 shift the movement involves two

adjacent atoms but moves over larger distances are possible. Examples are
the Wagner-Meerwein rearrangement:

and the Beckmann rearrangement:

Pericyclic reactions[edit]
Main article: pericyclic reactions
A pericyclic reaction is a type of reaction with multiple carbon-carbon bond making and
breaking wherein the transition state of the molecule has a cyclic geometry, and the
reaction progresses in a concerted fashion. Examples are hydride shifts

and the Claisen rearrangement:

Olefin metathesis[edit]
Main article: Olefin metathesis
Olefin metathesis is a formal exchange of the alkylidene fragments in two alkenes. It is
a catalytic reaction with carbene, or more accurately, transition metal carbene
complex intermediates.

In this example, a vinyl compound is dimerized with the expulsion of ethene.

See also[edit]

Beckmann rearrangement

Curtius rearrangement

Hofmann rearrangement

Lossen rearrangement

Schmidt reaction

Tiemann rearrangement

Wolff rearrangement

Photochemical rearrangements

Introduction to Rearrangement Reactions


by J AMES
in ORGANI C CH EMI STR Y 1 , ORGANI C R EACTI O NS

The previous four posts on acidbase, substitution, addition, and elimination covered the 4
main reactions in organic chemistry I. Now its time to go
beyond those mainstays to introduce a few of the less
common (but still important) reactions you learn in organic
chemistry 1. They will be rearrangements, radical
substitution, and cleavage (oxidative cleavage).
Lets look at rearrangements in this post. As with
everything in this series, the point is not to
understand why just yet, but to be able to see from the
diagrams what bonds are broken and formed. You need to
understand how to read line diagrams. But other than that
no further skills are required. The point here is to be able
to follow the plot to see what is happening. A later
series of posts will go into more detail as to why things
happen, but it takes time to build up that knowledge.
Rearrangement reactions are really interesting. They can
accompany many of the reactions weve previously
covered such as substitution, addition, and elimination
reactions. In fact, if you dont look closely, sometimes you
can miss the fact that a rearrangement reaction has
occurred. Lets look at a substitution reaction first.

On the top is a typical substitution reaction: were taking


an alkyl halide and adding water. The C-Br bond is broken
and a C-OH bond is formed. If you look at the table on the
right youll see this follows the typical pattern of
substitution reactions.
However if we change one thing about this alkyl halide
move the bromine to C-3 instead of C-2 now when we
run this reaction we see a different product emerge. It is
also a substitution reaction (were replacing Br with OH)
but its on a different carbon. Thats because if you look
closely, you can see there are actually 3 bonds broken
and 3 bonds formed. The C2-H bond broke and the C3-H
bond formed.
Very strange!

This represents a rearrangement reaction when you see


a group move from one carbon to another. Lets look at
another example.
Here we have an addition reaction. On top, nothing special
as with all additions, we break a C-C double bond
( bond )and form two new single bonds to the adjoining
carbons (H and Cl). But look at the bottom example. If we
use that alkene instead, we find that the Cl ends up on
C3, notC-2. Again, examining the bonds broken/formed,
we see that theres an extra pair of events: the C3-H bond
was broken and the C-2H bond was formed. In other
words, the hydrogen migrated from one carbon to
another. Weird!

Finally, lets look at an elimination reaction. If you take an


alcohol like the one below and add an acid (like H2SO4,

pictured) and help the reaction along with some heat, you
break the C1-OH and C2-H bonds, and form a new double
bond between C1-C2. This is, in other words, a typical
elimination reaction.
But if you take a slightly modified alcohol like the bottom
example (with an extra methyl group on C1) and try the
same reaction, something strange happens again.
Analyzing the bonds broken and formed, theres an
extra bond being broken and an extra bond being
formed here. If you look closely you can see that one of
the methyl groups on C1 (well call it C8) moved over to
C2.

So what can we conclude about rearrangement reactions?

1. they can accompany many of the reactions were


already familiar with, such as substitution, addition, and
elimination reactions.
2. They involve the movement of an atom (H in the top
two examples, C in the third) from one carbon to another.

What other insight might we glean from these examples?


Heres two questions.
1. look up, if you dont know already, what primary,
secondary and tertiary alcohols and alkyl halides are.

2. In the substitution reactions and the elimination


reactions, classify every alcohol (or alkyl halide) according
to whether it is primary, secondary or tertiary. Notice any
difference between the normal cases and the
rearrangement cases?
For more detail on rearrangement reactions, start
here: Rearrangement Reactions (1) Hydride Shifts
Next Post: Introduction To Free-Radical Substitution
Reactions

Carbocation rearrangements are extremely common in organic chemistry reactions are are defined
as the movement of a carbocation from an unstable state to a more stable state through the use of
various structural reorganizational "shifts" within the molecule. Once the carbocation has shifted over
to a different carbon, we can say that there is a structural isomer of the initial molecule. However,
this phenomenon is not as simple as it sounds.

Introduction
Whenever alcohols are subject to transformation into various carbocations, the carbocations are
subject to a phenomenon known as carbocation rearrangement. A carbocation, in brief, holds the
positive charge in the molecule that is attached to three other groups and bears a sextet rather than
an octet. However, we do see carbocation rearrangements in reactions that do not contain alcohol
as well. Those, on the other hand, require more difficult explanations than the two listed below.
There are two types of rearrangements: hydride shift and alkyl shift. These rearrangements usualy
occur in many types of carbocations. Once rearranged, the molecules can also undergo
further unimolecular substitution (SN1) or unimolecular elimination (E1). Though, most of the time we
see either a simple or complex mixture of products. We can expect two products before undergoing
carbocation rearrangement, but once undergoing this phenomenon, we see the major product.

Hydride Shift
Whenever a nucleophile attacks some molecules, we typically see two products. However, in most
cases, we normally see both a major product and a minor product. The major product is typically the
rearranged product that is more substituted (aka more stable). The minor product, in contract, is
typically the normal product that is less substituted (aka less stable).

The reaction: We see that the formed carbocations can undergo rearrangements called hydride
shift. This means that the two electron hydrogen from the unimolecular substitution moves over to
the neighboring carbon. We see the phenomenon of hydride shift typically with the reaction of an
alcohol and hydrogen halides, which include HBr, HCl, and HI. HF is typically not used because of
its instability and its fast reactivity rate. Below is an example of a reaction between an alcohol and
hydrogen chloride:

GREEN (Cl) = nucleophile

BLUE (OH) = leaving group ORANGE (H) = hydride shift


proton RED(H) = remaining proton

The alcohol portion (-OH) has been substituted with the nucleophilic Cl atom. However, it is not a
direct substitution of the OH atom as seen in SN2 reactions. In this SN1 reaction, we see that the
leaving group, -OH, forms a carbocation on Carbon #3 after receiving a proton from the nucleophile
to produce an alkyloxonium ion. Before the Cl atom attacks, the hydrogen atom attached to the
Carbon atom directly adjacent to the original Carbon (preferably the more stable Carbon), Carbon
#2, can undergo hydride shift. The hydrogen and the carbocation formally switch positions. The Cl
atom can now attack the carbocation, in which it forms the more stable structure because of
hyperconjugation. The carbocation, in this case, is most stable because it attaches to the tertiary
carbon (being attached to 3 different carbons). However, we can still see small amounts of the

minor, unstable product. The mechanism for hydride shift occurs in multiple stepsthat includes
various intermediates and transition states. Below is the mechanism for the given reaction above:

Hydration of Alkenes: Hydride Shift


In a more complex case, when alkenes undergo hydration, we also observe hydride shift. Below is
the reaction of 3-methyl-1-butene with H3O+ that furnishes to make 2-methyl-2-butanol:

Once again, we see multiple products. In this case, however, we see two minor products and one
major product. We observe the major product because the -OH substitutent is attached to the more
substituted carbon. When the reactant undergoes hydration, the proton attaches to carbon #2. The
carbocation is therefore on carbon #2. Hydride shift now occurs when the hydrogen on the adjacent
carbon formally switch places with the carbocation. The carbocation is now ready to be attacked by
H2O to furnish an alkyloxonium ion because of stability and hyperconjugation. The final step can be
observed by another water molecule attacking the proton on the alkyloxonium ion to furnish an
alcohol. We see this mechanism below:

Alkyl Shift
Not all carbocations have suitable hydrogen atoms (either secondary or tertiary) that are on adjacent
carbon atoms available for rearrangement. In this case, the reaction can undergo a different mode of
rearrangement known as alkyl shift (or alkyl group migration). Alkyl Shift acts very similarily to that
of hydride shift. Instead of the proton (H) that shifts with the nucleophile, we see an alkyl group that
shifts with the nucleophile instead. The shifting group carries its electron pair with it to furnish a bond
to the neighboring or adjacent carbocation. The shifted alkyl group and the positive charge of the
carbocation switch positions on the moleculeReactions of tertiary carbocations react much faster
than that of secondary carbocations. We see alkyl shift from a secondary carbocation to tertiary
carbocation in SN1 reactions:

We observe slight variations and differences between the two reactions. In reaction #1, we see that
we have a secondary substrate. This undergoes alkyl shift because it does not have a suitable
hydrogen on the adjacent carbon. Once again, the reaction is similar to hydride shift. The only
difference is that we shift an alkyl group rather than shift a proton, while still undergoing various
intermediate steps to furnish its final product.
With reaction #2, on the other hand, we can say that it undergoes a concerted mechanism. In short,
this means that everything happens in one step. This is because primary carbocations cannot be an
intermediate and they are relatively difficult processes since they require higher temperatures and
longer reaction times. After protonating the alcohol substrate to form the alkyloxonium ion, the water
must leave at the same time as the alkyl group shifts from the adjacent carbon to skip the formation
of the unstable primary carbocation.

Carbocation Rearrangements for E1


Reactions
E1 reactions are also affected by alkyl shift. Once again, we can see both minor and major products.
However, we see that the more substituted carbons undergo the effects of E1 reactions and furnish
a double bond. See practice problem #4 below for an example as the properties and effects of
carbocation rearrangements in E1 reactions are similar to that of alkyl shifts.

1,3-Hydride and Greater Shifts


Typically, hydride shifts can occur at low temperatures. However, by heating the solutionf of a cation,
it can easily and readily speed the process of rearrangement. One way to account for a slight barrier
is to propose a 1,3-hydride shift interchanging the functionality of two different kinds of methyls.
Another possibility is 1,2 hydride shift in which you could yield a secondary carbocation intermediate.
Then, a further 1,2 hydride shift would give the more stable rearranged tertiary cation.
More distant hydride shifts have been observed, such as 1,4 and 1,5 hydride shifts, but these
arrangements are too fast to undergo secondary cation intermediates.

Analogy
Carbocation rearrangements happen very readily and often occur in many organic chemistry
reactions. Yet, we typically neglect this step. Dr. Sarah Lievens, a Chemistry professor at the
University of California, Davis once said carbocation rearrangements can be observed with various
analogies to help her students remember this phenomenon. For hydride shifts: "The new friend
(nucleophile) just joined a group (the organic molecule). Because he is new, he only made two new
friends. However, the popular kid (the hydrogen) glady gave up his friends to the new friend so that
he could have even more friends. Therefore, everyone won't be as lonely and we can all be friends."
This analogy works for alkyl shifts in conjunction with hydride shift as well.

References
1. Vogel, Pierre. Carbocation Chemistry. Amsterdam: Elsevier Science Publishers B.V., 1985.
2. Olah, George A. and Prakash, G.K. Surya. Carbocation Chemistry. New Jersey: John Wiley & Sons,
Inc., 2004.
3. Vollhardt, K. Peter C. and Schore, Neil E. Organic Chemistry: Structure and Function. New York:
Bleyer, Brennan, 2007.

Outside links

http://en.wikipedia.org/wiki/Carboca..._rearrangement
Watch a short presentation on the carbocation rearrangement phenomenon

Problems

Answers to Practice Problems

Wagner-Meerwein
Rearrangement
A Wagner-Meerwein rearrangement is any reaction in which the carbon skeleton of a reactant
changes due to one or more rearrangements involving carbocations, e.g.:

mechanism:

Further Information
Literature
Related Reactions
Schmidt Reaction

Curtius Rearrangement

The Curtius Rearrangement is the thermal decomposition of carboxylic azides to produce


an isocyanate. These intermediates may be isolated, or their corresponding reaction or
hydrolysis products may be obtained.
The reaction sequence - including subsequent reaction with water which leads to amines is named the Curtius Reaction. This reaction is similar to the Schmidt Reaction with acids,
differing in that the acyl azide in the present case is prepared from the acyl halide and an
azide salt.

Mechanism of the Curtius Rearrangement

Preparation of azides:

Decomposition:

Reaction with water to the unstable carbamic acid derivative which will undergo
spontaneous decarboxylation:

Isocyanates are versatile starting materials:

Isocyanates are also of high interest as monomers for polymerization work and in the
derivatisation of biomacromolecules.
Recent Literature

Propylphosphonic Anhydride (T3P)-Mediated One-Pot Rearrangement of Carboxylic


Acids to Carbamates
J. K. Augustine, A. Bombrun, A. B. Mandal, P. Alagarsamy, R. N. Atta, P.
Selvam, Synthesis, 2011, 1477-1483.

Boc-Protected Amines via a Mild and Efficient One-Pot Curtius Rearrangement


H. Lebel, O. Leogane, Org. Lett., 2005, 7, 4107-4110.

One-Pot Synthesis of Ureido Peptides and Urea-Tethered Glycosylated Amino Acids


Employing Deoxo-Fluor and TMSN3
H. P. Hemantha, G. Chennakrishnareddy, T. M. Vishwanatha, V. V.
Sureshbabu, Synlett, 2009, 407-410.

Radical Azidonation of Aldehydes


L. Marinescu, J. Thinggaard, I. B. Thomsen, M. Bols, J. Org. Chem., 2003, 68, 9453-9455.

Iodobenzene Dichloride in Combination with Sodium Azide for the Effective Synthesis of
Carbamoyl Azides from Aldehydes
X.-Q. Li, X.-F. Zhao, C. Zhang, Synthesis, 2008, 2589-2593.

Radical Azidonation of Aldehydes


L. Marinescu, J. Thinggaard, I. B. Thomsen, M. Bols, J. Org. Chem., 2003, 68, 9453-9455.

An Efficient Synthesis of a Probe for Protein Function: 2,3-Diaminopropionic Acid with


Orthogonal Protecting Groups
E. A. Englund, H. N. Gopi, D. H. Appella, Org. Lett., 2004, 6, 213-215.

Further Information
Literature
Related Reactions
Curtius Rearrangement

Schmidt Reaction

The acid-catalysed reaction of hydrogen azide with electrophiles, such as carbonyl


compounds, tertiary alcohols or alkenes. After a rearrangement and extrusion of N2,
amines, nitriles, amides or imines are produced.

Mechanism of the Schmidt Reaction

Reaction of carboxylic acids gives acyl azides, which rearrange to isocyanates, and these
may be hydrolyzed to carbamic acid or solvolysed to carbamates. Decarboxylation leads to
amines.

The reaction with a ketone gives an azidohydrin intermediate, which rearranges to form an
amide:

Alkenes are able to undergo addition of HN3 as with any HX reagent, and the resulting alkyl
azide can rearrange to form an imine:

Tertiary alcohols give substitution by azide via a carbenium ion, and the resulting alkyl azide
can rearrange to form an imine.
Recent Literature

Chemoselective Schmidt Reaction Mediated by Triflic Acid: Selective Synthesis of Nitriles

from Aldehydes
B. V. Rokade, J. R. Prabhu, J. Org. Chem., 2012, 77, 5364-5370.

Efficient, One-Pot, BF3OEt2-Mediated Synthesis of Substituted N-Aryl Lactams


D. Caturvedi, A. K. Chaturvedi, N. Mishra, V. Mishra, Synlett, 2012, 23, 2627-2630.

Gold(I)-Catalyzed Intramolecular Acetylenic Schmidt Reaction


D. J. Gorin, N. R. Davis, F. D. Toste, J. Am. Chem. Soc., 2005, 127, 11260-11261.

Migratory aptitude is the relative ability of a migrating group to migrate in


a rearrangement reaction. This can be affected by the leaving group (whichever gives a
more stable carbocation)depends upon the electron density of the migrating group i.e.
Hydride > Phenyl = Tert-Alkyl > Sec-Alkyl > Primary Alkyl > Methyl
The migratory groups follow the order of increasing stability of carbocation instead of
carbanion.

1. Baeyer-Villiger Migratory Aptitude


Baeyer-Villiger oxidation of a ketone to an ester by a peroxy acid involves initial
formation of a sort of "hemiketal" between the reagents which then undergoes acidcatalyzed rearrangement by the mechanism shown by curved arrows in the following
scheme:

The key element of this rearrangment is migration of substituent group (A or B) of the


ketone with its electrons to attack the nearer peroxy oxygen atom and displace the
carboxylic acid. We formulate the rearrangement as, for example, a methide or a
hydride shift. [Note that protonation helped make the carboxylic acid a better leaving
group.]
An obvious question is "Which ester is formed?" that is "Which group migrates
more easily?"

A series of investigations in which one determines the ratio in which esters are formed
from a variety of different ketones allows one to establish a scale of "migratory
aptitudes." As the textbook states on page 412, the approximate order of decreasing
ease of migration is hydrogen > tertiary alkyl > secondary alkyl > phenyl > primary
alkyl > methyl.
At first glance this order seems crazy. One might think that if an anion is migrating,
the more stable anion should migrate more easily. What see instead that the more
stable cation migrates more easily.
The lesson we learn is that, although we may draw the curved arrow showing an anion
migration, the anion never breaks free of the rest of the molecule. What is really
important is the availability of the bonding electrons the the C-A (or C-B) bond to mix
with the sigma* orbital of the O-O bond and displace the carboxylic acid from the
nearer oxygen. The group that would form a more stable cation holds the bonding
electrons less tightly and makes them more available (higher HOMO) for attacking
the O-O bond.
Thus, paradoxically, the better cation is the better "anionic" migrating group. (Of
course a free anion is never truly involved and in the transition state for the migration
the migrating group looks more like a cation, surrendering electron density to the O-O
group, than like an anion.)

Migratory aptitude in pinacol-pinacolone


rearrangement

I am confused about the migratory aptitude of various groups, as there are many different
orders for the same given in different places, especially about -Ph and -H. I would like to know if
someone could reliably tell the order.

According to my FIITJEE textbook, the relative order of migratory aptitude of groups in


pinacol-pinacolone rearrangement is:

H>Ph>Me3C>MeCH2>Me
There does not seem to be a direct correlation between the mass and the migratory
aptitude of these groups.

In the pinacol rearrangement, a 1,2-diol is treated with acid and rearranges to a


carbonyl compound. Here is a reaction scheme showing a mechanism for the
rearrangement.

In this case the molecule is symmetric and methyl migration is the only reaction
pathway available. Methyl migration can 1) help stabilize the developing carbocation
center and 2) once fully migrated, produces a resonance stabilized hydroxylcarbocation.
If one or more of the methyl groups is replaced with some other substituent (for
example, alkyl, hydrogen, aryl), then our reaction possibilities become more complex;
which hydroxyl group will protonate, which group will migrate? Generally, both hydroxyls
will protonate and lead to separate products. In terms of which group will migrate, the

only thing agreed upon is that a tertiary carbon will migrate in preference to a secondary
carbon which will migrate in preference to a primary carbon.Clearly, the ability of the
migrating group to stabilize positive charge plays a role.
The reaction is reversible and product distribution can change as reaction conditions are
changed (thermodynamic control vs. kinetic control). All of these complexities make it
difficult to sort things out in detail, thereby making it difficult to determine the true
relative migratory preference of other groups like H and phenyl. This is why the
migratory aptitudes of these groups (H and phenyl) can change from textbook to
textbook.
See section 2. Pinacol Rearrangement in this reference for a very nice, straightforward
discussion of this reaction mechanism, and all its subtleties.

The pinacol rearrangement or pinacolpinacolone rearrangement is a method for


converting a 1,2-diol to a carbonyl compound in organic chemistry. This 1,2rearrangement takes place under acidic conditions. The name of the reaction comes
from the rearrangement of pinacol to pinacolone.

This reaction was first described by Wilhelm Rudolph Fittig in 1860.[1]


Contents
[hide]

1 An overview of mechanism(discussion)
2 Stereochemistry of the rearrangement
3 History
4 See also
5 References

An overview of mechanism(discussion)[edit]
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 >
tertiary carbocation (if formed by migration) > secondary carbocation (if formed by

migration) > methyl cation . The conclusion which group stabilizes carbocation more
effectively is migrated

Stereochemistry of the rearrangement[edit]


In cyclic systems, the reaction presents more features of interest. In these reactions,
the stereochemistry of the diol plays a crucial role in deciding the major product. An
alkyl group which is situated trans- to the leaving OH group alone may migrate. If
otherwise, ring expansion occurs, i.e. the ring carbon itself migrates to the
carbocation centre. This reveals another interesting feature of the reaction, viz. that
it is largely concerted. There appears to be a connection between the migration
origin and migration terminus throughout the reaction.
Moreover, if the migrating alkyl group has a chiral center as its key atom, the
configuration at this center is retained even after migration takes place.

History[edit]
Although Fittig first published about the pinacol rearrangement,it was not Fittig
but Aleksandr Butlerov who correctly identified the reaction products involved.[2]
In an 1859 publication Wilhelm Rudolph Fittig described the reaction
of acetone with potassium metal...[3] Fittig wrongly assumed a molecular formula of
(C3H3O)nfor acetone, the result of a long standing atomic weight debate finally settled
at the Karlsruhe Congress in 1860. He also wrongly believed acetone to be an
alcohol which he hoped to prove by forming a metal alkoxide salt. The reaction
product he obtained instead he called paraceton which he believed to be an
acetone dimer. In his second publication in 1860 he reacted paraceton with sulfuric
acid (the actual pinacol rearrangement).

Again Fittig was unable to assign a molecular structure to the reaction product
which he assumed to be another isomer or a polymer. Contemporary chemists

who had already adapted to the new atomic weight reality did not fare better.
One of them, Charles Friedel, believed the reaction product to be
the epoxidetetramethylethylene oxide[4] in analogy with reactions of ethylene
glycol. Finally Butlerov in 1873 came up with the correct structures after he
independently synthesised the compound trimethylacetic acid which Friedel had
obtained earlier by oxidizing with a dichromate.[5]
Some of the problems during the determination of the structure are because
carbon skeletal rearrangements were unknown at that time and therefore the
new concept had to be found. Butlerov theory allowed the structure of carbon
atoms in the molecule to rearrange and with this concept a structure for
pinacolone could be found.

See also[edit]

semipinacol rearrangement

Tiffeneau-Demjanov rearrangement, in which the leaving group is a diazo


(from amine) rather than oxonium (from hydroxyl)

References[edit]
1. Jump up^ Wilhelm Rudolph Fittig (1860). "ber einige Derivate des
Acetons". Annalen der Chemie und Pharmacie 114 (1): 54
63. doi:10.1002/jlac.18601140107.
2. Jump up^ Jerome A. Berson (2002). "What Is a Discovery? Carbon Skeletal
Rearrangements as Counter-Examples to the Rule of Minimal Structural
Change". Angewandte Chemie International Edition 41 (24): 4655
60. doi:10.1002/anie.200290007. PMID 12481317.
3. Jump up^ W. R. Fittig (1859). "Ueber einige Metamorphosen des Acetons
der Essigsure". Annalen der Chemie und Pharmacie 110 (1): 23
45. doi:10.1002/jlac.18591100104.
4. Jump up^ Charles Friedel (1869). "Recherches sur les actones et sur les
aldhydes". Annales de chimie et de physique. Srie 4 16: 310.
5. Jump up^ Aleksandr Butlerov (1873). "Ueber Trimethylessigsure". Justus
Liebigs Annalen der Chemie und Pharmacie 170 (12): 151
162. doi:10.1002/jlac.18731700114.

Вам также может понравиться