3.2.1 – Hydrohalogenation of Alkenes

Hydrohalogenation of Alkenes

The addition reaction of a hydrogen halide molecule, such as HCl or HBr, to an alkene produces an alkyl halide as a product. This process is known as a Hydrohalogenation reaction.

 

Hydrogen bonds to one carbon of the alkene and Cl bonds to the other carbon, forming an alkane.
Figure 3.2.1.a. Addition reaction of the hydrogen halide HCl to an alkene
Hydrogen bonds to one carbon of the alkene and Br bonds to the other carbon, forming an alkane (cycloalkane in the reaction).
Figure 3.2.1.b. Addition reaction of the hydrogen halide HBr to an alkene

The mechanism (Figure 3.2.1.c.) for hydrohalogenation is a two-step process. In the first step, the alkene is protonated by HBr, resulting in a carbocation intermediate and a halide anion (such as Br). In this protonation step, the π bond is broken to allow for formation of a sigma bond to the proton. This results in a carbon that loses π electron density and becomes cationic. In the second step, the halide ion formed in the first step acts as a nucleophile to attack the electrophilic carbocation intermediate. The result of the second step is that a new C–Br σ bond is formed.

Recall that in addition reactions, one π bond and one σ bond are broken, and two σ bonds are formed. In this mechanism, the first step involves breaking the H–Br σ-bond and the C=C π-bond and forming a new C–H σ-bond, while the second step involves forming a new C–Br σ bond. In addition, the hybridization at carbon changes: the C=C double bond of reactants have sp2 hybridized carbon atoms, which become sp3 hybridized in the products.

This reaction mechanism has some similarities with the SN1 mechanism discussed in Chapter 3.1.3. For example, hydrohalogenation and unimolecular nucleophilic substitution (SN1) are similar in that both mechanisms occur in two steps and involve a carbocation intermediate. In addition, the second step of both mechanisms is identical: the electrophilic carbocation is attacked by the nucleophilic halide to form a new C–X bond. SN1 reactions and hydrohalogenation differ in the first step of the reaction: an SN1 mechanism involves loss of a leaving group, whereas an alkene has no leaving group and instead is protonated in the first step.

Addition reaction of cyclohexene and HBr mechanism.
Figure 3.2.1.c. Mechanism of hydrohalogenation addition reaction to a symmetric alkene. The nucleophilic π electrons in the alkene attack the electrophilic proton, breaking the H–Br bond and forming a new C–H bond, resulting in a carbocation intermediate. After the breaking of the π bond, one of the carbon atoms has only 3 bonds and becomes cationic. The anionic, nucleophilic bromide can then attack the carbocation and form a new C–Br bond.

In the above examples, the alkenes are symmetrical in structure, with each alkene carbon existing in identical bonding environments. As a result. it is observed that there is no preference between which atom the halide or hydrogen atom are added to This is because the reaction pathway will proceed via an identical carbocation intermediate, so there is no preference on which side the C–H vs. C–X bond is formed, and the final product will be the same.

Cyclohexene and HBr can yield a compound where the H can go either on the carbon above in the alkene or below in the alkene forming carbocations, and the Br bonds to the other carbon. Although it may seem there are 2 different types of compounds, it’s actually the same compound 1-bromocyclohexane, just flipped.
Figure 3.2.1.d. Hydrohalogenation of symmetrical alkenes yields the same carbocation intermediate and the same product.

For an asymmetrical alkene, where the double bonded carbons atoms have different substituents, the placement of the H and X atoms across the π bond is critical. For example, in the following reaction (Figure 3.2.1.e.), two possible products could be produced: 1-bromo-1-methylcyclohexane and 1-bromo-2-methylcyclohexane. Experimental observations demonstrate that 1-bromo-1-methylcyclohexane is the main product for the reaction. To explain and understand the outcome of the reaction, we need to consider the mechanism of the reaction and the stability of the intermediates formed.

Hydrohalogenation of 1-methylcyclohexene with HBr.
Figure 3.2.1.e. Hydrohalogenation of 1-methylcyclohexene yields two possible products. The major product is 1-bromo-1-methylcyclohexane; the minor product is 1-bromo-2-methylcyclohexane.

Are You Wondering? The Difference Between a Major and a Minor Product

In reactions where multiple products can form, the product that is formed in over 50% of the theoretical yield is referred to as the major product. This distribution in products is due to the energy required to form it. The major product requires less energy to form, so more will form per unit time.

In reactions where multiple products can form, the product that is formed in less than 50% of the theoretical yield is referred to as the minor product. This distribution in products is due to the energy required to form it. The minor product requires more energy to form, so less will form per unit time.

When the nucleophilic π bond attacks the electrophilic proton  from H–X, a carbocation intermediate is formed. Recall carbocation stability (see Chapter 3.1.3): higher substituted carbocations are the most stable due to greater neighbouring electron density stabilization, while lowly substituted carbocations are less stable due to a lack of neighbouring electron density available to provide charge dissipation.

The energy profile diagram for the hydrohalogenation of symmetric alkenes is shown below:

Energy profile diagram of the reaction of cyclohexene and HBr. The curve is shaped like an Sn1 energy profile, where firstly one proton is bonded to the alkene, generating a carbocation intermediate. Then the bromide (Br-) is bonded to the other carbon generating a neutral bromocycloalkane.
Figure 3.2.1.f. Energy profile diagram for the hydrohalogenation reaction of the symmetric alkene, cyclohexene. The 2 maxima represent the two steps of the reaction. The first step (protonation of the alkene) is the slow step, hence the greater activation energy. The second step (nucleophilic attack of the carbocation by bromide) requires less energy and is therefore faster.

The first step of the reaction, protonation of the alkene, yields the carbocation intermediate. Since the carbocation intermediate is unstable and high in potential energy, there is a large activation barrier associated with this step resulting in it being the rate limiting step. The second step, nucleophilic attack of the carbocation by bromide, results in the final alkyl halide product.

Since the first step of the addition mechanism is rate limiting, we will focus on it to justify the differences in product formation. Just as carbocation stability is important when discussing the unimolecular nucleophilic substitution (SN1) mechanism, so too does it play a role in the hydrohalogenation mechanism (Figure 3.2.1.g). The first step of the hydrohalogenation mechanism adds a new C–H bond to one carbon atom and a formal positive charge to the other carbon atom (the carbocation). When looking at the possible carbocation intermediates formed, if the C–H bond was created on the more substituted carbon (tertiary position), then a secondary carbocation would form on the other carbon atom, leading to the 1–bromo–2–methylcyclohexane product (Figure 3.2.1.g, bottom). On the other hand, if the C–H bond was created on the less substituted carbon (secondary position), then a tertiary carbocation is formed on the other carbon atom, leading to the 1–bromo–1–methylcyclohexane product (Figure 3.2.1.g, top). Since a tertiary carbocation intermediate is more stable than a secondary carbocation intermediate, the tertiary carbocation intermediate will have a smaller activation energy barrier associated to its formation, and will form faster. As a result, more tertiary carbocation intermediate produced per unit time result in more alkyl halide product produced per unit time, leading to a higher abundance of 1-bromocyclohexane (major product).

Therefore, when performing hydrohalogenation reactions to alkenes, the electrophilic hydrogen atom is added to the least substituted carbon because it results a more substituted and more stable carbocation The halide will then attack the more substituted position in the second step of the reaction. This reaction is said to be regioselective – there is a preference to which carbon atoms the hydrogen and halide form sigma bonds too. In this case, we call this Markovnikov regioselectivity , named after Vladimir Markovnikov who studied alkene addition reactions extensively in the mid 1800s. As a result, we apply something known as Markovnikov’s  rule to hydrohalogenation reactions, which states that the carbon that is already bonded to more hydrogen atoms (less highly substituted position) gets the hydrogen. This rule is also cleverly coined as the “rich get richer” rule, where a carbon atom with more hydrogen atoms bonded to it is considered “rich”.

Are You Wondering? The Meaning of the Term Regioselectivity

Regioselectivity is when a reaction mechanism favours bond formation at a particular atom over other possible atoms. For alkene additions, we observe two types of regioselectivity – Markovnikov and Anti-Markovnikov. A Markovnikov regioselective product results from a mechanism that would result from the more stable of two possible carbocation intermediates. An Anti-Markovnikov regioselective product results from a mechanism that would result from the less stable of two possible carbocation intermediates.

Addition reaction of 1-methylcyclohexene and HBr generating 2 carbocations: a tertiary, more stable carbocation which leads to the major product, and a secondary, less stable carbocation leading to the minor product.
Figure 3.2.1.g. Hydrobromination of 1–methylcyclohexene. The π electrons in the alkene attack the electrophilic hydrogen, breaking the H-Br bond. This causes a new C–H bond to be formed. On the top pathway, the new C–H bond is formed on the less substituted carbon, generating a tertiary carbocation on the more substituted carbon. On the bottom pathway, the new C–H bond is formed on the more substituted carbon, generating a secondary carbocation on the less substituted carbon. In the second step, the nucleophilic bromide anion attacks the carbocation. This yields two products: a major (Markovnikov) product (top pathway) and a minor product (bottom pathway).

(The full solution to this problem can be found in Chapter 5.2)

Another way of naming the major and minor product of a reaction when there is Markovnikov regioselectivity is the Markovnikov and Anti-Markovnikov products.

1-Methylcyclohexene plus HBr reacting, indicated by a reaction arrow, to form 1-bromo-1-methylcyclohexane plus 1-bromo-2-methylcyclohexane.
Figure 3.2.1.h. Hydrobromination of 1–methylcyclohexene yields two products: 1–bromo–1–methylcyclohexane, the Markovnikov product, and 1–bromo–2-methylcyclohexane, the Anti-Markovnikov product.

The Markovnikov product is formed through a lower energy pathway, where a more stable tertiary carbocation intermediate is formed, and the hydrogen atom is added to the carbon bonded to more hydrogens. The Anti-Markovnikov product is formed through a higher energy pathway, where a less stable secondary carbocation intermediate is formed, and the hydrogen atom is added to the carbon bonded to less hydrogens.

The energy profile diagram for the hydrohalogenation of alkenes is shown below (Figure 3.1.2.i).

Energy profile diagram of HBr and 1-methylcyclohexene reagents. There are 2 curves stacked on each other with the higher energy one in pink having a greater peak than the bottom one in green. The pink pathway is more difficult to form leading to the minor product 1-bromo-2-methylcyclohexane. The green pathway is faster and forms the major product 1-bromo-1methylcyclohexane.
Figure 3.2.1.i. Energy profile diagram for the electrophilic addition of HBr to 1–methylcyclohexene. The 2 maxima represent the two steps of the reaction. The first step (protonation of the alkene) is the slow step, hence the greater activation energy. The second step (nucleophilic attack of the carbocation by bromide) requires less energy and is therefore faster. The blue pathway is higher in energy because the secondary carbocation intermediate is less stable; the green pathway is lower in energy because the tertiary carbocation intermediate is more stable.

The two maxima represent the two steps of the reaction, where the first step (protonation of the alkene) is the rate-limiting step. Carbocation stability dictates which product is formed: the more highly substituted the carbocation, the more stable it will be, and the reaction pathway that involves the more stable carbocation will require less energy to overcome the activation energy barrier and proceed faster.

(The full solution to this problem can be found in Chapter 5.2)

The following video includes a worked example from a previous CHEM 1AA3 test or exam that students struggled with. Try solving it on your own before looking at the solution.

 

Key Takeaways

  • Hydrohalogenation is a reaction using an alkene and a hydrogen halide (HX) as reactants, where the hydrogen and halogen are added to each alkene carbon
  • The mechanism involves two steps, the first being rate limiting:
    • Step 1: The π electrons of the double bond nucleophilically attack the proton and form a σ bond to it, creating a carbocation and a halide as intermediates
    • Step 2: The halide nucleophilically attacks the carbocation, creating the product: an alkyl halide
  • If the alkene is asymmetrical, then there’s two possible products: major and minor
  • Because of the carbocation stability, the more substituted carbocation intermediate will be the intermediate that leads to the major product
  • The major product is also called the Markovnikov product, while the anti–Markovnikov product is also the minor product

Key terms in this chapter:

Key term Definition
Hydrohalogenation reaction The addition reaction of a hydrogen halide to an alkene. This produces an alkyl halide as a product.

 

Any feedback or comments on this chapter? You may either email chemoer@mcmaster.ca, access this MS Form, or provide a comment in the feedback box below.

definition

License

Icon for the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License

Organic Chemistry and Chemical Biology for the Students by the Students! (and the Profs...) Copyright © 2023 by Emma Abreu; Anumta Amir; Anthony Chibba; Jim Ghoshdastidar; Sharonna Greenberg; Angela Liang; Layla Vulgan; and Shuoyang Wang is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License, except where otherwise noted.

Share This Book