Fact-checked by Grok 2 weeks ago

Aromatic compound

Aromatic compounds are cyclically conjugated molecular entities possessing a significantly greater than that of a hypothetical localized , such as a Kekulé , owing to electron delocalization. In the traditional sense, they exhibit a chemistry typified by , the prototypical aromatic with the molecular formula C₆H₆. Originally named for the often pleasant odors of early-discovered examples like and , the term "aromatic" now specifically denotes this class of highly stable, resonance-stabilized organic molecules featuring planar rings with delocalized π electrons./Arenes/Properties_of_Arenes/Aromaticity/What_does_aromatic_really_mean) Benzene was first isolated in 1825 by from the oily condensate produced during the compression of illuminating gas derived from . For decades, its structure puzzled chemists due to its high —empirical formula suggesting three degrees—yet resistance to typical reactions of alkenes like addition. In 1865, proposed the iconic cyclic structure: a regular of six carbon atoms, each bonded to one hydrogen, with three alternating double bonds representing the six π electrons. This model, inspired by Kekulé's dream of a snake biting its tail (the ), accounted for benzene's symmetry and stability but was later refined by the concept of , where the double bonds are delocalized in a continuous π cloud above and below the ring plane. The defining criteria for aromaticity in monocyclic systems, established in the 1930s by Erich Hückel, require a to be cyclic, planar, fully conjugated (with continuous p-orbital overlap), and to contain 4n + 2 π electrons, where n is a non-negative integer (). This electron count—such as 6 for (n=1), 10 for the cyclooctatetraene dianion (n=2), or 2 for the cyclopropenyl cation (n=0)—results in a closed-shell, or similar configuration that confers exceptional thermodynamic stability, often 30–40 kcal/mol greater than expected for localized bonds. Systems with 4n π electrons (e.g., cyclobutadiene with 4) are antiaromatic and destabilized, while those failing other criteria are nonaromatic./Arenes/Properties_of_Arenes/Aromaticity) Experimental evidence for aromaticity includes diatropic ring currents observed in NMR spectroscopy, where protons inside the ring are shielded and appear upfield. Beyond benzene, aromatic compounds encompass a vast array, including polycyclic aromatic hydrocarbons (PAHs) like (C₁₀H₈, two fused rings with 10 π electrons) and , which follow extended versions of and are ubiquitous in fossil fuels and combustion products. Heterocyclic aromatics, where one or more ring atoms are heteroatoms (e.g., nitrogen in , oxygen in , sulfur in ), also satisfy aromatic criteria and are essential in biochemistry; for instance, mimics benzene's reactivity but with basic nitrogen, while the rings in feature four units. Charged species like the (C₇H₇⁺, 6 π electrons) and (neutral, 6 π electrons) exemplify non-benzenoid aromatics. Aromatic compounds exhibit distinctive reactivity, favoring (e.g., , ) over addition, preserving the aromatic π system, due to the high for disrupting delocalization./01:_Chapters/1.30:_Aromatic_Compounds) Their stability and tunable electronics make them foundational in , serving as building blocks for pharmaceuticals (e.g., aspirin, antibiotics), dyes (e.g., azo compounds), polymers (e.g., ), and like and fullerenes./15:_Benzene_and_Aromaticity) However, some aromatics, particularly PAHs, pose environmental and health risks as carcinogens./09:_Organic_Chemistry/9.03:_Aromatic_Compounds-_Benzene_and_Its_Relatives)

Definition and Criteria

Classical Definition

The classical definition of aromatic compounds emerged from empirical observations of their distinctive chemical behavior in the , particularly their exceptional stability relative to alkenes and preference for over addition reactions. These compounds, unlike typical unsaturated hydrocarbons, resist decolorization by and do not readily undergo oxidation or under mild conditions that affect double bonds in aliphatic systems. This low reactivity toward addition preserves the integrity of the molecular framework, while substitution reactions, such as or , occur readily at the ring positions. The term "aromatic" originated in the early 19th century from the pleasant, sweet odor associated with and related derivatives, such as those isolated from fragrant resins like gum benzoin. , first isolated in 1825 by from whale oil, exemplified these traits through its resistance to oxidation—unlike , which oxidizes easily with —highlighting the unusual thermodynamic stability of such systems. This stability was quantified empirically through studies, revealing a energy of approximately 150 kJ/mol for compared to hypothetical localized structures, underscoring the delocalized nature contributing to aromatic character. A pivotal advancement in classical understanding came in 1865 when proposed a cyclic, hexagonal structure for , consisting of six carbon atoms alternately linked by single and double bonds, to account for its saturation-equivalent formula C₆H₆ and observed stability. This model established the foundational criteria for aromatic compounds as planar, cyclic systems with continuous conjugation, accounting for its empirical formula and observed stability through alternating single and double bonds.

Hückel's Rule and Aromaticity

, proposed by Erich Hückel in 1931, provides a quantum mechanical criterion for determining the of planar, monocyclic, conjugated polyenes. According to this rule, such a system is aromatic if it contains $4n + 2 π electrons, where n is a non-negative (0, 1, 2, ...), leading to exceptional stability due to a closed-shell configuration of molecular orbitals. Conversely, systems with $4n π electrons are antiaromatic, exhibiting destabilization and reactivity. The rule arises from Hückel molecular orbital (HMO) theory, which approximates the π-electron energies in cyclic conjugated systems using a simple secular determinant. For a of m atoms, the eigenvalues are given by \alpha + 2\beta \cos(2\pi j / m) for j = 0, 1, ..., m-1, where \alpha is the coulomb integral and \beta the resonance integral (negative). Systems satisfying $4n + 2 fill all bonding orbitals below the non-bonding level, maximizing delocalization energy. A graphical representation of these energy levels is provided by the Frost circle method, developed by Arthur A. Frost and Boris Musulin in 1953. To construct a Frost circle, inscribe a regular polygon with one vertex at the bottom in a circle whose diameter equals the energy difference between bonding and antibonding orbitals; the vertices represent the π molecular orbital energies, with the lowest point as the lowest-energy orbital. For benzene (m=6), the hexagon has the lowest vertex representing the fully bonding MO at \alpha + 2\beta, the next two vertices as a degenerate pair of bonding MOs at \alpha + \beta, the upper two vertices as a degenerate pair of antibonding MOs at \alpha - \beta, and the top vertex as the highest antibonding MO at \alpha - 2\beta. The six π electrons occupy the three bonding MOs below the midline (\alpha), confirming aromaticity with 6 π electrons (n=1). Illustrative examples include the cyclopropenyl cation, a three-membered ring with 2 π electrons (n=0), which is aromatic and stable in appropriate conditions due to its filled bonding orbital. In contrast, cyclobutadiene, with 4 π electrons (n=1), is antiaromatic, possessing two singly occupied non-bonding orbitals that result in high reactivity and a rectangular from planarity. The extends to larger annulenes, unbranched cyclic polyenes, predicting stability for those with $4n + 2 π electrons, such as annulene (14 electrons, n=3), which exhibits aromatic character despite conformational challenges. Hückel's analysis showed that larger annulenes follow the same electron-counting pattern, with delocalization energy increasing for aromatic cases but decreasing for antiaromatic ones.

Historical Development

Early Observations

In 1825, isolated from the oily residue produced during the compression of illuminating gas derived from , naming it "bicarburet of hydrogen" based on its empirical composition. This marked the first laboratory isolation of the compound, though it occurred naturally in and other sources. Faraday's work highlighted benzene's volatile, colorless liquid properties, setting the stage for further investigations into its structure. By 1834, Eilhard Mitscherlich synthesized through the of with lime, confirming its as C₆H₆ via , which revealed a 1:1 carbon-to-hydrogen ratio by mass. This formula puzzled chemists, as it suggested high unsaturation akin to three double bonds, yet exhibited unexpected stability; unlike alkenes, it resisted addition reactions such as decolorization of or reaction with without catalysts, instead undergoing under harsh conditions. In 1855, August Wilhelm von Hofmann coined the term "aromatic" to describe and related compounds like , attributing the name to their distinctive, often pleasant odors, such as 's sweet almond-like scent. The quest for benzene's structure intensified in 1865 when Friedrich August Kekulé proposed a cyclic arrangement of six carbon atoms with alternating single and double bonds, inspired by a daydream of a snake biting its tail, symbolizing a . This model accounted for the C₆H₆ and the compound's stability but faced challenges, as the predicted reactivity of three double bonds did not match observations. In 1869, Albert countered with a prismane structure—a bridged, three-dimensional of six carbons—to explain the same and substitution behavior, sparking debate that underscored the empirical puzzles of aromatic compounds.

Formulation of Aromaticity

In 1899, Johannes Thiele introduced the partial valence theory to explain the delocalized bonding in , proposing that in conjugated systems, adjacent share partial valences that neutralize internally, leaving reactive ends exposed while stabilizing the central structure. This concept addressed the limitations of Kekulé's alternating single and model by accounting for 's unexpected and reactivity patterns without invoking full bond alternation. The 1920s and 1930s marked a shift toward quantum mechanical frameworks, with developing resonance theory to describe aromatic compounds as hybrids of multiple Lewis structures, enhancing stability through delocalization in and related systems. Concurrently, Erich Hückel applied in 1931, calculating the π-electron energies of and predicting exceptional stability for cyclic, planar, conjugated systems with 4n+2 π electrons, laying the foundation for quantitative assessments of . These approaches complemented each other, with resonance emphasizing valence bond hybridization and Hückel's method highlighting orbital symmetry and energy lowering due to closed-shell configurations. In 1972, Erich Clar extended these ideas to polycyclic aromatic hydrocarbons through his sextet rule, positing that stability arises from localized aromatic sextets—six π electrons in benzene-like rings—rather than full delocalization across the entire molecule, as evidenced by UV patterns in compounds like . This empirical guideline prioritized "migrating" sextets in larger systems to predict reactivity and aromatic character more accurately than uniform delocalization models. A pivotal experimental confirmation came in the early 1950s with high-resolution (NMR) , which revealed equivalent proton environments in , directly supporting π-electron delocalization and ring current effects over localized structures. This milestone validated theoretical predictions and shifted focus toward spectroscopic probes of . Ongoing debates in the mid-20th century centered on extending definitions beyond planar, neutral hydrocarbons to include non-planar conformations and charged species, questioning whether criteria like apply universally or require modifications for systems exhibiting partial delocalization or conformational flexibility. These discussions highlighted tensions between structural planarity and electronic criteria, influencing refinements in quantum models for diverse aromatic analogs.

Structure and Bonding

Benzene as Prototype

, with the molecular formula C₆H₆, exemplifies the archetypal aromatic compound through its distinctive planar hexagonal ring structure, where six carbon atoms form the ring and each is bonded to a . This arrangement results in a fully with all internal bond angles measuring exactly 120 degrees, consistent with sp² hybridization of the carbon atoms. The six C-C bonds are identical in length at 1.39 , a value intermediate between the typical C-C length of 1.54 and C=C length of 1.34 , reflecting the partial double-bond character due to electron delocalization rather than alternating single and double bonds as in Kekulé's original proposal. The molecule exhibits D_{6h} symmetry, featuring a principal C_6 rotation axis passing through the ring center, six perpendicular C_2 axes, multiple vertical and horizontal mirror planes, and an inversion center, which underscores the equivalence of all six carbon positions. This high degree of not only enforces the planarity of the ring but also promotes uniform electron distribution, enhancing the overall molecular stability by minimizing steric strain and maximizing orbital overlap. The pi electron system arises from the sideways overlap of six unhybridized p_z orbitals—one per carbon—perpendicular to the ring plane, creating a cloud that encompasses the entire ring above and below the plane, housing six pi electrons in a stable, closed-shell configuration. Quantitatively, benzene's aromatic character imparts an stabilization energy of approximately 36 kcal/mol relative to a hypothetical cyclohexatriene with localized alternating bonds, as estimated through comparisons with acyclic polyenes like 1,3,5-hexatriene using block-localized wavefunction methods. This energy gain arises from the delocalized and is a key metric of 's enhanced thermodynamic stability. In disubstituted derivatives, positions are classified relative to the reference group at position 1: for adjacent carbons (positions 2 and 6), for positions separated by one carbon (3 and 5), and for the opposite carbon (4), that highlights the symmetry-equivalent sites and influences subsequent reactivity patterns.

Resonance and Orbital Models

The concept, developed by in the 1930s as part of , describes as a superposition of two equivalent Kekulé structures rather than a fixed single form with alternating single and double s. In this model, the actual electronic structure is a weighted of these forms, resulting in delocalized π electrons that distribute bond orders more evenly across the ring. This resonance delocalization imparts significant stabilization, with the energy of the hybrid being lower than that of either contributing structure by an amount known as the resonance energy. Molecular orbital theory provides an alternative framework for understanding electron delocalization in aromatic systems, with Erich Hückel's 1931 linear combination of atomic orbitals (LCAO) method applied specifically to π electrons in planar conjugated hydrocarbons like . Hückel's approach constructs s from the p_z atomic orbitals of the carbon atoms, yielding six π s for : a lowest-energy fully bonding orbital, two degenerate bonding orbitals, two degenerate antibonding orbitals, and a highest-energy fully antibonding orbital. The six π electrons fill the three bonding orbitals, making the highest occupied (HOMO) the degenerate pair at energy \alpha + \beta and the lowest unoccupied (LUMO) the degenerate antibonding pair at \alpha - \beta, where \alpha is the coulomb integral and \beta is the negative resonance integral. The total π-electron energy is calculated as $6\alpha + 8\beta, compared to $6\alpha + 6\beta for three isolated double bonds, yielding a delocalization energy of $2\beta. E_{\text{total}}^{\pi} = 6\alpha + 8\beta E_d = (6\alpha + 8\beta) - (6\alpha + 6\beta) = 2\beta Valence bond theory and molecular orbital theory both account for the enhanced stability of aromatic compounds but differ in emphasis: the former focuses on resonance among localized Kekulé-like structures to explain bond equalization and energy lowering, while the latter reveals global cyclic conjugation through delocalized orbitals that encircle the ring. These complementary perspectives emerged from the theoretical struggles of the 1930s, with valence bond providing intuitive structural hybrids and molecular orbital offering quantitative predictions of orbital symmetries and energies. For polycyclic aromatic hydrocarbons, Erich Clar's aromatic sextet rule, formulated in 1972, extends these ideas by prioritizing representations that maximize the number of independent benzene-like π-sextet units (six π electrons in a cyclic, delocalized ) while minimizing fixed double bonds. This empirical guideline, drawn from extensive spectroscopic and synthetic studies, identifies the most stable form by placing "Clar sextets" in disjoint rings, thereby rationalizing the distribution of electron pairs and predicting reactivity patterns in fused systems like and .

Classification of Aromatic Compounds

Carbocyclic Arenes

Carbocyclic arenes, also known as all-carbon aromatic hydrocarbons, consist of rings composed exclusively of carbon atoms exhibiting aromatic through delocalized π electrons. These compounds range from simple monocyclic structures like to complex fused polycyclic systems, where arises from adherence to criteria such as planarity and the Hückel rule for π electron count. Unlike heterocyclic , carbocyclic arenes lack heteroatoms, allowing pure carbon frameworks to demonstrate varying degrees of delocalization and . Monocyclic carbocyclic arenes include , the prototypical aromatic compound with a six-membered ring and six π electrons, which achieves exceptional stability through equal bond lengths and delocalized electrons in a planar conformation. Larger s, such as annulene, feature a 14-membered ring with 14 π electrons, satisfying the 4n+2 rule (n=3) and displaying aromatic properties including diatropicity in NMR spectra and bond equalization, though steric strain can challenge planarity in some derivatives. These systems highlight how increasing ring size can maintain if conjugation and electron count align with theoretical predictions. Polycyclic carbocyclic arenes arise from fused benzene rings, sharing two carbon atoms per fusion site, leading to extended π systems. , the simplest such compound, comprises two fused six-membered rings with 10 π electrons delocalized across the structure, rendering it aromatic overall despite unequal bond lengths between rings. exemplifies linear fusion, where three rings align in a straight chain, resulting in 14 π electrons but with reactivity concentrated at the central ring due to less effective delocalization compared to angular fusions like . Fusion patterns influence properties: linear arrangements often exhibit higher reactivity at terminal positions, while angular fusions enhance overall stability through better π overlap. Non-benzenoid carbocyclic arenes deviate from the six-membered motif yet pursue . , an eight-membered with eight π electrons, adopts a non-planar tub-shaped conformation to avoid , rendering it non-aromatic with localized double bonds and alternating single-double bond lengths. , a seven-membered with six π electrons in its , is debated for ; while it satisfies the 4n+2 rule (n=1) and shows some delocalization, its and reactivity suggest partial rather than full aromatic character. In fused systems, stability trends favor peripheral π electrons over internal ones, as articulated by Clar's rule, which posits that the most stable resonance structure maximizes disjoint aromatic sextets (6 π electron units) on the periphery, minimizing strain and enhancing delocalization. This peripheral prioritization explains why linearly fused systems like are less stable than angular counterparts, with internal bonds showing partial double-bond character. Representative large polycyclic aromatic hydrocarbons (PAHs) include , a five-ring angular system with 20 π electrons, noted for its planarity and use in dyes due to extended conjugation, and , a seven-ring circular PAH with 24 π electrons, exhibiting high symmetry and exceptional thermal stability akin to fragments.

Heterocyclic Arenes

Heterocyclic arenes represent a class of aromatic compounds in which one or more ring atoms are heteroatoms, typically , oxygen, or , replacing carbons in otherwise carbocyclic structures like . These heteroatoms introduce unique electronic perturbations, altering the distribution of pi electrons and influencing stability, reactivity, and properties compared to all-carbon analogs. The presence of heteroatoms can either enhance or disrupt depending on how their lone pairs interact with the , as assessed through adaptations of Hückel's 4n+2 rule. Prominent single-ring examples include , , , and , each satisfying the 4n+2 pi electron criterion (n=1, 6 electrons) for . , a six-membered ring with one atom, maintains aromaticity akin to , with the 's occupying an sp² hybrid orbital in the plane of the ring and not contributing to the delocalized ; this results in electron-withdrawing effects, rendering less nucleophilic than . In contrast, and feature five-membered rings with oxygen and sulfur heteroatoms, respectively, where one from the heteroatom resides in a p orbital, donating two electrons to achieve the required six pi electrons for ; these compounds exhibit electron-donating character from the heteroatoms, leading to higher reactivity toward electrophiles. , also five-membered with , follows a similar pattern, as its N-H contributes to the , fulfilling and conferring aromatic stability, though the N-H bond imparts slight electron-richness. Fused heterocyclic systems extend these principles to polycyclic frameworks, combining heterocyclic and carbocyclic rings while preserving overall aromaticity. Quinoline consists of a ring fused to a ring, possessing 10 pi electrons across the bicyclic structure that conform to for each constituent aromatic sextet, with the nitrogen exerting electron-withdrawing influence primarily in its pyridine portion. Indole, formed by fused to a ring, similarly achieves aromaticity through 10 pi electrons, where the pyrrole-like nitrogen donates its lone pair to the five-membered ring's , stabilizing the entire molecule and highlighting the heteroatom's role in electron donation. These fused arenes demonstrate how heteroatoms can modulate electronic density in extended systems without compromising the 4n+2 electron count..pdf) Cases of instability arise when heteroatoms lead to deviations from aromatic criteria, such as in smaller rings. Azete, a four-membered ring containing one atom, contains four pi electrons, violating Hückel's 4n+2 rule and rendering it ; this electron count results in high strain and reactivity, with the molecule prone to dimerization or ring-opening due to destabilizing cyclic conjugation. Such examples underscore the stringent requirements for in heterocyclic systems, where heteroatom incorporation can tip the balance toward if pi electron parity is unfavorable.

Physical and Spectroscopic Properties

Physical Characteristics

Aromatic compounds display distinctive macroscopic physical properties shaped by their planar, conjugated structures. Boiling and melting points vary with molecular size and ring fusion, often higher than those of comparable aliphatic hydrocarbons due to stronger intermolecular forces from pi-electron delocalization. , the prototypical aromatic compound, boils at 80.1 °C and melts at 5.5 °C. In contrast, the fused-ring system of exhibits elevated values of 218 °C for and 80.3 °C for melting point, reflecting enhanced stability from pi-stacking between aromatic rings. Solubility profiles underscore the nonpolar character of most aromatic compounds, rendering them poorly soluble in but highly miscible with solvents. , for example, dissolves to only about 1.8 g/L in at 25 °C, yet it readily mixes with nonpolar solvents like or . Substituents can modulate polarity; the nitro group in increases to approximately 1.9 g/L at 25 °C by enhancing moments, though it remains limited compared to fully polar molecules. Density and are notably influenced by the rigid aromatic framework. Aromatic hydrocarbons typically exhibit higher than aliphatic analogs of similar ; benzene's is 0.876 g/cm³ at 20 °C, exceeding that of (0.659 g/cm³). follows suit, with benzene at 0.62 mPa·s at 20 °C versus 0.30 mPa·s for , attributable to the compact ring structure impeding molecular flow. Simple aromatic compounds often emit a characteristic odor, evoking the term "aromatic." Toluene, a benzene derivative, possesses a sweet, pungent scent similar to paint thinners. In the solid phase, the planarity of aromatic rings promotes efficient crystal packing via pi-pi interactions, leading to ordered structures and influencing melting behavior.

Spectroscopic Identification

Spectroscopic methods provide definitive evidence for the presence of aromatic systems by detecting characteristic electronic and vibrational transitions arising from delocalized pi-electrons. These techniques exploit the unique symmetry and conjugation in aromatic compounds, allowing differentiation from non-aromatic unsaturated systems like alkenes. Ultraviolet-visible (UV-Vis), nuclear magnetic resonance (NMR), infrared (IR), and mass spectrometry (MS) are the primary tools, each offering complementary structural insights. In UV-Vis spectroscopy, aromatic compounds display intense absorptions due to pi-to-pi* electronic transitions within the . Benzene, as the prototype, exhibits a strong band near 180 nm (ε > 60,000 M⁻¹ cm⁻¹), a medium-intensity band at 200 nm (ε ≈ 7,400 M⁻¹ cm⁻¹), and weaker forbidden bands around 254 nm (ε ≈ 230 M⁻¹ cm⁻¹), reflecting the symmetry-imposed restrictions on transitions. Substituted benzenes show bathochromic shifts depending on the substituents, with electron-donating groups intensifying and red-shifting the bands, enabling confirmation of extended conjugation beyond simple alkenes, which absorb at shorter wavelengths (typically <200 nm) with lower molar absorptivities. Nuclear magnetic resonance spectroscopy highlights the diatropic ring current in aromatic rings, which generates a secondary magnetic field that deshields protons in the plane, shifting their signals downfield. Aromatic protons typically resonate between 6.5 and 8.5 ppm, as seen in benzene at 7.27 ppm in CDCl₃, contrasting with alkene protons at 4.6–5.7 ppm. In ¹³C NMR, aromatic carbons appear in the 110–150 ppm range due to sp² hybridization and anisotropic effects, with unsubstituted benzene showing a single peak at 128.4 ppm. Integration and multiplicity patterns further distinguish monosubstituted versus polysubstituted rings, providing evidence of aromatic symmetry. Infrared spectroscopy identifies aromatic functional groups through specific vibrational modes. Aromatic C–H stretching occurs above 3000 cm⁻¹, typically as sharp bands near 3030 cm⁻¹, distinguishing them from aliphatic C–H stretches below 3000 cm⁻¹. The ring C=C stretching vibrations appear as medium-intensity bands between 1450 and 1600 cm⁻¹, often as a doublet around 1500 and 1580 cm⁻¹ for . Out-of-plane C–H bending modes in the 650–900 cm⁻¹ region serve as fingerprints for substitution patterns: monosubstituted benzenes show strong bands at 690–710 and 730–770 cm⁻¹, while para-disubstituted rings exhibit a characteristic band near 810–840 cm⁻¹. These features contrast with alkene C=C stretches at 1620–1680 cm⁻¹ and their broader C–H bends. Mass spectrometry reveals the stability of aromatic molecular ions, which resist fragmentation due to resonance delocalization. Aromatic compounds often display prominent molecular ion (M⁺) peaks with intensities exceeding 50% of the base peak, as in benzene where the M⁺ at m/z 78 is abundant and stable. Common fragments include tropylium ion (m/z 91) from benzyl cleavage in substituted cases, aiding identification. In contrast, alkenes produce weaker M⁺ peaks and favor allylic cleavages, underscoring the diagnostic value for aromatics.

Chemical Reactivity

Electrophilic Aromatic Substitution

Electrophilic aromatic substitution (EAS) represents the characteristic reaction pathway for aromatic compounds, in which a hydrogen atom on the aromatic ring is replaced by an electrophilic species while the aromatic system's stability is maintained through a substitution rather than addition process. This reactivity stems from the high electron density of the aromatic π-system, which attracts electrophiles, leading to reactions that are typically slower than analogous aliphatic substitutions due to the partial loss of aromaticity in the transition state. The overall transformation can be represented as Ar–H + E⁺ → Ar–E + H⁺, where Ar denotes the aromatic moiety and E⁺ is the electrophile. The mechanism of EAS, established by Christopher K. Ingold and coworkers in the mid-20th century, proceeds via a two-stage addition-elimination sequence. In the first, rate-determining step, the electrophile adds to the aromatic ring, forming a resonance-stabilized carbocation intermediate known as the or σ-complex, in which the aromaticity is temporarily disrupted as the ring adopts a sp³-hybridized carbon at the substitution site. This intermediate features delocalized positive charge across the ring, with the electrophile bonded to one carbon. The second step involves the rapid loss of a proton from the σ-complex, facilitated by a base, which restores the aromatic π-system and yields the substituted product. The was conceptually formalized by George W. Wheland in his 1942 analysis of aromatic reactivity, building on earlier kinetic studies. Substituents on the aromatic ring profoundly influence both the rate and regioselectivity of EAS by modulating the electron density and stabilizing (or destabilizing) the Wheland intermediate. Electron-donating groups, such as the hydroxy (-OH) in , activate the ring toward substitution and direct the electrophile preferentially to ortho and para positions relative to themselves, as these orientations allow better resonance stabilization of the positive charge in the intermediate. In contrast, electron-withdrawing groups like the nitro (-NO₂) deactivate the ring, slowing the reaction rate, and direct substitution to the meta position, where the Wheland intermediate experiences less charge buildup on the substituent-bearing carbon. For instance, in nitration, undergoes reaction approximately 25 times faster than overall, with partial rate factors indicating ortho and para positions are 42 and 67 times more reactive, respectively, relative to a single position in . Key examples of EAS illustrate its versatility in synthesis. Halogenation involves treatment with X₂ (X = Cl, Br) in the presence of a Lewis acid catalyst like FeX₃ to generate X⁺, yielding aryl halides; bromination of benzene, for example, proceeds cleanly under these conditions. Sulfonation employs fuming sulfuric acid or oleum to produce the electrophile SO₃, introducing a sulfonic acid group (-SO₃H) that is reversible under heating, useful for directing in polysubstitution sequences. Friedel-Crafts alkylation uses an alkyl halide (R–X) with AlCl₃ to form a carbocation electrophile, attaching an alkyl group, though rearrangements can occur with secondary or tertiary halides. Friedel-Crafts acylation, employing an acid chloride (R–COCl) and AlCl₃ to generate an acylium ion (R–CO⁺), introduces acyl groups without rearrangement and is widely used for ketone synthesis, as originally developed by and in 1877. Nitration, a cornerstone example, utilizes a mixture of HNO₃ and H₂SO₄ to produce the nitronium ion (NO₂⁺), enabling the preparation of nitroarenes central to explosives and pharmaceuticals.

Reduction and Addition Reactions

Aromatic compounds, despite their thermodynamic stability, can undergo reduction and addition reactions that disrupt their delocalized π-electron systems, leading to loss of aromaticity. Catalytic hydrogenation represents a primary method for fully saturating the aromatic ring, converting to using hydrogen gas in the presence of metal catalysts such as or . The reaction proceeds stepwise, with each addition of H₂ across the double bonds, but requires elevated temperatures (typically 100–200°C) and pressures due to the kinetic barrier imposed by aromatic stabilization./Aromatic_Compounds/Properties_of_Aromatic_Compounds/Hydrogenation_of_Arenes) The balanced equation for benzene hydrogenation is: \mathrm{C_6H_6 + 3H_2 \rightarrow C_6H_{12}} This process is highly exothermic, with a standard enthalpy change of approximately -208 kJ/mol, reflecting the release of strain from the planar aromatic system into the more flexible chair conformation. However, the reaction's slow kinetics without catalysts stem from the high activation energy needed to break the aromatic π-bonds. In substituted benzenes, the hydrogenation rate varies with substituent effects on adsorption to the catalyst surface; electron-withdrawing groups like nitro accelerate the process by enhancing binding, while electron-donating alkyl groups slightly deactivate the ring, leading to selectivity where the unsubstituted ring hydrogenates preferentially in mixtures. The Birch reduction offers a selective partial reduction, employing alkali metals (e.g., sodium or lithium) dissolved in liquid ammonia with a proton donor like ethanol, to convert into 1,4-cyclohexadiene. This method adds two electrons and two protons, targeting the meta positions relative to substituents in disubstituted cases, and preserves two isolated double bonds while eliminating . The mechanism involves initial electron addition to form a radical anion, followed by protonation and a second electron transfer, resulting in the unconjugated diene product that is thermodynamically favored over the conjugated 1,3-isomer due to reduced strain. This reaction is particularly useful for preparing non-aromatic cyclohexadienes from , with high yields under mild conditions (–78°C). Dearomatization via electrophilic addition occurs readily in electron-rich heterocycles like furan, where the highly reactive ring undergoes addition rather than substitution. For instance, furan reacts with bromine to form 2,5-dibromo-2,5-dihydrofuran, an addition product that saturates the ring and abolishes aromaticity through electrophilic attack at the 2,5-positions, stabilized by the oxygen lone pair. This contrasts with benzene's preference for substitution and highlights how heteroatom activation lowers the barrier for addition in such systems. In polycyclic aromatic hydrocarbons (PAHs), partial reduction to dihydro derivatives is achievable using selective catalysts like or rhodium complexes, targeting specific peripheral rings while preserving central aromaticity. For example, can be selectively hydrogenated to 1,4-dihydronaphthalene under controlled conditions (moderate pressure, 50–100°C), yielding the non-aromatic dihydro product with up to 90% selectivity. Such transformations are valuable in refining processes to reduce PAH toxicity and improve fuel stability, with catalyst choice dictating the extent of saturation.

Specific Classes and Examples

Benzene Derivatives

Benzene derivatives are compounds in which one or more hydrogen atoms on the benzene ring are replaced by substituents, leading to a wide array of monosubstituted and polysubstituted structures with distinct chemical behaviors. Common monosubstituted derivatives include toluene (methylbenzene), phenol (hydroxybenzene), aniline (aminobenzene), and nitrobenzene, each exhibiting properties influenced by the attached group. Toluene serves as a key solvent and precursor in organic synthesis, while phenol is valued for its role in resins and antiseptics. Aniline is essential in dye production, and nitrobenzene acts as an intermediate for explosives and pharmaceuticals. Nomenclature for benzene derivatives follows IUPAC guidelines, where the parent chain is benzene, and substituents are prefixed with locants for positioning. For monosubstituted compounds, common names like , , , and are retained and acceptable under IUPAC rules, especially when the substituent defines the compound's identity. In polysubstituted benzenes, particularly disubstituted ones, IUPAC names use numerical locants (e.g., 1-bromo-2-chlorobenzene) to indicate positions, with the lowest possible numbers assigned and substituents listed alphabetically. Common nomenclature employs directional terms: (1,2-), (1,3-), and (1,4-) for disubstitution patterns relative to the primary substituent. These patterns are crucial for describing spatial arrangements and reactivity./15%3A_Benzene_and_Aromaticity%3A_Electrophilic_Aromatic_Substitution/15.01%3A_Naming__the__Benzenes) Substituents on benzene significantly alter the ring's electronic properties, affecting acidity, basicity, and reactivity. For instance, the hydroxy group in enhances acidity compared to aliphatic alcohols due to resonance stabilization of the , with phenol having a pKa of approximately 10 versus 's pKa of 16. This delocalization spreads the negative charge into the ring, making the conjugate base more stable. Similarly, the amino group in decreases basicity relative to aliphatic amines because the nitrogen lone pair is delocalized into the benzene ring by resonance, reducing its availability for protonation. Electron-withdrawing groups like in decrease basicity and increase acidity of nearby protons through inductive effects. In electrophilic aromatic substitution, substituent directing effects determine product distribution, favoring ortho/para or meta positions based on electronic influence. Electron-donating substituents, such as methyl in or hydroxy in , activate the ring and direct incoming electrophiles to ortho and para positions due to increased electron density there, with para often preferred for steric reasons. Electron-withdrawing substituents, like nitro in , deactivate the ring and direct to meta positions to avoid destabilizing the intermediate at ortho/para sites. This regioselectivity influences isomer stability and yield in synthetic applications.Complete_and_Semesters_I_and_II/Map%3A_Organic_Chemistry(Wade)/18%3A_Reactions_of_Aromatic_Compounds/18.06%3A_Substituent_Effects_on_the_EAS_Reaction) A notable industrial example is cumene (isopropylbenzene), used primarily as a precursor in the Hock process for producing phenol and acetone, accounting for over 95% of global phenol synthesis. This derivative highlights how alkyl substituents enable large-scale petrochemical applications while maintaining the aromatic stability of benzene.

Polycyclic and Non-Benzenoid Arenes

Polycyclic aromatic hydrocarbons (PAHs) consist of two or more fused benzene rings sharing adjacent carbon-carbon bonds, resulting in extended conjugated π-systems that enhance stability through delocalization. Naphthalene, the simplest PAH, features two linearly fused six-membered rings with a total of 10 π-electrons delocalized across the structure, satisfying for aromaticity in a planar conformation. Phenanthrene represents an angular fusion variant with three rings and 14 π-electrons distributed in a Clar sextet pattern, contributing to its lower reactivity compared to linear counterparts like anthracene. Non-benzenoid arenes deviate from the standard benzene motif while exhibiting aromatic properties. Azulene, a bicyclic isomer of naphthalene, comprises a five-membered ring fused to a seven-membered ring, yielding 10 π-electrons in a non-alternant system that imparts a distinctive blue color due to intramolecular charge transfer and a dipole moment of 1.08 D. This asymmetry leads to polarized reactivity, with the five-membered ring acting as electron-rich and the seven-membered as electron-poor, enabling applications in materials with unique optical properties. Metallocenes extend aromaticity to organometallic frameworks. Ferrocene features two cyclopentadienyl anions sandwiching an iron(II) center, each ring contributing 6 π-electrons in an 18-electron configuration that confers remarkable thermal stability and aromatic delocalization, as evidenced by its resistance to decomposition up to 400°C. The parallel rings maintain a staggered conformation in the solid state, with bond lengths indicative of equalized C-C bonds consistent with aromaticity. Aromaticity also manifests in ionic species. The tropylium cation, a seven-membered ring with a positive charge, possesses 6 π-electrons in a planar, fully conjugated system, rendering it exceptionally stable as a salt like tropylium tetrafluoroborate, which resists nucleophilic attack under ambient conditions. This stability arises from equal bond lengths and delocalized charge, contrasting with typical carbocation reactivity. Certain PAHs pose significant health risks due to their environmental persistence and bioaccumulation. Benzopyrene, a five-ring PAH, is a potent carcinogen that forms DNA adducts upon metabolic activation, inducing tumors in lung, skin, and other tissues across animal models and linked to human cancers via occupational exposure. Chronic exposure to PAH mixtures, including benzopyrene, correlates with respiratory disorders, cardiovascular disease, and immune suppression in epidemiological studies. Synthesizing larger non-benzenoid systems like annulenes presents formidable challenges. Large annulenes, such as annulene and annulene, struggle with conformational flexibility that disrupts planarity, leading to twisted geometries that diminish aromatic stabilization despite satisfying the 4n+2 π-electron criterion; for instance, annulene adopts a non-planar twist to alleviate transannular steric repulsion between internal hydrogens. These distortions result in bond length alternation and reduced delocalization, complicating isolation of purely aromatic forms and requiring low-temperature or derivatized syntheses to enforce planarity.

Applications and Biological Significance

Industrial and Synthetic Uses

Aromatic compounds, particularly benzene and its derivatives, serve as foundational feedstocks in the petrochemical industry, enabling the production of a wide array of synthetic materials and chemicals. Global benzene production reached approximately 64 million tonnes in 2024, with projections for steady growth driven by demand in plastics and resins. This output underscores benzene's role as a key intermediate, derived primarily from catalytic reforming of naphtha and steam cracking of hydrocarbons. In petrochemical applications, benzene is predominantly used to synthesize styrene, which polymerizes into polystyrene for packaging, insulation, and consumer goods, accounting for about 50% of benzene consumption. Another major pathway involves benzene's alkylation to form cumene, which is oxidized to produce phenol and acetone—essential monomers for phenolic resins, adhesives, and epoxy coatings. These processes often employ electrophilic aromatic substitution reactions to introduce alkyl groups, highlighting the reactivity of aromatic rings in industrial synthesis. Aromatic amines like , derived from via hydrogenation, are critical for dye production, particularly synthesized through and coupling reactions, which color textiles, leather, and inks with vibrant hues. In explosives manufacturing, undergoes stepwise nitration to yield 2,4,6-trinitrotoluene (), a stable high explosive used in mining, demolition, and munitions due to its reliable detonation properties. Advanced materials leverage aromatic structures for enhanced performance; for instance, terephthalic acid from p-xylene oxidation forms , a polyester used in bottles, fibers, and films, representing over 60% of purified terephthalic acid applications. Heterocyclic aromatics such as polymerize into , which exhibit electrical conductivity upon doping and find use in organic electronics, sensors, and solar cells owing to their conjugated π-systems. Recent advances include bio-based production of aromatic compounds using engineered microbes, contributing to sustainable alternatives in chemical manufacturing as of 2025. Zeolite catalysts, prized for their shape-selective pores and acidity, facilitate efficient alkylation of aromatics with olefins or alkanes, minimizing polyalkylation side products in processes like cumene production and improving yields in detergent alkylate synthesis. These solid acid catalysts enable greener, continuous operations compared to traditional homogeneous systems, reducing environmental impact in large-scale aromatic derivatization.

Natural Occurrence and Roles

Aromatic compounds are ubiquitous in nature, primarily synthesized through the shikimate pathway, a seven-step metabolic process that converts phosphoenolpyruvate and erythrose-4-phosphate into chorismate, the precursor for phenylalanine, tyrosine, and tryptophan. This pathway, absent in animals but present in bacteria, fungi, algae, and plants, enables the production of thousands of secondary metabolites derived from these aromatic amino acids, including over 8,000 phenylpropanoids, contributing to structural integrity and environmental adaptation. In plants, the shikimate pathway feeds into the phenylpropanoid pathway, where phenylalanine is deaminated by phenylalanine ammonia-lyase to form trans-cinnamic acid, leading to diverse compounds like lignin and flavonoids. In plants, phenylpropanoids such as lignin provide mechanical support and water impermeability in cell walls, forming complex polymers from monolignols like coniferyl alcohol, which constitute up to 30% of dry plant biomass. Flavonoids, another major class, accumulate in epidermal tissues and absorb ultraviolet (UV) radiation, protecting photosynthetic tissues from UV-B damage by scavenging reactive oxygen species and preventing DNA strand breaks. For instance, anthocyanins and flavonols in leaves and fruits act as antioxidants under high UV exposure, enhancing plant survival in intense sunlight environments. Additionally, indole-derived auxins like indole-3-acetic acid (IAA) regulate growth processes such as cell elongation, apical dominance, and root development through auxin signaling pathways involving TIR1/AFB receptors and Aux/IAA repressors. In broader biological systems, aromatic compounds fulfill essential roles in proteins and cofactors. The aromatic amino acids phenylalanine, tyrosine, and tryptophan absorb UV light at 260–280 nm due to their conjugated ring systems, thereby shielding proteins from photodegradation and oxidative stress in organisms exposed to solar radiation. Tryptophan, with its indole side chain, is particularly effective in this UV-protective function and serves as a precursor for signaling molecules like serotonin in animals. Porphyrins, macrocyclic aromatics, form the core of heme in hemoglobin and myoglobin, facilitating oxygen transport and storage by coordinating iron in a planar conjugated system. These compounds also appear in chlorophyll for photosynthesis, underscoring their role in energy transfer across kingdoms. However, certain aromatic compounds pose risks as environmental toxins. Polycyclic aromatic hydrocarbons (PAHs), formed during incomplete combustion of organic matter in wildfires, vehicle exhaust, and industrial processes, persist in soil and water, bioaccumulating in food chains. Exposure to PAHs like benzopyrene induces DNA adducts, leading to carcinogenesis, respiratory diseases, and immune suppression in humans and wildlife, with the International Agency for Research on Cancer classifying several as Group 1 carcinogens. These pollutants highlight the dual nature of aromatics, balancing vital biological functions with ecological hazards.

Intermolecular Interactions

Arene-Arene Bonding

Arene-arene bonding refers to the non-covalent interactions between aromatic rings, primarily driven by the overlap of their π-electron systems, which arise from the delocalized π orbitals characteristic of . These interactions play a crucial role in stabilizing molecular assemblies without altering the aromaticity of the rings involved. The main types of arene-arene interactions include π-π stacking in a parallel-displaced configuration, where the aromatic rings are offset to maximize overlap while minimizing repulsion; edge-to-face or T-shaped geometry, in which the edge of one ring approaches the face of another perpendicularly; and , where a C-H bond acts as a hydrogen bond donor to the electron-rich π cloud of an adjacent ring. The parallel-displaced arrangement allows for favorable dispersion and electrostatic contributions, while the T-shaped form often dominates in sterically constrained environments due to reduced overlap repulsion. CH-π bonds, classified as weak hydrogen bonds between a soft acid (C-H) and soft base (π-system), further diversify these motifs in molecular assemblies. Energetically, these interactions are relatively weak, with binding energies for configurations typically ranging from 2 to 5 kcal/mol. For instance, high-level calculations on the yield interaction energies of approximately -1.5 kcal/mol for the parallel configuration, -2.5 kcal/mol for the T-shaped, and -2.5 kcal/mol for the slipped-parallel (parallel-displaced) form, highlighting their modest but cumulative stabilizing effect in larger systems. The attractive nature of arene-arene bonding stems from the quadrupole moment of aromatic rings, where the π-electron cloud creates a region of partial negative charge above and below the ring plane, contrasted by partial positive charge at the periphery due to the σ-framework. This electrostatic complementarity drives attraction: in parallel-displaced stacking, the negative π face of one ring aligns with the positive edge of another, while in T-shaped orientations, the positive hydrogen atoms of one ring interact with the negative π face of the partner. Such quadrupole-quadrupole interactions favor offset geometries over direct face-to-face stacking to avoid repulsion between the negative π clouds. In molecular crystals, arene-arene bonding significantly influences packing motifs and can lead to polymorphism, where the same molecule adopts different crystal forms due to varying interaction geometries. For example, π-π stacking often promotes layered or herringbone arrangements that enhance density and stability, while competing motifs like edge-to-face can result in distinct polymorphs with altered physical properties such as solubility or melting points. These interactions contribute to the overall lattice energy, guiding self-assembly in organic solids. Computational modeling of arene-arene bonding relies heavily on dispersion-corrected density functional theory (DFT), as standard DFT functionals underestimate the vital dispersion forces in π-π stacking. Methods incorporating London dispersion corrections, such as DFT-D3 or DFT-D4, accurately predict interaction energies and geometries for benzene dimers and larger aromatic systems, with double-hybrid functionals like PWPB95-D4 providing benchmark-level precision for non-covalent motifs. These approaches enable reliable simulations of crystal packing and polymorphism by balancing electrostatic, dispersion, and exchange contributions.

Stacking and Dimerization

In the benzene dimer, the most stable configuration is the parallel displaced geometry, where the aromatic rings are offset to maximize π-orbital overlap while minimizing electrostatic repulsion, as determined by high-level quantum chemical calculations and spectroscopic studies. This structure exhibits a binding energy of approximately 2.5 kcal/mol, measured through rotational spectroscopy and corroborated by coupled-cluster computations. Aromatic stacking interactions play crucial roles in biological systems, particularly in stabilizing macromolecular structures. In DNA, the π-π stacking between adjacent aromatic purine (adenine, guanine) and pyrimidine (thymine, cytosine) bases contributes significantly to the double helix's stability, with stacking energetics following the order purine-purine > purine-pyrimidine > pyrimidine-pyrimidine. Similarly, in , interactions between and residues facilitate core packing and secondary structure formation; for instance, T-shaped or parallel displaced orientations between these aromatic side chains enhance folding kinetics and thermodynamic stability in enzymes and structural proteins. Coronene, a consisting of seven fused rings, serves as an ideal model for the layered stacking in , where its disc-like molecules align in eclipsed or slightly displaced parallel orientations to mimic the AB stacking of sheets. Quantum mechanical studies of dimers reveal binding energies around 20 kcal/mol per pair (as of recent CCSD(T)/ calculations), reflecting the cumulative π-π and van der Waals forces that underpin graphite's cohesive interlayer interactions. Experimental evidence for aromatic dimerization has been provided by of systems, where photodimerization yields [2+2] cycloadducts from initially stacked monomers in the crystal lattice, confirming parallel or herringbone arrangements with intermolecular distances of about 3.5-4.0 . These structures highlight how photoexcitation promotes dimer formation from preorganized stacked geometries in the solid state. In , aromatic stacking drives the assembly of host-guest complexes, where π-π interactions between aromatic hosts (e.g., calixarenes or porphyrins) and guest molecules enable selective binding and molecular recognition. For example, coronene-based hosts encapsulate planar aromatic guests through multilayered stacking, achieving binding affinities up to 10-20 kcal/mol and facilitating applications in sensors and systems.

References

  1. [1]
    aromatic (A00441) - IUPAC Gold Book
    A cyclically conjugated molecular entity with a stability (due to delocalization ) significantly greater than that of a hypothetical localized structure.
  2. [2]
    Faraday - University of Oregon
    In 1825, as a result of research on illuminating gases, Faraday isolated and described benzene. In the 1820s he also conducted investigations of steel ...
  3. [3]
    [PDF] Lecture Summary 09 February 4, 2004 - Cook Group
    Feb 4, 2004 · Benzene Structure​​ Benzene was first isolated from whale oil in 1825 by Michael Faraday. It was found to be highly unsaturated, but did not show ...
  4. [4]
    Kekule
    This paper was the first one where he published his idea of the "ring" structure for benzene. In the second paper he drew it as a hexagon. Presumably 27 is n- ...
  5. [5]
    Kekulé's Dreams - MIT
    Nov 30, 2015 · Genius has been spoken of, and the Benzene Theory has been designated a work of genius.
  6. [6]
    Chapter 8: Conjugated compounds and aromaticity – OCLUE
    These are known as Huckel's Rule: Aromatic compounds are planar, cyclic, conjugated, and have 4n+2 π electrons in the π electron cloud.
  7. [7]
  8. [8]
    [PDF] Aromatic Compounds
    Hückel's Rule states that if the number of π electrons in the cyclic system is equal to (4N+2), where N is a whole number integer, then the system is aromatic. ...
  9. [9]
    Benzene and Other Aromatic Compounds - MSU chemistry
    Evidence for the enhanced thermodynamic stability of benzene was obtained from measurements of the heat released when double bonds in a six-carbon ring are ...
  10. [10]
    [PDF] Structure of benzene
    Aromatic properties are those properties of benzene that distinguish it from aliphatic hydrocarbons. Benzene: • A liquid that smells like gasoline. • Boils at ...
  11. [11]
    [PDF] Aromatic Compounds: Understanding the Fragrant World of Organic ...
    Jun 30, 2023 · The term “aromatic” originated from the fact that many early aromatic compounds were characterized by strong and pleasant odors. However ...<|separator|>
  12. [12]
    [PDF] Quantentheoretische Beiträge zum Benzolproblem
    II. Quantentheorie der induzierten Polaritäten. Von Erich Hückel in Stuttgart. Mit 12 Abbildungen. (Eingegangen am 19. August 1931).
  13. [13]
    Michael Faraday's sample of benzene | Royal Institution
    Benzene is a natural hydrocarbon and a component of crude oil. Faraday isolated this substance for the first time in 1825 while investigating an oily residue ...
  14. [14]
    Benzene - American Chemical Society
    Jul 3, 2023 · Seven years later, George Barger* and Frederick Philip Coyne at the University of Edinburgh determined its structure and synthesized it. Ever ...
  15. [15]
    4.2 Reactivity, Stability and Structure of Benzene - KPU Pressbooks
    The benzene ring shows special reactivity and stability. First, the most common reaction of benzene is substitution, instead of the addition reaction for ...<|control11|><|separator|>
  16. [16]
    Aromaticity and Antiaromaticity: How to Define Them - MDPI
    In 1855, August Wilhelm Hofmann was the first to use the term “aromatic” to designate a family of acids related to benzene [3]. The empirical formula of ...
  17. [17]
    [PDF] The 150th Anniversary of the Kekul Benzene Structure
    Sep 24, 2014 · The first page of the 1865 article.[4]. Angewandte. Chemie. 47. Angew ... That first short French paper on benzene theory[4, 10, 14] is.
  18. [18]
    Snakes, sausages and structural formulae | Feature - Chemistry World
    Oct 8, 2015 · Alternative suggestions like James Dewar's bridged molecule (1867) and Albert Ladenburg's prism (1869) gained only limited support. (Both were ...
  19. [19]
    Introduction: DelocalizationPi and Sigma | Chemical Reviews
    Thiele's theory was widely applied for over a decade until Willstätter synthesized cyclooctatetraene (1911, 1913) and found it to lack aromatic character ...
  20. [20]
    Quest for the Most Aromatic Pathway in Charged Expanded ...
    Oct 4, 2022 · In nonplanar and charged macrocycles, a discrepancy between electronic and magnetic descriptors is observed. Nevertheless, our work demonstrates ...Missing: debates articles
  21. [21]
    13.2 The Structure of Benzene - Chemistry LibreTexts
    Jun 5, 2019 · Because of the aromaticity of benzene, the resulting molecule is planar in shape with each C-C bond being 1.39 Å in length and each bond angle ...
  22. [22]
    Character table for the D 6h point group - gernot-katzers-spice-pages.
    Molecules with D6h symmetry are surprisingly rare. The common example is benzene, but relatively few derivatives can maintain the high symmetry, e. g ...
  23. [23]
    11.6: Delocalized Electrons: Bonding in the Benzene Molecule
    Jul 12, 2023 · Molecular orbital theory predicts (accurately) that the four π electrons are to some extent delocalized, or 'spread out', over the whole π system.Learning Objectives · Benzene · Molecular Orbitals and... · The Chemistry of Vision
  24. [24]
    An Energetic Measure of Aromaticity and Antiaromaticity Based on ...
    Feb 10, 2006 · Based on 1,3,5-hexatriene, which also has three double bonds, the ECRE of benzene is 36.7 kcal mol−1, whereas based on 1,3,5,7-octatetraene, ...
  25. [25]
    15.3: Nomenclature of Benzene Derivatives - Chemistry LibreTexts
    Jun 5, 2019 · Ortho-, Meta-, Para- (OMP) Nomenclature for Disubstituted Benzenes ; ortho- (o-): 1,2- (next to each other in a benzene ring) ; meta- (m): 1,3- ( ...
  26. [26]
    Bonding - Resonance - ChemTeam
    Starting around 1930, Linus Pauling developed what today is called "resonance theory," the currently accepted way to explain the bonding in these substances.
  27. [27]
    [PDF] Linus Pauling - Nobel Lecture
    In the resonance discussion of the benzene molecule the two Kekulé structures have to be described as hypothetical: it is not possible to synthesize molecules ...Missing: 1930s | Show results with:1930s
  28. [28]
    [PDF] HÜCKEL MOLECULAR ORBITAL THEORY
    E = 3EC=C = 6α + 6β which is off by 2β. We recall that β is negative, so that the πelectrons in benzene are more stable than a collection of three double bonds.
  29. [29]
    Aromaticity as a Cornerstone of Heterocyclic Chemistry
    The aromaticity concept is a cornerstone to rationalize and understand the structure and thus the behavior of heterocyclic compounds.Introduction · Aromaticity as an Enduring... · The Extent and Classification...
  30. [30]
    Aromaticity of Heterocycles - ScienceDirect.com
    An aromatic compound is cyclic and unsaturated with enhanced stability over simple olefinic compounds. Pyridine reacts less with electrophiles than benzene.
  31. [31]
    Rules for Aromaticity: The 4 Key Factors - Master Organic Chemistry
    Feb 23, 2017 · The four key rules for aromaticity are: cyclic, conjugated, [4n+2] pi electrons, and flat. If any is violated, no aromaticity is possible.
  32. [32]
    [PDF] Synthesis and Chemistry of Indole
    ➢ Aromatic: Indole is a planar molecule and follows Huckel's rule [(4n+2) π electrons]. All atoms in indole are sp2 hybridized and each of them possesses one ...
  33. [33]
    Azete - an overview | ScienceDirect Topics
    Azete is defined as the simplest antiaromatic heteroannulene, characterized by its thermal instability and extreme reactivity, with tri-tert-butylazete being ...
  34. [34]
    Why isn't azete aromatic? - Chemistry Stack Exchange
    Dec 13, 2013 · The compound is anti-aromatic. While counting the number of π-electrons, you count the electrons which are delocalized over the ring.Is cyanidin aromatic? - Chemistry Stack ExchangeIs biphenyl considered to be aromatic? - Chemistry Stack ExchangeMore results from chemistry.stackexchange.com
  35. [35]
    Naphthalene | C10H8 | CID 931 - PubChem - NIH
    The bicyclic aromatic fraction, with a boiling range of, e.g., 220 - 270 °C, is dealkylated under hydrogen pressure, either thermally above 700 °C, or ...
  36. [36]
    Nitrobenzene | C6H5NO2 | CID 7416 - PubChem - NIH
    Nitrobenzene appears as a pale yellow to dark brown liquid. Flash point 190 °F. Very slightly soluble in water. Toxic by inhalation and by skin absorption.
  37. [37]
    Hydrocarbons, Linear Alcohols and Acids - Densities
    Benzene - Density and Specific Weight vs. Temperature and Pressure. Online calculator, figures and table showing density and specific weight of benzene, C6H6 ...
  38. [38]
    Liquids - Dynamic Viscosities - The Engineering ToolBox
    Absolute (dynamic) viscosity values for some common fluids. Absolute or dynamic viscosities for some common liquids at temperature 300 K are indicated below.
  39. [39]
    Toluene | C6H5CH3 | CID 1140 - PubChem - NIH
    Toluene appears as a clear colorless liquid with a characteristic aromatic odor. Flash point 40 °F. Less dense than water (7.2 lb / gal) and insoluble in water.
  40. [40]
    UV-Visible Spectroscopy - MSU chemistry
    Benzene exhibits very strong light absorption near 180 nm (ε > 65,000) , weaker absorption at 200 nm (ε = 8,000) and a group of much weaker bands at 254 nm (ε = ...
  41. [41]
    IR Absorption Table
    Aromatic C-H Stretch Aromatic C-H Bending Aromatic C=C Bending, ~3030 (v) 860 - 680 (s) 1700 - 1500 (m,m) ; Alcohol/Phenol O-H Stretch, 3550 - 3200 (broad, s) ...
  42. [42]
    [PDF] IR_lectureNotes.pdf
    An absorption above 3000 cm-1 indicates C=C, either alkene or aromatic. Confirm the aromatic ring by finding peaks at 1600 and 1500 cm-1 and C-H out-of-plane ...
  43. [43]
    Mass Spectrometry - MSU chemistry
    Among simple organic compounds, the most stable molecular ions are those from aromatic rings, other conjugated pi-electron systems and cycloalkanes.
  44. [44]
    Mass Spectrometry - Examples - Chemistry and Biochemistry
    Aromatic: Molecular ion peaks are strong due to the stable structure. Mass spectrum of an aromatic. Naphthalene C10H8. MW = 128.17. Aromatic molecule. Take the ...
  45. [45]
    Electrophilic Aromatic Substitution - Wiley Online Library
    Nov 27, 2015 · Ingold, C. K. (1969) Structure and Mechanism in Organic Chemistry, 2nd Ed., Cornell University Press, Ithaca, NY, p. 330. Google Scholar.Missing: original | Show results with:original
  46. [46]
    RELATIVE REACTIVITY OF TOLUENE-BENZENE IN NITRONIUM ...
    A study was made to determine the relative reactivity of the toluene–benzene pair in a nitronium salt nitration, using the method of competitive rate ...
  47. [47]
    Benzene Selectivity in Competitive Arene Hydrogenation: Effects of ...
    Benzene selective hydrogenation, a potential approach for carcinogenic benzene removal from gasoline, is probed using benzene/toluene mixtures.
  48. [48]
    [PDF] Alshehri, Feras (2017) The hydrogenation of substituted benzenes ...
    The selectivity towards these products can be controlled by varying catalyst, support and reaction parameters [50, 51]. It is an important process from an.
  49. [49]
    [PDF] Selective Catalytic Hydrogenation of Polycyclic Aromatic ...
    Feb 13, 2015 · These compounds were partially hydrogenated with good to excellent selectivities by just optimizing the reaction conditions. The influence ...
  50. [50]
    CUMENE - NCBI - NIH
    Cumene is used primarily (95%) as an intermediate in the production of phenol and acetone. Other uses include: the manufacture of styrene, α-methylstyrene, ...
  51. [51]
    Organic Nomenclature - MSU chemistry
    The IUPAC nomenclature system is a set of logical rules devised and used by organic chemists to circumvent problems caused by arbitrary nomenclature.Alkanes · Cycloalkanes · Alkenes And Alkynes<|separator|>
  52. [52]
    Naming Aromatic Compounds - Chemistry Steps
    Many monosubstituted derivatives of benzene are named systematically by adding the name of the substituent to “benzene” which is the parent.
  53. [53]
    [PDF] Chapter 17: Alcohols and Phenols
    Alcohols have an OH group on a saturated carbon (sp3), while phenols have an OH group on a benzene ring carbon. Phenols are aromatic alcohols.
  54. [54]
    [PDF] CHAPTER 8. ACID-BASE REACTIONS - Organic Chemistry
    The alcohol proton has a pKa of about 15, and the phenol proton has a pKa of about 10: thus, the most acidic group on the molecule above is the phenol.
  55. [55]
    Aromatic Reactivity - MSU chemistry
    The chemical reactivity of benzene contrasts with that of the alkenes in that substitution reactions occur in preference to addition reactions.
  56. [56]
    [PDF] Cumene - U.S. Environmental Protection Agency
    Cumene is used in a variety of petroleum products. Acute (short-term) inhalation exposure to cumene may cause headaches, dizziness, drowsiness, ...
  57. [57]
    Polycyclic Aromatics - Chemistry LibreTexts
    Jan 22, 2023 · Anthracene is fused linearly, whereas phenanthrene is fused at an angle. This difference in fusions causes the phenanthrene to have five ...Missing: patterns | Show results with:patterns
  58. [58]
    2.4 Polycyclic aromatic hydrocarbons - Organic Chemistry II - Fiveable
    Linear PAHs (anthracene) generally more reactive than angular ones (phenanthrene) · Angular PAHs often more stable due to better electron delocalization ...
  59. [59]
    Polycyclic Aromatic Hydrocarbons: Sources, Toxicity, and ...
    Nov 5, 2020 · As per Clar's resonance structure rule, aromaticity of PAHs strongly depends on the number of aromatic π-sextets. Aromatic π-sextet rings ...
  60. [60]
    [PDF] Non-benzenoid-aromatics-part-2.pdf
    Azulene is blue in colour and has large dipole moment of 1.08 D (unusual for small aromatic compounds)- this is due to the intramolecular charge transfer from ...
  61. [61]
    Recent advances and future challenges in the bottom-up synthesis ...
    Jun 26, 2025 · Azulene, an isomer of naphthalene, is the smallest non-alternant, non-benzenoid aromatic compound (Figure 1b). It consists of an electron ...
  62. [62]
    (IUCr) Synthesis and physical properties of ferrocene derivatives ...
    Introduction. Ferrocene is a metallocene that shows a remarkable aromaticity combined with thermal stability. It is well known that ferrocene derivatives ...
  63. [63]
    31.1: Metallocenes - Chemistry LibreTexts
    Mar 5, 2021 · A metallocene is a compound typically consisting of two cyclopentadienyl anions bound to a metal center (M) in the oxidation state II.Missing: aromaticity | Show results with:aromaticity
  64. [64]
    Tropylium Ion, an Intriguing Moiety in Organic Chemistry - PMC - NIH
    May 15, 2023 · Thiophene-fused tropylium cations with terminal amino groups displayed exceptional stability (because of aromaticity) and high pka values. A ...
  65. [65]
    [PDF] Tropylium Tetrafluoroborate: Synthesis of a Stable, Aromatic Cation
    According to Hückel's rule, in order for a molecule to be aromatic it must have 4n+2 pi electrons which are in conjugation with one another. Cycloheptatriene.
  66. [66]
    BENZO[a]PYRENE - Chemical Agents and Related Occupations
    Benzo[a]pyrene is a carcinogen that induces tumours in many animal species. Some of the examples relevant for this review are: lung tumours in mice, rats, and ...Exposure Data · Cancer in Experimental Animals · Other Relevant Data
  67. [67]
    Benzo[a]pyrene—Environmental Occurrence, Human Exposure, and ...
    Recent research works have also implied that the exposure to PAHs is associated with increased risk of larynx, kidney, prostate, breast, blood (leukemia), brain ...
  68. [68]
    Structure of [18]Annulene Revisited: Challenges for Computing ...
    The continuing interest in the exact structures of these larger annulenes arises from attempts to understand the rivalry between, on one side, aromaticity and ...
  69. [69]
    Structure of [18]Annulene Revisited: Challenges for Computing ...
    The continuing interest in the exact structures of these larger annulenes arises from attempts to understand the rivalry between, on one side, aromaticity and ...
  70. [70]
    Benzene Market Size, Share | Global Industry Report, 2032
    The global Benzene market was approximately 51.5 million tonnes in 2022 and is anticipated to grow at a healthy CAGR of 4.23% during the forecast period ...
  71. [71]
    BENZENE - Chemical Agents and Related Occupations - NCBI - NIH
    The primary use of benzene today is in the manufacture of organic chemicals. In Europe, benzene is mainly used to make styrene, phenol, cyclohexane, aniline, ...
  72. [72]
    Benzene - Chevron Phillips Chemical
    Benzene is an industrial chemical ... Creation of styrene monomer is the largest use of benzene, followed by cumene/phenol, cyclohexane, and nitrobenzene.
  73. [73]
    [PDF] Benzene, Toluene, Xylene - eere.energy.gov
    Phenol is an intermediate in the production of phenolic resins, pharmaceuticals, and various plastics. Benzene, along with xylene and toluene, is also used as ...
  74. [74]
    Classifications, properties, recent synthesis and applications of azo ...
    Jan 31, 2020 · Most azo dyes are synthesized by diazotization of an aromatic primary amine, followed by coupling with one or more electron-rich nucleophiles ...
  75. [75]
    2,4,6-Trinitrotoluene - NCBI - NIH
    2,4,6-Trinitrotoluene is produced commercially by the nitration of toluene. It is used mainly as a high explosive in military and industrial applications.
  76. [76]
    Applications | INEOS Aromatics
    The largest application for PTA is in polyethylene terephthalate (PET) for the polyester industry to produce PET-bottles, textiles, film and moulded product ...Fibres For Clothing And Home... · Industrial And High... · Polyester Film And Tape
  77. [77]
    Conducting polymers: a comprehensive review on recent advances ...
    Typical conducting polymers include polyacetylene (PA), polyaniline (PANI), polypyrrole (PPy), polythiophene (PTH), poly(para-phenylene) (PPP), poly( ...
  78. [78]
    [PDF] Zeolite-Containing Catalysts in Alkylation Processes - lidsen
    Jul 5, 2022 · In the industry, zeolite-containing catalysts are used for isomerization, alkylation, catalytic cracking, hydrocracking, hydrotreatment, and ...
  79. [79]
    Mechanistic Characterization of Zeolite-Catalyzed Aromatic ...
    Feb 3, 2022 · (5) Moreover, zeolites have also been shown to effectively catalyze the dealkylation of the alkylphenolic monomers derived from lignin ...Introduction · Computational details · Results and Discussion · Conclusions
  80. [80]
    The shikimate pathway: gateway to metabolic diversity - PMC
    The shikimate pathway is the metabolic process responsible for the biosynthesis of the aromatic amino acids phenylalanine, tyrosine, and tryptophan.Missing: paper | Show results with:paper
  81. [81]
    THE SHIKIMATE PATHWAY - Annual Reviews
    Jun 1, 1999 · The shikimate pathway links metabolism of carbohydrates to biosynthesis of aromatic compounds. In a sequence of seven metabolic steps, phosphoenolpyruvate and ...Missing: paper | Show results with:paper
  82. [82]
    Phenylpropanoid Biosynthesis - ScienceDirect.com
    This review aims to give an update on the various facets of the general phenylpropanoid pathway, which is not only restricted to common lignin or flavonoid ...Review Article · Phenylpropanoid Esters And... · Anthocyanins, Aurones...
  83. [83]
    Phenylpropanoids metabolism: recent insight into stress tolerance ...
    Here, we provide a comprehensive update on the biosynthesis of phenylpropanoids, mainly lignin and flavonoids, their roles in biotic and abiotic interaction, ...3.1. Phenylpropanoid... · 4.1. Phenylpropanoid... · 4.2. Salinity Stress...
  84. [84]
    Recent advances on the roles of flavonoids as plant protective ...
    Aug 28, 2021 · All flavonoids absorb in the UV wavelengths, they mostly accumulate in the epidermis of plant cells and their biosynthesis is generally ...
  85. [85]
    Recent advances on the roles of flavonoids as plant protective ...
    All flavonoids absorb in the UV wavelengths, they mostly accumulate in the epidermis of plant cells and their biosynthesis is generally activated after UV ...
  86. [86]
    Biosynthetic Pathways and Functions of Indole-3-Acetic Acid in ... - NIH
    Aug 12, 2023 · Indole-3-acetic acid (IAA) belongs to the family of auxin indole derivatives. IAA regulates almost all aspects of plant growth and development, ...
  87. [87]
    Mycosporine-like amino acid and aromatic amino acid transcriptome ...
    Nov 26, 2020 · In particular, tryptophan is known to play a number of stress response and UV protective roles within plant and animal cells. Tryptophan is the ...
  88. [88]
    The Uniqueness of Tryptophan in Biology: Properties, Metabolism ...
    Nov 20, 2020 · Its side chain is indole, which is aromatic with a binuclear ring structure, whereas those of Phe, Tyr, and His are single-ring aromatics. In ...
  89. [89]
    Porphyrins—valuable pigments of life - Frontiers
    Porphyrin complexes are present in many natural systems and have significant biological roles, such as light harvesting, oxygen transport, and catalysis.
  90. [90]
    Chemistry of porphyrins in fossil plants and animals - RSC Publishing
    Feb 17, 2021 · Due to their extended conjugated aromatic ring system, porphyrins absorb light in the visible range and therefore show characteristic colors.
  91. [91]
    Polycyclic aromatic hydrocarbons. A review - Taylor & Francis Online
    PAHs are harmful persistent organic pollutants, while the high-molecular-weight (HMW) PAHs are even more detrimental to the environment and human health. The ...
  92. [92]
    Polycyclic Aromatic Hydrocarbons: Sources, Toxicity, and ... - Frontiers
    Naphthalene, anthracene, and benzo(a)pyrene are direct skin irritants ... Angular arrangement is thermodynamically the most stable configuration of PAHs ...
  93. [93]
    Pi-Pi Interaction - an overview | ScienceDirect Topics
    Pi-Pi interactions are attractive interactions between aromatic π systems through face-to-face stacking, involving electrostatic and dispersion forces.
  94. [94]
    Substituent Effects in Parallel‐Displaced π–π Stacking Interactions ...
    May 11, 2017 · Both host–guest geometries, namely π–π stacking and edge-to-face alignment of the pyridine and platform arene rings, were observed for the co- ...
  95. [95]
    CH/π hydrogen bonds in crystals - RSC Publishing
    May 10, 2004 · CH/π hydrogen bonds are interactions between a soft acid CH and a soft base π-system, mainly stabilized by dispersion forces.
  96. [96]
    Model Chemistry Calculations of Benzene Dimer Interaction
    The calculated interaction energies of the parallel, T-shaped ,and slipped-parallel benzene dimers are −1.48, −2.46, and −2.48 kcal/mol, respectively.
  97. [97]
    Polymorphism Driven by π–π Stacking and van der Waals ...
    Relatively strong and directional hydrogen-bonding interactions and partly directional π–π stacking interactions are most commonly utilized for rational design ...
  98. [98]
    Benchmarking London dispersion corrected density functional ...
    It is shown that dispersion-uncorrected DFT underestimates ion–π interactions significantly, even though electrostatic interactions dominate the overall binding ...
  99. [99]
    Highly Accurate Coupled Cluster Potential Energy Curves for the ...
    The small binding energy of the gas-phase benzene dimer (∼2−3 kcal mol-1) and the shallowness of the potential energy surface make it a challenge for both ...Introduction · Results and Discussion · Supporting Information Available
  100. [100]
    High accuracy benchmark calculations on the benzene dimer ...
    Oct 15, 2007 · There are two experimental estimates for the binding energy of the benzene dimer, 1.6 ± 0.2 kcal/mol [6] and 2.4 ± 0.4 kcal/mol [7]. Based on ...
  101. [101]
    Influence of the π–π interaction on the hydrogen bonding capacity of ...
    The experimentally observed stacking arrangement between consecutive bases in DNA and RNA/DNA double helices is shown to enhance their hydrogen bonding ability.
  102. [102]
    Aromatic-aromatic interactions in structures of proteins and ... - NIH
    These phe-tyr interactions have been found to play important roles in protein folding and stability [18]. Since T-shaped interactions are preferred, it can be ...
  103. [103]
    Morphology of a columnar stack of coronene molecules ...
    Nov 5, 2015 · Podeszwa28 named this stacking 'graphite' structure for the coronene dimer. He suggested that the 'graphite' structure cannot be the global ...
  104. [104]
    A benchmark quantum chemical study of the stacking interaction ...
    We estimate the binding of coronene on a graphite surface to be 37. 4 or 1. 56 kcal/mol per carbon atom (67. 5 meV/C atom). This is also our estimate for ...
  105. [105]
    X-Ray observations of orientations of photodimer in single crystals of ...
    Each dimer crystal forms is one of several (at least six) orientations permitted by the anthracene crystal, each orientation being associated with a specific ...
  106. [106]
    π-π stacking interactions: Non-negligible forces for stabilizing ...
    Jan 10, 2020 · We present a porous supramolecular framework (π-1) stabilized only by intermolecular π-π stacking interactions.
  107. [107]
    Exploring π–π interactions and electron transport in complexes ...
    This study investigates π-π interactions and electron transport in host-guest complexes, showing increased conductance when guests are embedded inside the host.