You are on page 1of 23

Deciding SN1/SN2/E1/E2 (1)

– The Substrate
Deciding SN1/SN2/E1/E2: The Key Role Of The Alkyl Halide (“Substrate”)

Having gone through the SN1, the SN2, the E1, and the E2 reactions in turn, we can now say the
following:

 Both substitution reactions and elimination reactions occur with alkyl halides (and


related species). They do not occur with alkenyl (sp2-hybridized) or alkynyl (sp-
hybridized) halides
 Whether an alkyl halide is primary, secondary, or tertiary has a tremendous impact
on which reaction pathway the reaction  will follow. We have seen how the SN2
pathway is retarded by steric hindrance, and carbocation formation is favored by
increasing substitution by carbon   (i.e. tertiary > secondary > primary).
  A wide variety of nucleophiles/bases can be used to perform substitution and
elimination reactions.
 A wide variety of solvents can be used in substitution and elimination reactions
 We also have to gauge the importance of factors such as the leaving group and
temperature.
This is a lot of different factors to think about. Let’s look at some examples of situations you
might encounter:
This is often one of the most difficult parts of organic chemistry for new students: how to weigh
multiple (and often contradictory) factors? How do we know which factor is most important? Do
we pay attention to the base, substrate, temperature, solvent? How do we go about sorting
through a problem like this?

In this post and the next few, we’ll walk through one way of thinking about how to evaluate
whether a reaction will proceed through SN1/SN2/E1/E2. It’s not 100% foolproof*, but it’s a
decent enough framework for our purposes. Think of it as a set of 80/20 guidelines.  I call it this:

The Quick N’ Dirty Guide To Determining


SN1/SN2/E1/E2
It starts by asking questions. In order of importance, I think they are:
1. The substrate
2. The nucleophile/base
3. The solvent
4. The temperature
It’s also an approach where I tend to (at least in the beginning) ruling things out rather than
ruling things “in”. In other words, seek to decide what options are not possible, rather than
decide which are possible. It’s a subtle distinction, but a valuable one. Once you’ve crossed
certain reactions off the list, you can then start asking yourself which reactions would be most
consistent with the reaction conditions.

Remember: this is the “Quick N’ Dirty” guide!  There will be some exceptions! (more on


those at the bottom)

Before getting specific with each of those 4 questions, let’s start with the most important
question you can ask in any situation like the ones above.

The most important step in evaluating any reaction is first to ask yourself “what type of
functional group(s) are present in this molecule? This is because the type of functional group
will dictate the type of reaction(s) that can occur. Note that in the questions above, all of the
starting materials are alkyl halides or alcohols. Substitution/elimination reactions are possible
for these substrates; many other reaction types (addition, for example) are not.

Quick N’ Dirty Question 1: The Substrate


Given that we’re looking at alkyl halides/alcohols, it’s a reasonable expectation that we should
evaluate SN1/SN2/E1/E2. The next step is to identify the type of alkyl halide we are dealing
with.
Look at the carbon that contains the best leaving group. Typically this is Cl, Br, I or some
other group that can act as a good leaving group.
Ask yourself: is this carbon primary, secondary, or tertiary?

Given what we know about SN1, SN2, E1, and E2 reactions, we can say the following:

 The “big barrier” to the SN2 reaction is steric hindrance. The rate of SN2 reactions
goes primary > secondary > tertiary
 The “big barrier” to the SN1 and E1 reactions is carbocation stability. The rate of
SN1 and E1 reactions proceeds in the order tertiary > secondary > primary.
 The E2 reaction has no “big barrier”, per se (although later we will have to worry
about the stereochemistry)
So how can we apply what we know about each of these reactions to simplify our decision?

For Primary Carbons, Rule Out The SN1 and


E1
Quick N’ Dirty Rule #1: If the substrate is primary, we can rule out SN1 and E1,
because primary carbocations are unstable* (see below for exceptions).  You
cannot definitively rule out E2 yet, although I will spill the beans and say that it’s almost
certainly going to be SN2 unless you are using a very sterically hindered (“bulky”) base such as
tert-butoxide ion (e.g. potassium t-butoxide KOtBu).

For Tertiary Carbons, Rule Out The SN2


Quick N’ Dirty Rule #2: If the substrate is tertiary, we can rule out SN2, because tertiary
carbons are very sterically hindered.

If the substrate is secondary, we can’t rule out anything (yet).

As you can see, based on the information we’ve evaluated so far, we can’t make a definitive
decision on SN1/SN2/E1/E2. We’ll have to look at some other factors before we can make a
final decision. Next, we’ll evaluate the role of the nucleophile/base.

Next Post: The Role of The Nucleophile

—————-END OF QUICK N’ DIRTY GUIDE, PART 1 ————————–


NOTES: One final word of warning on the substrate: SN1/SN2/E1/E2 reactions tend not to
occur on alkenyl or alkynyl halides. So if you see one of the substrates below, it is highly
likely that no reaction will occur. 

Why are alkenyl and alkynyl halides so bad? Well, the SN1, SN2, and E1 mechanisms all
involve considerable build-up of positive charge on the carbon bearing the leaving group, and the
stability of sp2 and sp hybridized carbocations is much lower than that for sp3 hybridized
carbocations [for the same reason that sp and sp2 anions are more stable than sp3 carbanions!].

E2 reactions are also more difficult due to the stronger C-H bonds of alkenes. [We’ll see later
that there is one example of an E2 that can occur on alkenyl halides, but the point remains that
they are very rare!]

—————— EXCEPTIONS ————–

* One question that comes up a lot is this: are there exceptions? Keeping in mind the two themes
of “steric hindrance” and “carbocation stability”, there are edge cases where we can have a
particularly sterically hindered primary alkyl halide, or a particularly stable primary carbocation.

For instance, the alkyl halide below (“neopentyl bromide”) is indeed primary, but is so crowded
on the carbon adjacent to the primary alkyl halide that it is essentially inert in SN2 reactions. On
the SN1/E1 side, the allyl halide below, while primary, can undergo SN1/E1 reactions because
the resulting carbocation is stabilized through resonance. As long as you keep in mind the “big
barriers” for each reaction, you should be fine.
Deciding SN1/SN2/E1/E2 (2)
– The Nucleophile/Base
SN1/SN2/E1/E2 Decision: The Role of the Nucleophile
Last time I talked about the process of deciding if a reaction goes through SN1, SN2, E1, or E2 as
asking a series of questions. I call it The Quick N’ Dirty Guide To SN1/SN2/E1/E2. This is the
second instalment.

After Determining Whether Your Substrate Is


Primary, Secondary, Tertiary, or Methyl,
Examine The Nucleophile/Base
Once we’ve looked at a reaction and recognized that it has the potential for proceeding through
SN1/SN2/E1/E2 – that is, is it an alkyl halide, alkyl sulfonate (abbreviated as OTs or OMs) or
alcohol – and asked whether the carbon attached to the leaving group is primary, secondary, or
tertiary, we next can look at the reagent for the reaction.

In substitution reactions, a nucleophile forms a new bond to carbon, and a bond between the
carbon and the leaving group is broken. In elimination reactions, a base forms a new bond with a
proton from the carbon, the C-H bond breaks,  a C-C π bond forms, and a bond between carbon
and leaving group is broken.

There’s a lot of confusion from students on this point. “How do I know what’s a nucleophile and
what’s a base?”.

Whether something is a nucleophile or a base depends on the type of bond it is forming in the


reaction. Take a species like NaOH. It’s both a strong base and a good nucleophile. When it’s
forming a bond to hydrogen (in an elimination reaction, for instance), we say it’s acting as a
base. Similarly, when it’s forming a bond to carbon (as in a substitution reaction) we say
it’s acting as a nucleophile. 

It’s a relationship, in other words. For instance, when I’m interacting with my wife, I’m
interacting with her as a husband. When I’m talking to my mom, I’m interacting with her as a
son. I’m the same person, but depending on whom I’m interacting with, our relationship has
different names.

Anyway. All this is prelude to making the key determination for today, which is:

1. Charged bases/nucleophiles will tend to perform S 2/E2 reactions.


N

2. Reactions where neutral bases/nucleophiles are involved tend to go


through carbocations (i.e. they tend to be S 1/E1). N

Again: Quick N’ Dirty is an 80/20 set of principles. There are exceptions!!! But


this framework will help us in most situations.
What Is A “Charged”  Nucleophile or Base?
Let’s talk about charged nucleophiles first. It’s important to be able to recognize charged
nucleophiles. The charges are often not written in, but “implied”. For example, NaOEt (sodium
ethoxide) actually has an ionic bond between Na(+)  and (-)OEt, even though the charges
themselves aren’t written in (us chemists are lazy that way). So if you see Na, K, or Li, for
instance, you’re looking at a charged nucleophile/base. Whether it’s K, Na, or Li doesn’t matter
for our purposes – these are just spectator ions.

In both the SN2 and E2 pathways the reaction is “concerted” – that is, the nucleophile/base forms
a bond as the C-LG bond is broken. Since there is significant bond-breaking occurring in the
transition state, the energy barrier for this step is higher than in the (second) step of the E1 or
SN1; we’re going to require a stronger nucleophile/base to perform these reactions. Recall
that the conjugate base is always a stronger nucleophile. Negatively charged species have a
higher electron density and are more reactive than their neutral counterparts.

Quick N’ Dirty Rule #3:


If you see a charged nucleophile/base, you can rule out
carbocation formation (i.e. rule out SN1/E1). In other
words, the reaction will be SN2/E2. 
We can break things down even more, depending on how strong a base the charged species is; go
to the section at the bottom of this post for some examples where we can use base strength to
rule out E2.

Reactions of Neutral Bases/Nucleophiles

Neutral bases/nucleophiles tend to be weaker than negatively charged bases/nucleophiles. In


order for them to participate in substitution or elimination reactions, generally the leaving group
must depart first, giving a carbocation.

Quick N’ Dirty Rule #4: If you don’t see a charged species present, you’re likely looking at a
reaction that will go through a carbocation (i.e. an SN1 or E1).
One special case worth noting is if you see a strong acid such as H2SO4 or HCl with an alcohol as
a substrate. Unless you’re looking at a primary alcohol (where carbocations are very unstable)
the reactions in these cases will almost always proceed through carbocations.

It’s not uncommon to see a neutral nucleophile in the presence of a charged one (see example 2,
below). In that case it’s likely acting as the solvent. We’ll talk about solvents next. 

Here’s a chart where we evaluate this second question for deciding if a reaction is SN1, SN2, E1,
or E2 (below).
What’s the biggest weakness of the Quick N’ Dirty approach? It’s an oversimplification. To
conclude that a reaction “proceeds SN2” or “proceeds E2” might give the impression that it gives
100% SN2 or 100% E2, and that is surely not the case! Often, these reactions compete with each
other, and can therefore give mixtures products. When I say “SN2” , for instance, I mean mostly
SN2. There are likely other products in there.  

The key lesson here is to understand the concepts – “what conditions favor each reaction?” and
then to be able to apply the rules you know about each reaction to draw the proper product.

Next Post: The Role of Solvent

——–END QUICK N’ DIRTY GUIDE TO SN1/SN2/E1/E2 PART 2 ——————

Elaboration: Good Nucleophiles That Are Weak Bases

Some charged nucleophiles are actually poor bases. Here’s a good rule of thumb: if the conjugate
acid of the base/nucleophile is less than 12, an E2 reaction will be extremely unlikely. So if you
see a nucleophile like NaCl, NaBr, KCN, and so on, it will favor SN2 over E2.

(This is a more rigorous way of saying that weak bases don’t perform E2 reactions).
In contrast, the bulky base below (tert-butoxide ion) is a strong base but a poor nucleophile due
to its great steric hindrance, so an E2 reaction is much more likely than SN2.

Exception: Neutral Nucleophiles in SN2 and E2 Reactions

One class of neutral nucleophiles/bases that readily perform E2 reactions (and SN2) are amines.
For example, the tertiary alkyl halide below will undergo elimination through E2 here, although
the Quick N’ Dirty rules call for SN1/E1. Amines are generally not the most useful nucleophiles
for doing SN2 however because they lead to over-alkylation and ammonium salt
formation. [See: Alkylation of Amines (Sucks!)]

Finally, there are also neutral species which are good nucleophiles (and poor bases) such as PPh3,
below.

Exception: Charged Nucleophiles In SN1 Reactions

It’s also possible to use charged nucleophiles in SN1 reactions under certain conditions. If you
have, for instance a tertiary alkyl halide in the presence of a high concentration of a  good
nucleophile (but weak base) such as those above, the carbocation that forms can be intercepted
by that nucleophile. For example:
Here, the good nucleophile (cyanide ion), if present in large excess, can overpower the weak
nucleophile  (solvent). Of course the ultimate arbiter of such statements are actual experiments.

Deciding SN1/SN2/E1/E2 (3)


– The Solvent
The Quick N’ Dirty Guide To SN1/SN2/E1/E2 Reactions, Part 3: The Role of Solvent
Let’s continue with our Quick N Dirty guide to SN1/SN2/E1/E2 – a quick walkthrough of
thinking through this reaction decision.

After having examined the substrate and the base/nucleophile in an SN1/SN2/E1/E2 reaction,


this post is about the next question to ask:

What’s The Solvent? 

Recall that there are two important types of solvents to consider: polar protic solvents and polar
aprotic solvents. [See: All About Solvents]

Let’s do a little review,  looking at polar protic solvents first.

Polar protic solvents are capable of hydrogen bonding. Recall that hydrogen bonding occurs
where we have a highly electronegative atom such as O or N directly bonded to hydrogen:

Quick N’ Dirty Tip: Solvents with OH or NH


groups are polar protic solvents
Hydrogen bonding is directly responsible for the high boiling points of solvents such as water
and ethanol;  the partial positive charges on hydrogen are attracted to the partial negative charges
on the electronegative atoms.  This is also why water is such an excellent solvent for charged
species such as halide ions;  hydrogen bonding solvents surround negatively charged ions like a
jacket.
Polar Protic Solvents “Cling” To Nucleophiles
via Hydrogen Bonding, And Nucleophilicity
Goes Up As We Go Down  The Periodic Table
This “jacket” of solvent molecules – much like a protective crowd of bobbies –  means that these
anions do not have the freedom of action they would normally have if they weren’t surrounded
by clingy hordes of solvent molecules. That is, nucleophiles are made less nucleophilic!  The
propensity to form hydrogen bonds is highest for small, highly electronegative ions such as
fluorine and decreases as ions get larger (and the charge is more diffuse).

This means that in polar protic solvents, the nucleophilicity of halide anions increases as we


go down the periodic table.

Polar Aprotic Solvents Do Not Hydrogen-Bond


With Nucleophiles, And Therefore
Nucleophilicity In These Solvents Correlates
With Basicity
Now let’s talk about polar aprotic solvents; polar aprotic solvents are polar enough to dissolve
charged species (such as halide ions) but do not donate hydrogen bonds. This means that in
solvents such as DMSO, DMF, acetone, or acetonitrile, nucleophilicity correlates much better
with basicity  (and bond strength, as C-F > C-Cl > C-Br  > C-I ) – and therefore
nucleophilicity decreases as we go down the periodic table.

So what’s the bottom line when it comes to SN1/SN2/E1/E2?

The bottom line is this:

 Quick N’ Dirty Rule #5: Polar protic solvents tend to favor elimination (E2)


over substitution (SN2). Polar aprotic solvents tend to favor substitution (SN2)
relative to elimination (E2)

Let’s go back to the examples we were looking at in the first two posts.
Practically, you’ll only need to consider the solvent in an SN1/SN2/E1/E2 decision when
you’ve already analyzed the substrate and the nucleophile/base.
This is usually the case when you have a secondary alkyl halide with a strongly basic
nucleophile such as NaOCH3 or NaOH.
That’s especially the case in example #2, where we really can’t make the call as to whether it’s
SN2 or E2 until we’ve looked at the solvent. The fact that we are using a polar protic solvent
(EtOH) is the crucial piece of information that clinches the reaction as E2.
In the last post in this series we’ll look at the impact of temperature on these reactions.
Next Post: The Role Of Temperature
[Edit Dec 13]: What’s the weakness of the Quick N’ Dirty approach? Well, it doesn’t really
capture the true situation in these reactions, that is, that they don’t always give one product
exclusively. So keep in mind that when the Quick N’ Dirty approach says that E2 is the major
product, you may also have a small amount of competing SN2 occurring as well, or if SN1 is
favored, you might have small amounts of other reaction products also.  The key points of this
exercise is to 1) remember the major factors that affect each type of reaction, and 2) once you
know the type of reaction, be able to apply it.
Deciding SN1/SN2/E1/E2 (4)
– The Temperature
The Quick N’ Dirty Guide To Determining SN1/SN2/E1/E2, Part 4 : The Role of
Temperature

In previous installations of the Quick N’ Dirty Guide, we’ve examined the substrate,


the base/nucleophile, and the solvent. Today, we’ll address the final variable to consider: the
temperature.

If you’ve been following so far, you may have noticed that by this point we should be able to
differentiate all cases where SN2 is favored over E2 (and vice versa) but are still left with this
dilemma: when a carbocation is formed, how do we determine whether SN1 or E1 products are
favored?

How Do We Determine Whether SN1 versus E1


Will Dominate Once A Carbocation Has
Formed?
First of all, note that the first step of the SN1 and E1 reactions is the same:  loss of a leaving
group to give a carbocation. Since both of these reactions proceed via the same intermediate, in
practice a mixture of both SN1 and E1 products will be found whenever the reaction proceeds
through a   carbocation [where possible].  Given that, however, we would still like to have a rule
of thumb that tells us what type of product should be the major product in most cases.

Generally speaking, SN1 products tend to predominate over E1 products at lower


temperatures.  However, recall that elimination reactions are favored by heat.  In cases where
substitution reactions and elimination reactions are in competition with each other, increasing
the temperature tends to increase the amount of elimination products produced. 

Here is a representative example:


Loss of bromide ion from the substrate leads to the formation of a tertiary carbocation [stable,
hence no rearrangement]. At low temperatures, the SN1 pathway (above) will dominate: attack at
the carbocation by CH3OH, followed by loss of proton to give the ether. The bottom pathway –
removal of hydrogen from the carbon adjacent to the carbocation – will be minor at low
temperature [note the formation of the more substituted alkene here – Zaitsev’s rule in action].
As temperature is increased, the amount of elimination relative to substitution should gradually
increase.

This leads to the following Quick N’ Dirty rule of thumb.

If “Heat” Is Noted, The Reaction Will Favor E1


Over SN1
Quick N’ Dirty Rule #6:  When carbocations are formed, at low temperatures, the SN1 pathway
will dominate over the E1 pathway. At higher temperatures, more E1 products will be formed.

(Note: before applying these reaction patterns to the substrate, make sure to examine the
carbocation that is formed. If a more stable carbocation can be formed through a hydride or alkyl
shift, do this rearrangement first!)

Let’s go back to the examples we’ve been working on.

The third case – addition of H2SO4 to a tertiary alcohol – is a case where a carbocation is
formed in the absence of a good nucleophile [the negatively charged oxygen on the conjugate
base, [HSO4(-)] is stabilized through resonance, reducing its reactivity]. The fact that heat is
being applied helps to tip the balance even further toward E1 being dominant over SN1.

In the fourth example we have a tertiary halide [which will form a stable carbocation] in a polar
protic solvent [will help to stabilize the intermediate carbocation] and heat is not indicated.
Therefore using Quick N’ Dirty Rule #6, we can say that SN1 products will dominate. [E1
products will form as well, but they will not be the major products].

This is truly a Quick N’ Dirty rule. It is not applied evenly and there are plenty of exceptions.
Your mileage will vary widely. I wish I had a hard, concrete example to show you that clearly
demonstrates the relationship between increased heat and a greater proportion of elimination
(E1) versus substitution (SN1) products. Sadly, I cannot find a good example at this time. So a
hand-wavy “elimination increases as heat is applied” will have to do for now.*

In the next post we’ll summarize all the Quick N’ Dirty Rules for determining whether a reaction
goes SN1/SN2/E1/E2.

Next Post: Wrapup of the Quick N’ Dirty Guide


———–END QUICK N’ DIRTY DISCUSSION ————-

Notes
* I am annoyed by the lack of hard data available when making pronouncements about the
SN1/SN2/E1/E2 decision. Wohler’s quote about the  “monstrous and boundless thicket… into
which one may well dread to enter” seems appropriate here.  I would love to see the control
experiments with various substrates run under identical sets of conditions that clearly delineate
the impact of each variable. I have not seen this. Any undergraduate labs out there with a desire
for performing this valuable public service?

A final note/editorial.

One Difference Between “Real ” Chemistry and “Exam Question” Chemistry

The post above is not so much about understanding some facet of organic chemistry as it is about
how to answer some arbitrary question from a textbook or exam. For the purpose
of understanding organic chemistry, it’s enough to know that heat favors elimination reactions.
For the purpose of knowing how to answer a particular kind of question on an exam, it will of
course depend on the examiner. There are wide variations. However, I will share with you one
common observation that I’ve seen in 2.5 years of seeing exams from all over the country.

First, background.  In the laboratory, it is extremely common to heat reactions to get them to go
at a reasonable speed, such as in this example [from March’s Advanced Organic Chemistry 5th
ed; Cooper, K.A. et. al. J. Chem. Soc. 1948, 2038:]

However, on an exam, instructors – for various reasons, including a well-intentioned desire not
to overwhelm the student – will often omit some of the data. For exam purposes, if the above
reaction were written as a  question it will often look like this:
 Note how “heat” has been omitted, which is in accord with the principle of least effort. The
expected answer in this instance would be t-BuOH, the product of an SN1 reaction. Depending
on the question wording, some instructors will also insist that the E1 product be drawn as well.

Here’s the observation I see in many (but certainly not all) courses.  If the word “heat” is
written in the exam question, it is often a clue from the instructor that an elimination is to
take place. In the following reaction, for example, the question would point to elimination (2-
methyl propene) being the major product.

If you are a student and your goal is to answer a particular type of question on an exam
correctly, I advise you to double check this issue with your instructor and get their answer on it.
There is tremendous inconsistency in this practice nationwide.

 
Wrapup: The Quick N’
Dirty Guide To
SN1/SN2/E1/E2
The basic premise is this: given 15-20 minutes to describe the basic principles by which one
could figure out if a given reaction goes down one of these pathways, these are, in my opinion,
the key factors to consider.

Quick N’ Dirty rules, by their nature,  do not cover exceptions. To learn about some of the
exceptions, I advise you to go back and read the individual posts [One Two Three Four].

Even further back, I urge you to understand the key concepts behind each reaction, such
as nucleophilicity, leaving group ability, carbocation stability, and the mechanism of each of
these reactions. [SN1] [SN2] [E1] [E2]

Finally, I will preface this by saying that the best way to learn and understand how these
reactions work is to do a lot of practice problems and pay particular attention to situations where
you get the wrong answer – they are instructive.

Here goes:

Question 1: Is the carbon containing the leaving


group methyl (only one carbon), primary,
secondary, or tertiary?
 Quick N’ Dirty Rule #1: If primary, the reaction will almost certainly be
SN2 [prominent, commonly encountered exceptions: 1) a bulky base such as tBuOK
will tend to give elimination products [E2]; 2) primary carbons that can form
relatively stable carbocations (i.e. allylic/benzylic) may proceed through the SN1/E1
pathway.] Also – methyl carbons always proceed through SN2.
 Quick N’ Dirty Rule #2: If tertiary, the reaction cannot be SN2. [Because tertiary
alkyl halides are too hindered for the SN2.  Depending on the type of nucleophile/base,
it will either proceed with concerted elimination [E2] or through carbocation
formation [SN1/E1]
Question 2: Does the nucleophile/base bear a
negative charge?
 Quick N’ Dirty Rule #3: Charged nucleophiles/bases will favor SN2/E2
pathways  [i.e. rule out SN1/E1]. [So, for example, if SN2 has already been ruled out
[e.g. for a tertiary carbon, according to Question 1] then the reaction will therefore be
E2. This is the case for tertiary alkyl halides in the presence of strong bases such as
NaOEt, etc. The strength of the [charged] nucleophile/base can be important!  An
important special case is to be aware of charged species that are weak bases [such as
Cl, N3, –CN, etc.] these will favor SN2 reactions over E2 reactions].
 Quick N’ Dirty Rule #4: If a charged species is not present, the reaction is likely
to be SN1/E1. [so if the only reagent is, say, H2O or CH3OH you are likely looking at
carbocation formation  resulting in an SN1/E1 reaction.]

Question 3: Is the solvent polar protic or polar


aprotic?
 Quick N’ Dirty Rule #5: All else being equal, polar aprotic solvents favor
substitution [SN2] over elimination [E2]. Polar protic solvents favor elimination
[E2] over substitution [SN2]. [Note that this rule is generally only important in the
case of trying to distinguish SN2 and E2 with a secondary alkyl halide and a charged
nucleophile/base. This is not meant to distinguish SN1/E1 since these reactions tend to
occur in polar protic solvents, which stabilize the resulting carbocation better than
polar aprotic solvents.]

Question 4: Is heat being applied to the


reaction?
 Quick N’ Dirty Rule #6: Heat favors elimination reactions. [This only becomes an
important rule to apply when carbocation formation is indicated and we are trying to
decide whether SN1 or E1 will dominate. At low temperatures SN1 products tend to
dominate over E1 products; at higher temperatures, E1 products become more
prominent.]
Writing this post makes me feel like a nun giving out condoms.  I realize there will be many who
are reading this an hour before their exam and are completely clueless on this subject. All I have
to say is, God help you. And do more fricking practice problems so you don’t put yourself in this
situation next time.

You might also like