Frequently Asked Questions
An electrophilic substitution reaction is a type of chemical reaction in which an electrophile selectively replaces an atom or group in a molecule, commonly a hydrogen atom, through the attack on an electron-rich site. This class of reactions is especially prominent in aromatic compounds, where the stable π-electron cloud of the aromatic ring provides a nucleophilic center that electrophiles can attack.
In the context of aromatic chemistry, electrophilic substitution reactions maintain the aromaticity of the ring by temporarily disrupting it during the formation of an intermediate arenium ion, followed by re-aromatization through loss of a proton. This allows functional groups to be introduced onto the ring without compromising its unique electronic stability.
Examples include:
- Nitration of Benzene: Using a mixture of concentrated nitric acid and sulfuric acid generates the nitronium ion (NO₂⁺), which substitutes a hydrogen on benzene to give nitrobenzene.
- Sulfonation: Benzene reacts with fuming sulfuric acid (oleum) to introduce a sulfonic acid (-SO₃H) group on the ring.
- Halogenation: Benzene reacts with halogens (Cl₂ or Br₂) in the presence of a Lewis acid catalyst (FeCl₃ or FeBr₃), resulting in chloro- or bromobenzene.
- Friedel-Crafts Alkylation: An alkyl group is introduced onto benzene using an alkyl halide and a Lewis acid catalyst like AlCl₃.
- Friedel-Crafts Acylation: An acyl group (RCO-) is attached to benzene using an acyl chloride and AlCl₃ catalyst.
These examples show the diversity of functional groups that can be introduced via electrophilic substitution, enabling synthesis of a wide range of aromatic derivatives with applications in dyes, pharmaceuticals, and polymers.
Benzene prefers electrophilic substitution over addition because addition reactions disrupt its aromatic stability.
Benzene’s aromaticity arises from a delocalized π-electron system, which lowers its energy and stabilizes the molecule. Electrophilic substitution temporarily disturbs this aromatic system but quickly restores it, keeping benzene’s inherent stability intact.
Addition reactions, on the other hand, saturate the double bonds and break aromatic conjugation, resulting in a much less stable, non-aromatic product. This loss of aromaticity makes addition energetically unfavorable.
Thus, benzene undergoes substitution to preserve aromaticity, which drives the reaction preference.
The typical steps in an electrophilic aromatic substitution (EAS) reaction are:
- Generation of the Electrophile: The electrophile (E⁺) is prepared or activated, often requiring a catalyst. For example, in nitration, sulfuric acid generates the nitronium ion.
- Electrophilic Attack: The aromatic ring acts as a nucleophile and donates electrons to the electrophile, forming a sigma complex or arenium ion intermediate. This step disrupts aromaticity and is the rate-determining step.
- Deprotonation: A base removes a proton from the sigma complex to restore the aromatic system.
- Formation of Substituted Aromatic Compound: The final substituted product is formed with the electrophile replacing a hydrogen on the ring.
Between anisole and phenol, phenol typically shows a faster rate of electrophilic aromatic substitution.
Phenol (-OH group) is a stronger activating group than the methoxy group (-OCH₃) in anisole because the lone pair on the oxygen is more directly available for resonance donation into the ring. Additionally, phenol’s ability to engage in hydrogen bonding can stabilize the transition state and intermediates further.
As a result, phenol’s ring is highly activated, making it more reactive toward electrophiles. This is why phenol readily undergoes bromination in water without a catalyst, whereas anisole requires slightly more forcing conditions.
In electrophilic aromatic substitution, C (para-substituted) product is often the most favored due to steric and electronic factors.
The para position is opposite the substituent already present on the ring, allowing the incoming electrophile to attach with minimal steric hindrance compared to the ortho positions, which are adjacent and more sterically congested.
Electronically, when the substituent is an activating group (such as -OH or -OCH₃), resonance structures show significant electron density at the para position, stabilizing the intermediate arenium ion formed during substitution.
Therefore, the combination of steric accessibility and resonance stabilization leads to para substitution being favored in many EAS reactions.
Electrophilic substitution is a broad class of reactions where an electrophile replaces a substituent, usually hydrogen, on an organic molecule.
Electrophilic aromatic substitution (EAS) is a specific type of electrophilic substitution that occurs on aromatic rings, such as benzene or heteroaromatic compounds. The defining feature of EAS is that the aromatic system’s aromaticity is preserved after substitution. The mechanism involves formation of a resonance-stabilized arenium ion intermediate and the eventual restoration of the aromatic π-electron system.
In contrast, electrophilic substitution on non-aromatic compounds typically does not involve maintaining aromaticity and may proceed through different mechanisms without such stabilized intermediates.
Thus, EAS is a subset of electrophilic substitution reactions distinguished by its occurrence on aromatic systems and preservation of aromaticity.
Between anisole and ethylbenzene, anisole is more reactive in electrophilic aromatic substitution due to the strong electron-donating effect of its methoxy (-OCH₃) group.
The methoxy group in anisole donates electron density into the aromatic ring via resonance. This resonance donation increases the electron density primarily at the ortho and para positions of the ring, making these sites more nucleophilic and therefore more reactive toward electrophiles. As a result, anisole undergoes EAS reactions more rapidly and under milder conditions than benzene or ethylbenzene.
Ethylbenzene contains an ethyl (-CH₂CH₃) substituent, which is an electron-donating group as well but only through the inductive effect and hyperconjugation, both of which are much weaker than the resonance effect from the methoxy group. Hence, ethylbenzene’s ring is less activated toward electrophilic attack compared to anisole.
Practically, anisole can be nitrated or brominated more quickly and often yields predominantly ortho and para-substituted products, whereas ethylbenzene reacts more slowly and with less regioselectivity.
Benzene undergoes electrophilic substitution rather than addition primarily because of its aromatic stability.
Benzene’s six π-electrons are delocalized in a planar, cyclic structure, giving it exceptional stability, known as aromaticity. Any reaction that disrupts this conjugation and aromatic sextet requires significant energy.
In electrophilic substitution, the aromatic ring temporarily loses aromaticity upon electrophile attack but quickly regains it after the proton is removed. This substitution preserves the aromatic π-system, maintaining the ring’s overall stability.
In contrast, an electrophilic addition reaction breaks the π-bond permanently, converting the aromatic system into a non-aromatic cyclohexadiene or similar species, which is energetically unfavorable due to loss of aromatic stabilization.
Thus, benzene prefers substitution reactions that replace a hydrogen with an electrophile while maintaining aromaticity, rather than addition reactions that disrupt the conjugated π-system and lose aromatic stability.
Adding a nitro group (-NO₂) to benzene involves the nitration of benzene, a classic example of electrophilic aromatic substitution.
The procedure involves:
- Generating the Electrophile: Mix concentrated nitric acid (HNO₃) and concentrated sulfuric acid (H₂SO₄). Sulfuric acid protonates nitric acid, producing the nitronium ion (NO₂⁺), which is the actual electrophile that attacks the benzene ring. HNO3 +H2 SO4 → NO2+ +HSO4- + H2O
- Electrophilic Attack: The electron-rich benzene ring donates a pair of electrons to the nitronium ion, forming a resonance-stabilized arenium ion intermediate.
- Deprotonation: A base (often HSO₄⁻) removes a proton from the arenium ion, restoring aromaticity and yielding nitrobenzene.
This reaction is typically conducted at controlled temperatures (around 50°C) to avoid over-nitration or side reactions.
Key practical tips:
- Use freshly prepared acid mixtures for maximum electrophile generation.
- Control temperature to favor monosubstitution.
- Nitrobenzene is an important intermediate in dyes, pharmaceuticals, and explosives.
This method is highly efficient and widely used in organic synthesis to introduce the nitro functional group onto aromatic systems.
An electrophile is a species that is electron-deficient and positively charged or neutral but with an electron-poor center, meaning it seeks electrons and acts as an electron pair acceptor in chemical reactions.
In electrophilic aromatic substitution (EAS), electrophiles can be positively charged ions like the nitronium ion (NO₂⁺), carbocations, or neutral molecules with polarized bonds that render certain atoms electron-poor. For example, in halogenation, the halogen molecule (Br₂ or Cl₂) becomes polarized in the presence of a Lewis acid catalyst, creating a partial positive charge on one halogen atom, enabling it to act as an electrophile.
Because electrophiles are attracted to electron-rich sites, such as the π-electrons in an aromatic ring, their positive or partial positive nature allows them to interact with the nucleophilic aromatic system to form new bonds.
In summary, electrophiles are always electron-poor and either carry a positive charge or possess a region of partial positive charge, making them the reactive centers in substitution reactions with aromatic rings.
The theory of electrophilic aromatic substitution (EAS) is grounded in understanding the unique stability and reactivity of aromatic compounds due to their conjugated π-electron systems.
Aromatic rings like benzene are exceptionally stable because of aromaticity—a continuous, cyclic, conjugated system of π-electrons following Hückel’s rule (4n+2 π electrons). This delocalization lowers the energy of the system and makes the ring less reactive toward many typical addition reactions that would break aromaticity.
EAS theory explains that aromatic rings undergo substitution reactions with electrophiles rather than addition to preserve aromatic stability. The key steps are:
- The aromatic ring’s high electron density allows it to act as a nucleophile toward electrophiles.
- The electrophile attacks the π-electron cloud, forming a resonance-stabilized carbocation intermediate known as the arenium ion or sigma complex. This step disrupts aromaticity and is the rate-determining step due to the high energy of the intermediate.
- Loss of a proton from the intermediate restores aromaticity, completing the substitution.
The theory also incorporates substituent effects on reactivity and orientation. Electron-donating groups stabilize the positively charged intermediate and activate the ring, directing substitution to ortho and para positions. Electron-withdrawing groups destabilize the intermediate, deactivate the ring, and direct substitution to the meta position.
This theoretical framework not only explains why EAS occurs but also predicts reaction rates, regioselectivity, and the effects of catalysts, enabling rational design of synthetic routes involving aromatic compounds.
The five classic Electrophilic Aromatic Substitution (EAS) reactions commonly studied and widely used in organic synthesis are:
- Nitration: Introduction of a nitro group (-NO₂) into the aromatic ring, using nitric acid and sulfuric acid to generate the electrophile nitronium ion (NO₂⁺). This reaction is fundamental for producing nitroaromatics, which serve as precursors to amines and explosives.
- Halogenation: Substitution of a hydrogen by a halogen (Cl, Br) using halogen molecules and Lewis acid catalysts (FeCl₃, FeBr₃). Halogenated aromatics are useful intermediates in further cross-coupling reactions and as intermediates in drug synthesis.
- Sulfonation: Introduction of the sulfonic acid group (-SO₃H) by reacting with sulfur trioxide or fuming sulfuric acid (oleum). Aromatic sulfonic acids find applications as detergents, dyes, and ion-exchange resins.
- Friedel-Crafts Alkylation: Alkyl groups are attached to the aromatic ring using alkyl halides and Lewis acid catalysts like AlCl₃. This reaction allows the synthesis of alkylated aromatic compounds, commonly found in pharmaceuticals and fragrances.
- Friedel-Crafts Acylation: Acyl groups (RCO-) are introduced using acyl chlorides and AlCl₃. The resulting aromatic ketones are important intermediates in pharmaceuticals, perfumes, and polymers.
These five reactions form the foundational toolkit for modifying aromatic rings in a controlled manner, each with distinct electrophiles and conditions but sharing the same mechanistic principle of substitution while preserving aromaticity.
Electrophilic Aromatic Substitution (EAS) is a fundamental reaction mechanism widely used in organic chemistry to introduce various substituents onto an aromatic ring, such as benzene or its derivatives. Aromatic rings are chemically stable due to their conjugated π-electron system, which resists many types of reactions. However, through EAS, chemists can strategically replace a hydrogen atom on the aromatic ring with an electrophile, thereby functionalizing the ring without disrupting its aromaticity.
This reaction is extensively employed in the synthesis of pharmaceuticals, dyes, agrochemicals, and polymers. For example, nitration of benzene via EAS introduces a nitro group (-NO₂), which can later be reduced to an amine, an essential intermediate for making aniline-based drugs or dyes. Similarly, sulfonation allows the introduction of sulfonic acid groups, used to make detergents and dyes. Halogenation, another EAS variant, helps incorporate halogen atoms (Cl, Br), enabling further substitution or coupling reactions useful in medicinal chemistry.
In industrial organic synthesis, EAS reactions are crucial for modifying aromatic scaffolds to tailor their chemical and physical properties, such as altering solubility, reactivity, or biological activity. Additionally, the ability of EAS to proceed under relatively mild conditions and maintain aromaticity makes it a preferred method over other harsher aromatic modification reactions.
In practical terms, mastering EAS techniques allows chemists to build complex molecules step-by-step, starting from simple aromatic compounds. Understanding the reactivity and directing effects of substituents enables precise control over the position and type of substitution, which is essential for designing target molecules in synthetic organic chemistry.
Electrophilic Aromatic Substitution (EAS) is a cornerstone reaction in organic synthesis due to its ability to selectively functionalize aromatic rings without destroying their aromatic stability. This makes it invaluable for constructing complex organic molecules in pharmaceuticals, agrochemicals, dyes, and materials science.
The importance of EAS lies in its versatility. It allows chemists to introduce a variety of functional groups—such as nitro, sulfonyl, halogen, alkyl, and acyl groups—onto aromatic rings, which serve as key intermediates or final functional moieties. These substituents often act as handles for further synthetic transformations, enabling multi-step synthesis pathways.
Moreover, EAS reactions offer control over regioselectivity through the influence of existing substituents on the ring. Activating and deactivating groups direct new substituents to ortho, meta, or para positions, enabling precision in molecular design. This control is critical in drug development, where subtle changes in substitution patterns can drastically alter biological activity.
From an industrial perspective, EAS reactions are conducted on a large scale to produce intermediates for polymers, pigments, and pharmaceuticals efficiently and economically. The mild conditions often used preserve sensitive functional groups elsewhere in the molecule, reducing side reactions.
In summary, EAS is essential in organic synthesis because it provides a practical, predictable, and versatile approach to modify aromatic compounds, facilitating the development of a wide array of chemical products.
Electrophilic Aromatic Substitution (EAS) proceeds through a well-defined multi-step mechanism designed to replace a hydrogen atom on an aromatic ring with an electrophile while preserving aromaticity. The main steps are:
- Generation of the Electrophile: Often, the electrophile (E⁺) is generated in situ or activated by a catalyst. For example, nitration requires the formation of the nitronium ion (NO₂⁺) from a mixture of nitric and sulfuric acids. Halogenation uses a Lewis acid catalyst like FeBr₃ to activate Br₂.
- Electrophilic Attack and Formation of the Arenium Ion: The electrophile attacks the aromatic π-electron cloud, forming a sigma complex called the arenium ion. This intermediate is resonance-stabilized but temporarily loses aromaticity, making it higher in energy. This step is slow and the rate-determining step of the reaction.
- Deprotonation and Restoration of Aromaticity: A base (often the counterion or solvent) removes a proton from the arenium ion intermediate, restoring the aromatic system. This step is fast and drives the reaction toward completion.
- Product Formation: The substituted aromatic compound is formed, where the electrophile replaces a hydrogen on the ring.
These steps are general for most EAS reactions, including nitration, sulfonation, halogenation, alkylation, and acylation. Understanding this mechanism allows chemists to predict the effects of substituents, optimize reaction conditions, and control regioselectivity for desired synthetic outcomes.
Anisole is more reactive than ethylbenzene toward electrophilic aromatic substitution (EAS) due to the difference in the nature and strength of their substituents attached to the benzene ring.
Anisole contains a methoxy (-OCH₃) group, which is a strong electron-donating substituent via resonance. The oxygen atom donates lone pair electrons into the aromatic ring, increasing the electron density predominantly at the ortho and para positions. This elevated electron density makes the ring more nucleophilic and reactive toward electrophiles, significantly accelerating EAS reactions. The resonance effect also stabilizes the arenium ion intermediate formed during substitution, reducing the activation energy and speeding up the reaction.
Ethylbenzene, on the other hand, has an ethyl (-CH₂CH₃) group, which is an electron-donating group but primarily through the inductive effect, which is weaker than resonance donation. Alkyl groups donate electron density via hyperconjugation and inductive effects, but these are less effective at increasing the ring’s nucleophilicity compared to oxygen’s lone pair resonance donation in anisole.
Practically, anisole undergoes substitution reactions faster and under milder conditions compared to ethylbenzene. For instance, anisole nitrates more readily and yields predominantly ortho and para products, while ethylbenzene reacts more slowly with less regioselectivity.
Therefore, the strong resonance donation from the methoxy group in anisole makes it significantly more reactive toward EAS than the alkyl-substituted ethylbenzene.
Between pyrrole and pyridine, pyrrole undergoes electrophilic aromatic substitution (EAS) much faster than pyridine.
Pyrrole is a five-membered heterocyclic aromatic compound containing a nitrogen atom with a lone pair of electrons that is part of the aromatic sextet. This lone pair is delocalized into the ring, significantly increasing the overall electron density of the ring and making it highly nucleophilic. Due to this enhanced electron density, pyrrole reacts rapidly with electrophiles, often under mild conditions, to give substitution products primarily at the C-2 position.
In contrast, pyridine is a six-membered heteroaromatic ring with a nitrogen atom whose lone pair is in an sp² orbital orthogonal to the π-system, meaning it does not contribute to the aromatic sextet. Moreover, the nitrogen atom in pyridine is electronegative and withdraws electron density inductively from the ring, making the ring less nucleophilic. As a result, pyridine is much less reactive toward electrophiles in EAS reactions. Instead, it often undergoes nucleophilic substitution at the carbon atoms ortho and para to nitrogen.
Thus, the electron-rich pyrrole ring is significantly more reactive toward electrophilic aromatic substitution than the electron-poor pyridine ring. This difference is fundamental in heterocyclic chemistry and has practical implications in designing synthesis routes for nitrogen-containing aromatic compounds.
Solvents play a critical role in influencing the reaction rates of electrophilic aromatic substitution (EAS) by affecting the solubility, stability, and reactivity of the reactants, intermediates, and catalysts involved.
Polar protic solvents like acetic acid often accelerate EAS reactions because they can stabilize charged intermediates and transition states through hydrogen bonding and dipole interactions. Acetic acid, with its ability to donate and accept hydrogen bonds, helps stabilize the positively charged arenium ion intermediate, lowering the activation energy and increasing the reaction rate. Additionally, acetic acid can solubilize both electrophiles and aromatic substrates, enhancing molecular collisions and thus reaction efficiency.
In contrast, non-polar solvents like cyclohexane provide a less interactive environment. Cyclohexane is a non-polar, aprotic solvent that does not stabilize charged species well. As a result, reactions proceeding through charged intermediates, such as the arenium ion in EAS, tend to be slower in cyclohexane. The reduced solubility of polar or ionic species in cyclohexane can also limit reaction rate by decreasing the effective concentration of reactants in the reaction medium.
Moreover, the choice of solvent influences the selectivity of substitution, the stability of electrophiles, and even the catalyst’s effectiveness. For example, in nitration reactions, a polar acidic solvent mixture is often used to generate the nitronium ion electrophile more efficiently.
In summary, polar protic solvents like acetic acid generally enhance the rate of EAS by stabilizing intermediates and increasing reactant solubility, while non-polar solvents like cyclohexane tend to slow down these reactions due to poor stabilization and solubility effects.
Benzene and aromatic compounds readily undergo electrophilic aromatic substitution (EAS) but are generally unreactive towards nucleophilic aromatic substitution (NAS) due to differences in their electronic structures and reaction mechanisms.
In EAS, the electron-rich aromatic ring acts as a nucleophile, attacking electrophiles. The π-electron cloud of benzene is delocalized and relatively high in electron density, which facilitates interaction with electrophiles. The reaction proceeds via a resonance-stabilized arenium ion intermediate, and importantly, the aromatic system is restored after substitution, preserving the system’s stability. This retention of aromaticity drives the reaction forward and allows substitution to occur under relatively mild conditions.
In contrast, nucleophilic aromatic substitution requires the aromatic ring to be electrophilic enough to be attacked by nucleophiles. However, benzene’s electron-rich nature and stable conjugation make it resistant to nucleophilic attack. For NAS to occur, the aromatic ring typically must bear strong electron-withdrawing groups (like nitro groups) that stabilize the negatively charged intermediates formed during the reaction, such as Meisenheimer complexes. These groups increase the ring’s susceptibility to nucleophiles by lowering electron density.
Furthermore, NAS mechanisms generally require either a good leaving group and electron-withdrawing substituents or alternative conditions like benzyne intermediates, which are less common and require harsher conditions.
Therefore, the inherent electron density and aromatic stabilization of benzene favor electrophilic substitution and render nucleophilic substitution difficult without additional activating factors.
The rate-determining step (RDS) in electrophilic aromatic substitution (EAS) is the formation of the sigma complex, also known as the arenium ion or the sigma complex intermediate.
During EAS, the electrophile first approaches the aromatic ring. Because the aromatic ring is electron-rich, the electrophile attacks one of the ring carbons, temporarily disrupting aromaticity. This step involves the electrophile forming a sigma bond with a carbon atom of the aromatic ring, resulting in a resonance-stabilized carbocation intermediate (the arenium ion). This intermediate is less stable than the aromatic starting material because it has lost aromatic stabilization, and its formation requires overcoming a significant energy barrier.
The formation of this arenium ion is slow and energy-demanding, making it the rate-limiting step of the entire substitution process. Once this intermediate forms, the subsequent step—loss of a proton (deprotonation) to restore aromaticity—is much faster and energetically favorable.
The significance of this rate-determining step is that it governs the reaction kinetics and can be influenced by factors such as the nature of the electrophile, substituents on the ring, and reaction conditions. Electron-donating groups on the aromatic ring stabilize the arenium ion intermediate via resonance, thereby lowering the activation energy and increasing the reaction rate. Conversely, electron-withdrawing groups destabilize the intermediate and slow the reaction.
Understanding that the rate-determining step involves breaking aromaticity to form the arenium ion provides insight into why aromatic substitution requires specific catalysts or activating groups to proceed efficiently.
The rate determining step (RDS) in electrophilic aromatic substitution (EAS) is the formation of the sigma complex, also known as the arenium ion or the sigma complex intermediate.
During EAS, the electrophile first approaches the aromatic ring. Because the aromatic ring is electron-rich, the electrophile attacks one of the ring carbons, temporarily disrupting aromaticity. This step involves the electrophile forming a sigma bond with a carbon atom of the aromatic ring, resulting in a resonance-stabilized carbocation intermediate (the arenium ion). This intermediate is less stable than the aromatic starting material because it has lost aromatic stabilization, and its formation requires overcoming a significant energy barrier.
The formation of this arenium ion is slow and energy-demanding, making it the rate-limiting step of the entire substitution process. Once this intermediate forms, the subsequent step—loss of a proton (deprotonation) to restore aromaticity—is much faster and energetically favorable.
The significance of this rate-determining step is that it governs the reaction kinetics and can be influenced by factors such as the nature of the electrophile, substituents on the ring, and reaction conditions. Electron-donating groups on the aromatic ring stabilize the arenium ion intermediate via resonance, thereby lowering the activation energy and increasing the reaction rate. Conversely, electron-withdrawing groups destabilize the intermediate and slow the reaction.
Understanding that the rate-determining step involves breaking aromaticity to form the arenium ion provides insight into why aromatic substitution requires specific catalysts or activating groups to proceed efficiently.
When comparing the reactivity of chlorobenzene and cyclopentadiene toward electrophilic aromatic substitution, cyclopentadiene is generally much more reactive.
Cyclopentadiene is a five-membered conjugated diene, and although it is not aromatic in the same way as benzene, it exhibits higher reactivity toward electrophilic substitution due to its electron-rich diene system. The conjugated double bonds provide a high electron density that readily interacts with electrophiles, facilitating faster reactions. Cyclopentadiene can also undergo Diels-Alder and other addition reactions due to its diene nature, contributing to its overall higher reactivity.
On the other hand, chlorobenzene is a substituted benzene where the chlorine atom exerts both inductive and resonance effects. Chlorine is an electron-withdrawing group by induction (due to its electronegativity), which tends to reduce the electron density of the aromatic ring, thereby making the ring less reactive toward electrophiles compared to benzene itself. However, chlorine also has lone pairs that can participate in resonance, donating electron density into the ring at ortho and para positions, but this effect is weaker than its inductive withdrawal. Overall, chlorobenzene is less reactive than benzene toward EAS, and significantly less reactive than cyclopentadiene.
Thus, cyclopentadiene’s non-aromatic but electron-rich conjugated diene system makes it more nucleophilic and reactive toward electrophiles compared to the electron-deficient chlorobenzene aromatic ring.
Electrophilic Aromatic Substitution (EAS) and Electrophilic Addition are two fundamentally different reaction types involving electrophiles, and their distinction is crucial in organic chemistry.
EAS occurs on aromatic rings, such as benzene, where the aromatic system remains intact after the reaction. In EAS, an electrophile replaces one hydrogen atom on the aromatic ring without destroying the aromatic π-electron conjugation. This preservation of aromaticity is essential because the aromatic system's stability provides the driving force for the reaction's specificity and selectivity. The general mechanism involves the formation of a resonance-stabilized arenium ion intermediate, followed by the loss of a proton to restore aromaticity.
In contrast, electrophilic addition typically occurs on alkenes or alkynes, which are non-aromatic unsaturated compounds. During electrophilic addition, the electrophile adds across the double or triple bond, breaking the π-bond and converting the molecule into a saturated or partially saturated compound. This reaction does not preserve any special electronic structure like aromaticity but rather transforms the unsaturated bond into a single or double bond. For example, bromination of ethene adds bromine atoms across the C=C double bond, converting it into a dibromoalkane.
The key difference lies in the outcome: EAS maintains aromaticity by substituting, while electrophilic addition destroys the unsaturation by adding electrophiles directly. This difference also influences reaction conditions and reactivity. Aromatic rings resist addition reactions due to the loss of aromatic stabilization, whereas alkenes readily undergo addition reactions because they lack such stabilization.
Understanding this distinction helps chemists predict reaction pathways and design synthesis routes accordingly, selecting conditions that favor substitution on aromatic rings or addition on unsaturated non-aromatic systems.