You are on page 1of 9

SN1 reaction

From Wikipedia, the free encyclopedia


Jump to: navigation, search
For other uses, see SN1 (disambiguation).

The SN1 reaction is a substitution reaction in organic chemistry. "SN" stands for nucleophilic
substitution and the "1" represents the fact that the rate-determining step is unimolecular.[1][2]
Thus, the rate equation is often shown as having first-order dependence on electrophile and zero-
order dependence on nucleophile. This relationship holds for situations where the amount of
nucleophile is much greater than that of the carbocation intermediate. Instead, the rate equation
may be more accurately described using steady-state kinetics. The reaction involves a
carbocation intermediate and is commonly seen in reactions of secondary or tertiary alkyl halides
under strongly basic conditions or, under strongly acidic conditions, with secondary or tertiary
alcohols. With primary alkyl halides, the alternative SN2 reaction occurs. In inorganic chemistry,
the SN1 reaction is often known as the dissociative mechanism. This dissociation pathway is
well-described by the cis effect. A reaction mechanism was first proposed by Christopher Ingold
et al. in 1940.[3] This reaction does not take account much on the strength of the nucleophile
unlike the SN2 mechanism.

Contents
[hide]

 1 Mechanism
 2 Scope
 3 Stereochemistry
 4 Side reactions
 5 Solvent effects
 6 See also
 7 References
 8 Further reading
 9 External links

Mechanism[edit]
An example of a reaction taking place with an SN1 reaction mechanism is the hydrolysis of tert-
butyl bromide with water forming tert-butanol:
This SN1 reaction takes place in three steps:

 Formation of a tert-butyl carbocation by separation of a leaving group (a bromide anion)


from the carbon atom: this step is slow and reversible.[4]

 Nucleophilic attack: the carbocation reacts with the nucleophile. If the nucleophile is a
neutral molecule (i.e. a solvent) a third step is required to complete the reaction. When
the solvent is water, the intermediate is an oxonium ion. This reaction step is fast.

 Deprotonation: Removal of a proton on the protonated nucleophile by water acting as a


base forming the alcohol and a hydronium ion. This reaction step is fast.

Scope[edit]
The SN1 mechanism tends to dominate when the central carbon atom is surrounded by bulky
groups because such groups sterically hinder the SN2 reaction. Additionally, bulky substituents
on the central carbon increase the rate of carbocation formation because of the relief of steric
strain that occurs. The resultant carbocation is also stabilized by both inductive stabilization and
hyperconjugation from attached alkyl groups. The Hammond-Leffler postulate suggests that this
too will increase the rate of carbocation formation. The SN1 mechanism therefore dominates in
reactions at tertiary alkyl centers and is further observed at secondary alkyl centers in the
presence of weak nucleophiles.

An example of a reaction proceeding in a SN1 fashion is the synthesis of 2,5-dichloro-2,5-


dimethylhexane from the corresponding diol with concentrated hydrochloric acid:[5]
As the alpha and beta substitutions increase with respect to leaving groups the reaction is
diverted from SN2 to SN1.

Stereochemistry[edit]
The carbocation intermediate formed in the reaction's rate limiting step is an sp2 hybridized
carbon with trigonal planar molecular geometry. This allows two different avenues for the
nucleophilic attack, one on either side of the planar molecule. If neither avenue is preferentially
favored, these two avenues occur equally, yielding a racemic mix of enantiomers if the reaction
takes place at a stereocenter.[6] This is illustrated below in the SN1 reaction of S-3-chloro-3-
methylhexane with an iodide ion, which yields a racemic mixture of 3-iodo-3-methylhexane:

However, an excess of one stereoisomer can be observed, as the leaving group can remain in
proximity to the carbocation intermediate for a short time and block nucleophilic attack. This
stands in contrast to the SN2 mechanism, which is a stereospecific mechanism where
stereochemistry is always inverted as the nucleophile comes in from the rear side of the leaving
group.

Side reactions[edit]
Two common side reactions are elimination reactions and carbocation rearrangement. If the
reaction is performed under warm or hot conditions (which favor an increase in entropy), E1
elimination is likely to predominate, leading to formation of an alkene. At lower temperatures,
SN1 and E1 reactions are competitive reactions and it becomes difficult to favor one over the
other. Even if the reaction is performed cold, some alkene may be formed. If an attempt is made
to perform an SN1 reaction using a strongly basic nucleophile such as hydroxide or methoxide
ion, the alkene will again be formed, this time via an E2 elimination. This will be especially true
if the reaction is heated. Finally, if the carbocation intermediate can rearrange to a more stable
carbocation, it will give a product derived from the more stable carbocation rather than the
simple substitution product.

Solvent effects[edit]
See also: Solvent effects

Since the SN1 reaction involves formation of an unstable carbocation intermediate in the rate-
determining step, anything that can facilitate this will speed up the reaction. The normal solvents
of choice are both polar (to stabilize ionic intermediates in general) and protic (to solvate the
leaving group in particular). Typical polar protic solvents include water and alcohols, which will
also act as nucleophiles and the process is known as solvolysis.

The Y scale correlates solvolysis reaction rates of any solvent (k) with that of a standard solvent
(80% v/v ethanol/water) (k0) through

with m a reactant constant (m = 1 for tert-butyl chloride) and Y a solvent parameter.[7] For
example 100% ethanol gives Y = −2.3, 50% ethanol in water Y = +1.65 and 15% concentration
Y = +3.2.[8]

SN2 reaction
From Wikipedia, the free encyclopedia
Jump to: navigation, search
Ball-and-stick representation of the SN2 reaction of CH3SH with CH3I

Structure of the SN2 transition state

The SN2 reaction is a type of reaction mechanism that is common in organic chemistry. In this
mechanism, one bond is broken and one bond is formed synchronously, i.e., in one step. SN2 is a
kind of nucleophilic substitution reaction mechanism. Since two reacting species are involved in
the slow (rate determining) step, this leads to the term substitution nucleophilic (bi-molecular) or
SN2, the other major kind is SN1.[1] Many other more specialized mechanisms describe
substitution reactions.
The reaction type is so common that it has other names, e.g. "bimolecular nucleophilic
substitution", or, among inorganic chemists, "associative substitution" or "interchange
mechanism".

Contents
[hide]

 1 Reaction mechanism
 2 Factors affecting the rate of the reaction
o 2.1 Partial positive charge on carbon
o 2.2 Substrate
o 2.3 Nucleophile
o 2.4 Solvent
o 2.5 Leaving group
 3 Reaction kinetics
 4 E2 competition
 5 Roundabout mechanism
 6 See also
 7 References

Reaction mechanism[edit]
The reaction most often occurs at an aliphatic sp3 carbon center with an electronegative, stable
leaving group attached to it (often denoted X), which is frequently a halide atom. The breaking
of the C–X bond and the formation of the new bond (often denoted C–Y or C–Nu) occur
simultaneously through a transition state in which the carbon under nucleophilic attack is
pentacoordinate, and approximately sp2 hybridised. The nucleophile attacks the carbon at 180° to
the leaving group, since this provides the best overlap between the nucleophile's lone pair and
the C–X σ* antibonding orbital. The leaving group is then pushed off the opposite side and the
product is formed.

If the substrate under nucleophilic attack is chiral, this will lead, to an inversion of
stereochemistry called a Walden inversion (the nucleophile attacks the electrophilic carbon
center, inverting the tetrahedron, much like an umbrella turning inside out in the wind).

In an example of the SN2 reaction, the attack of Br− (the nucleophile) on an ethyl chloride (the
electrophile) results in ethyl bromide, with chloride ejected as the leaving group.:
SN2 reaction of chloroethane with bromide ion

SN2 attack occurs if the backside route of attack is not sterically hindered by substituents on the
substrate. Therefore this mechanism usually occurs at an unhindered primary carbon centre. If
there is steric crowding on the substrate near the leaving group, such as at a tertiary carbon
centre, the substitution will involve an SN1 rather than an SN2 mechanism, (an SN1 would also be
more likely in this case because a sufficiently stable carbocation intermediary could be formed).

Factors affecting the rate of the reaction[edit]


Four factors affect the rate of the reaction:[2]

Partial positive charge on carbon[edit]

More the partial positive charge on carbon more will be reactivity towards SN2

Substrate[edit]

The substrate plays the most important part in determining the rate of the reaction. This is
because the nucleophile attacks from the back of the substrate, thus breaking the carbon-leaving
group bond and forming the carbon-nucleophile bond. Therefore, to maximise the rate of the SN2
reaction, the back of the substrate must be as unhindered as possible. Overall, this means that
methyl and primary substrates react the fastest, followed by secondary substrates. Tertiary
substrates do not participate in SN2 reactions, because of steric hindrance. Moreover, compounds
like '1-chloro 1-ethene' too do not undergo nucleophillic substitution easily because the carbon to
chlorine bond is said to be of partial double bond character, thus is harder to break. Another
factor leading to an SN2 reaction due to substrate involves the stability and ease by which the
carbocation is formed after removing the leaving group. This means the more stable the
carbocation is after removing the leaving group, the more likely it is an SN1 reaction will occur
instead of an SN2. Among the stabilization methods to be considered are: resonance stabilization,
hyper-conjugative stabilization, inductive effect stabilization, or the formation of an aromatic
ring molecule (as in the case of 7-chloro cyclohept-1, 3, 5-triene, as it will form a tropolium
carbocation which is aromatic).

Nucleophile[edit]

Like the substrate, steric hindrance affects the nucleophile's strength. The methoxide anion, for
example, is both a strong base and nucleophile because it is a methyl nucleophile, and is thus
very much unhindered. tert-Butoxide, on the other hand, is a strong base, but a poor nucleophile,
because of its three methyl groups hindering its approach to the carbon. Nucleophile strength is
also affected by charge and electronegativity: nucleophilicity increases with increasing negative
charge and decreasing electronegativity. For example, OH- is a better nucleophile than water, and
I- is a better nucleophile than Br- (in polar protic solvents). In a polar aprotic solvent,
nucleophilicity increases up a column of the periodic table as there is no hydrogen bonding
between the solvent and nucleophile; in this case nucleophilicity mirrors basicity. I- would
therefore be a weaker nucleophile than Br- because it is a weaker base. Verdict - A
strong/anionic nucleophile always favours SN2 manner of nucleophillic substitution.

Solvent[edit]

The solvent affects the rate of reaction because solvents may or may not surround a nucleophile,
thus hindering or not hindering its approach to the carbon atom. Polar aprotic solvents, like
tetrahydrofuran, are better solvents for this reaction than polar protic solvents because polar
protic solvents will hydrogen bond to the nucleophile, hindering it from attacking the carbon
with the leaving group. A polar aprotic solvent with low dielectric constant or a hindered dipole
end will favour SN2 manner of nucleophillic substitution reaction. Examples: DMSO, DMF,
acetone etc. In polar aprotic solvent, nucleophilicity parallels basicity.

Leaving group[edit]

The leaving group affects the rate of reaction because the more stable it is, the more likely that it
will take the two electrons of its carbon-leaving group bond with it when the nucleophile attacks
the carbon. Therefore, the weaker the leaving group is as a conjugate base, and thus the stronger
its corresponding acid, the better the leaving group. Examples of good leaving groups are
therefore the halides (except fluoride) and tosylate, whereas HO- and H2N- are not.

Reaction kinetics[edit]
The rate of an SN2 reaction is second order, as the rate-determining step depends on the
nucleophile concentration, [Nu−] as well as the concentration of substrate, [RX].

r = k[RX][Nu−]

This is a key difference between the SN1 and SN2 mechanisms. In the SN1 reaction the
nucleophile attacks after the rate-limiting step is over, whereas in SN2 the nucleophile forces off
the leaving group in the limiting step. In other words, the rate of SN1 reactions depend only on
the concentration of the substrate while the SN2 reaction rate depends on the concentration of
both the substrate and nucleophile. In cases where both mechanisms are possible (for example at
a secondary carbon centre), the mechanism depends on solvent, temperature, concentration of the
nucleophile or on the leaving group.

SN2 reactions are generally favored in primary alkyl halides or secondary alkyl halides with an
aprotic solvent. They occur at a negligible rate in tertiary alkyl halides due to steric hindrance.

It is important to understand that SN2 and SN1 are two extremes of a sliding scale of reactions, it
is possible to find many reactions which exhibit both SN2 and SN1 character in their mechanisms.
For instance, it is possible to get a contact ion pairs formed from an alkyl halide in which the
ions are not fully separated. When these undergo substitution the stereochemistry will be
inverted (as in SN2) for many of the reacting molecules but a few may show retention of
configuration. Sn2 reactions are more common than Sn1 reactions[citation needed].
E2 competition[edit]
A common side reaction taking place with SN2 reactions is E2 elimination: the incoming anion
can act as a base rather than as a nucleophile, abstracting a proton and leading to formation of the
alkene. This is more common when the incoming ion is sterically hindered in which case
abstracting a proton is much easier. Elimination reactions are usually favoured at elevated
temperatures.[3] This effect can be demonstrated in the gas-phase reaction between a sulfonate
and a simple alkyl bromide taking place inside a mass spectrometer:[4][5]

With ethyl bromide, the reaction product is predominantly the substitution product. As steric
hindrance around the electrophilic center increases, as with isobutyl bromide, substitution is
disfavored and elimination is the predominant reaction. Other factors favoring elimination are
the strength of the base. With the less basic benzoate substrate, isopropyl bromide reacts with
55% substitution. In general, gas phase reactions and solution phase reactions of this type follow
the same trends, even though in the first, solvent effects are eliminated.

Roundabout mechanism[edit]
A development attracting attention in 2008 concerns a SN2 roundabout mechanism observed in
a gas-phase reaction between chloride ions and methyl iodide with a special technique called
crossed molecular beam imaging. When the chloride ions have sufficient velocity, the energy of
the resulting iodide ions after the collision is much lower than expected, and it is theorized that
energy is lost as a result of a full roundabout of the methyl group around the iodine atom before
the actual displacement takes place.[

You might also like