Fact-checked by Grok 2 weeks ago

Cyclohexane conformation

Cyclohexane conformation encompasses the dynamic spatial arrangements adopted by the molecule (C₆H₁₂), a saturated six-membered ring, to minimize energetic inherent in cyclic structures. Unlike smaller cycloalkanes, avoids significant and torsional by assuming non-planar, puckered forms rather than a flat . The predominant and most stable conformation is the chair form, where all carbon-carbon bonds are staggered, bond s approximate the ideal tetrahedral value of 109.5°, and atoms are fully eclipsed-free, resulting in zero net . Less stable conformations include the boat and twist-boat forms, which serve as intermediates or transition states during ring inversion. The boat conformation features four pairs of eclipsed C–H bonds and additional steric repulsion from "flagpole" hydrogens, elevating its energy to approximately 6.5 kcal/mol above the chair. The twist-boat, a slightly distorted variant, alleviates some of this torsional and transannular strain, with an energy of about 5.5 kcal/mol relative to the chair, making it a local minimum but still far less populated at room temperature. These energy differences arise from ab initio computational analyses and underpin the rapid interconversion between equivalent chair forms via a pseudorotation pathway, occurring on the microsecond timescale with a rate constant of about 10^5 s^{-1} at room temperature. The understanding of cyclohexane conformations originated from early 20th-century structural studies, with Norwegian chemist Odd Hassel using to confirm the chair preference in the 1940s, building on Adolf von Baeyer's 19th-century planar model that overestimated . British chemist extended this in the 1950s by applying conformational principles to predict reactivity in complex molecules like steroids, distinguishing axial (perpendicular to the ring plane) and equatorial (roughly parallel) substituent positions in the , where equatorial orientations minimize steric interactions and thus dominate stability. Their foundational work earned the 1969 and established conformational analysis as a cornerstone of organic stereochemistry, influencing predictions of molecular behavior in rings and beyond.

Fundamentals of Cyclohexane

Molecular Structure and Bonding

has the molecular formula C₆H₁₂ and consists of a six-membered ring composed entirely of carbon atoms, each bonded to two hydrogen atoms. All six carbon atoms in the ring are sp³ hybridized, forming a saturated with no multiple bonds. The carbon-carbon (C-C) bond lengths in are approximately 1.54 Å, consistent with single bonds between sp³-hybridized carbons in alkanes. The ideal bond angle for sp³-hybridized carbons is 109.5°, but in the cyclic structure, these angles experience distortion due to the constraints of the ring framework. The bonding in is characterized by a (σ) framework, where all C-C and C-H bonds are formed by the overlap of sp³ orbitals, resulting in a tetrahedral local around each carbon. arises within this framework from the eclipsing of bonds on adjacent carbons, contributing to the energetic preferences in ring conformations. To visualize the ring structure, is often represented using Newman projections, which depict the molecule by looking along a C-C bond to show the relative positions of substituents, or sawhorse models, which provide a three-dimensional of the carbon skeleton and attached hydrogens. These representations highlight the potential for torsional and angle , which drive conformational flexibility in the ring.

Strain and Flexibility in Rings

In cyclic hydrocarbons, ring strain manifests in three primary forms: angle strain, torsional strain, and steric strain. Angle strain results from the deviation of internal bond angles from the ideal tetrahedral value of 109.5° associated with sp³-hybridized carbon atoms. In smaller rings, this deviation is pronounced; for instance, cyclopropane enforces C-C-C bond angles of 60°, leading to significant angle strain that destabilizes the molecule./Alkanes/Properties_of_Alkanes/Cycloalkanes/Ring_Strain_and_the_Structure_of_Cycloalkanes) Torsional strain arises in planar or nearly planar rings due to eclipsing interactions between adjacent bonds, which prevent the preferred staggered conformation and increase electron repulsion. Steric strain occurs from close non-bonded contacts between hydrogen atoms or other groups, exacerbating the overall energy penalty in constrained geometries. These strain types combine to elevate the total ring strain energy, particularly in rings smaller than six members. Quantitative measures of total strain energy highlight the relative stabilities of cycloalkanes. Cyclopentane exhibits a strain energy of about 6.5 kcal/mol, primarily from torsional contributions in its puckered envelope conformation. In contrast, cyclohexane possesses negligible strain energy, approximately 0 kcal/mol, positioning it as the archetypal strain-free cyclic hydrocarbon. To alleviate torsional and angle strains, six-membered rings like employ puckering or non-planar distortions, which enable staggered bond arrangements while maintaining bond angles close to 109.5°. This flexibility allows to achieve minimal overall , underscoring its conformational adaptability compared to more rigid smaller rings.

Principal Conformations of Cyclohexane

Chair Conformation

The chair conformation of cyclohexane features a puckered ring structure in which the carbon-carbon bonds alternate between pointing upward and downward relative to a hypothetical plane through the ring, resulting in a three-dimensional shape that resembles a lounge chair. This arrangement allows all C-C-C bond angles to measure approximately 111.5°, which is very close to the ideal tetrahedral angle and minimizes angle strain. Furthermore, the bonds are fully staggered, eliminating torsional strain as there are no eclipsing interactions between adjacent C-H bonds. In this conformation, the twelve atoms are distinctly oriented: six axial hydrogens are aligned parallel to the ring's threefold (three pointing upward and three downward), while the six equatorial hydrogens extend roughly to this , lying near the ring's equatorial . This positioning arises from the chair's inherent , classified under the D3d , which includes a center of inversion, a principal C3 , and C2 axes, contributing to its overall stability. The conformation represents the global minimum for , with a relative of 0 kcal/ compared to higher-energy forms, due to the effective relief of both angle and torsional strain through ring puckering. Although minor steric interactions, such as gauche butane-like overlaps and 1,3-diaxial contacts between hydrogens, are present, they are negligible and do not significantly elevate the . This strain-free profile was first elucidated through studies by Odd Hassel in the , establishing the as the predominant structure in the gas phase.

Boat and Twist-Boat Conformations

The boat conformation of features a structure where four adjacent carbon atoms lie in a , with the remaining two carbons elevated above and below this at the "bow" and "stern" positions. This arrangement results in significant steric repulsion between the flagpole hydrogens at the bow and , which are approximately 1.8 apart, contributing an estimated 2.7 kcal/ to the overall . Additionally, the boat exhibits partial eclipsing along the C2–C3 and C5–C6 bonds, introducing torsional strain of about 3.7 kcal/, for a total energy approximately 6.5 kcal/ higher than the chair conformation. The boat possesses C2v , reflecting its molecular and a C2 axis bisecting the ring. The twist-boat conformation arises as a of the , where the ring is twisted to alleviate the flagpole steric repulsion by increasing the distance between those hydrogens. This adjustment lowers the energy relative to the , positioning the twist-boat at about 5.5 kcal/mol above the , as determined by direct spectroscopic measurement of the difference. Despite this relief, the twist-boat retains partial eclipsing, which sustains some torsional , though reduced compared to the . At , the high energies of these conformers result in negligible populations: the boat is effectively 0%, while the twist-boat accounts for less than 1% of the equilibrium mixture.

Half-Chair Transition State

The half-chair conformation of is characterized by a in which four consecutive carbon atoms lie approximately in a , while the two adjacent carbons are displaced out of this in opposite directions—one above and one below—leading to partial eclipsing of bonds along the . This arrangement distorts the ideal tetrahedral angles and introduces torsional strain from the eclipsed interactions, distinguishing it from the staggered bonds in the more stable chair form. calculations confirm this structure with C1 symmetry, a puckering of about 0.57 , and angles such as approximately 35° and -12° around the . As a rather than an energy minimum, the half-chair lies approximately 10-12 kcal/mol above the conformation, representing the highest point on the during conformational interconversions. This elevated energy stems primarily from the eclipsing of vicinal hydrogens and angle deformations, making it unstable and short-lived. Computational studies, including those using /6-31G* level, place its energy at around 12 kcal/mol relative to the chair, underscoring its role as a barrier rather than a populated . The half-chair plays a crucial role in the conformational pathways of , serving as the that links the to the and subsequent twist-boat forms during inversion. This facilitates the pseudorotation and overall chair-chair interconversion by allowing the to flex without breaking bonds. In the inversion process, the progresses from the through the half-chair to a twist-boat minimum before reaching the symmetric , enabling axial-equatorial exchanges. Spectroscopic evidence for the transient half-chair is derived from NMR studies of cyclohexane and its derivatives, where signal broadening and coalescence occur at low temperatures due to slowing of the inversion process through this high-energy state. For instance, variable-temperature ^1H NMR reveals the barrier height by monitoring the averaging of axial and equatorial protons, with coalescence temperatures indicating rates consistent with a 10.8 kcal/mol activation energy for passage via the half-chair. These observations confirm the half-chair's involvement without direct observation, as its lifetime is too brief for resolution.

Conformational Interconversions

Chair-Chair Inversion Mechanism

The chair-chair inversion mechanism in represents a dynamic that interconverts the two equivalent chair conformations of the , allowing for the exchange of axial and equatorial positions among all hydrogen atoms or substituents. This inversion occurs via a multistep pathway involving transitional forms, beginning with the distortion of the chair into a half-chair , where one carbon atom is elevated out of the ring plane while adjacent carbons adjust accordingly. From the half-chair, the ring progresses to a twist-boat local minimum, followed by a , then another twist-boat local minimum, before returning through a second half-chair to the inverted chair form. Central to this pathway is the involvement of twist-boat intermediates, which facilitate a pseudorotation—a continuous deformation of the ring without breaking bonds—that smooths the transition and avoids higher-energy barriers. During the full inversion, every axial position becomes equatorial, and vice versa, effectively inverting the of substituents around the ring while preserving their relative up or down orientation. This exchange is a direct consequence of the symmetric nature of the chair forms and the transitional geometries. The mechanism has been experimentally observed through low-temperature nuclear magnetic resonance (NMR) spectroscopy, where distinct signals for axial and equatorial protons are resolvable below approximately -60°C, indicating slowed inversion rates that allow the conformers to be distinguished on the NMR timescale. As temperature increases, these signals coalesce due to rapid interconversion, confirming the between the chairs via the described pathway.

Boat-Twist-Boat Pseudorotation

The boat-twist-boat pseudorotation in refers to a concerted, vibration-like motion in which the two adjacent pseudoequatorial carbon atoms that are twisted relative to the plane of the ring migrate continuously around the six-membered ring, interconverting equivalent twist-boat forms without passing through a true intermediate as a stable minimum. This process was first conceptualized as part of the conformational flexibility in six-membered rings, where the ring adopts an infinite number of intermediate geometries along a pseudorotational pathway defined by a phase angle varying from 0° to 360°. The pseudorotation preserves the C2 symmetry inherent to the twist-boat geometry, ensuring that all twist-boat conformers are identical in energy and structure, with no distinct "starting" or "ending" position distinguishable on the ring. In the twist-boat conformation, which features a puckering Q ≈ 0.737 and relieves some of the torsional and steric strain present in the boat form, this migration of twist sites occurs seamlessly. The energy barrier opposing this pseudorotation is notably low at approximately 1.4 kcal/mol, as determined by calculations at the HF/VDZ+P level, allowing for extremely rapid interconversions even at with rates on the order of picoseconds. Theoretical estimates place this barrier in the range of 0.8–1.7 kcal/mol, confirming the fluxional nature of the twist-boat without significant energetic cost. In contrast to a literal ring , pseudorotation involves no net inversion or reorientation of substituents; positions that are pseudoaxial or pseudoequatorial in one twist-boat remain so throughout the , distinguishing it as a pseudorotational rather than rotational process.

Energy Barriers and Rates

The for the –chair interconversion in is 10.8 kcal/mol, determined through low-temperature () spectroscopy by observing the coalescence of proton signals in deuterated . At 25 °C, this barrier corresponds to an interconversion rate of approximately 10^5 s^{-1}, allowing the two equivalent chair forms to equilibrate rapidly on the NMR timescale at . In contrast, the twist-boat conformation serves as a local minimum approximately 5.5 kcal/ above the , with pseudorotation among equivalent twist-boat forms occurring over a low barrier of 1.3 kcal/, rendering this process significantly faster than inversion and facilitating rapid averaging of positions within the twist-boat manifold. The conformation, lying about 6.5 kcal/ above the , represents a along the pseudorotation pathway, with the barrier for distortion from to adjacent twist-boat forms estimated at around 5 kcal/ in early conformational analyses, though more recent computations refine this to lower values near 1.6 kcal/. These interconversion rates exhibit strong temperature dependence, governed by the k = A \exp\left(-\frac{E_a}{RT}\right), where k is the rate constant, A is the (typically 10^{12}–10^{13} s^{-1} for conformational processes), E_a is the , R is the (1.987 cal ^{-1} ^{-1}), and T is the absolute temperature; consequently, a 10 °C rise can roughly double the chair inversion rate due to the exponential term. This temperature sensitivity enables experimental probing of barriers via variable-temperature NMR, where rates slow sufficiently at sub-ambient conditions to resolve conformational signals.

Substituted Cyclohexanes

Axial and Equatorial Substituents

In the chair conformation of , each of the six carbon atoms bears two substituents oriented in distinct directions: axial and equatorial positions. Axial bonds are oriented nearly parallel to the ring's of , extending vertically upward or downward; there are three axial hydrogens pointing up and three pointing down, alternating around the ring. Equatorial bonds, in contrast, are directed outward at an angle, roughly in the plane of the ring, with a slight tilt to accommodate the tetrahedral . These positions interconvert rapidly through chair inversion, a process with an energy barrier of approximately 45 kJ/mol that occurs on the timescale (rate constant ≈ 10^5 s^{-1}) at , rendering all twelve atoms equivalent on average in unsubstituted —each spending half its time in an axial position and half in an equatorial one. Structural studies reveal subtle differences in bond lengths between these positions. The equilibrium axial C-H bond length is 1.098 ± 0.001 Å, slightly longer than the equatorial C-H bond length of 1.093 ± 0.001 Å, as determined by femtosecond rotational coherence spectroscopy combined with calculations. To visualize these orientations, the chair conformation is commonly represented using a flat hexagonal drawing where axial bonds are depicted as vertical lines (up or down) and equatorial bonds as angled lines slanting outward from the hexagon's edges; alternatively, three-dimensional wedge-dash notation emphasizes the spatial arrangement, with solid wedges for bonds coming out of the plane and dashed lines for those receding behind.

Monosubstituted Derivatives

In monosubstituted derivatives of , a single exhibits a strong preference for the equatorial position in the chair conformation, as the axial orientation incurs unfavorable steric strain from 1,3-diaxial interactions. This preference is quantified by the , defined as the difference ΔG° between the axial and equatorial conformers (with the axial being higher in energy). For , the is 1.74 kcal/mol, resulting in an equilibrium of approximately 95% equatorial conformer at 25°C. The conformational equilibrium is governed by the equation K = \frac{[\text{equatorial}]}{[\text{axial}]} = e^{-\Delta G^\circ / RT} where K is the , R is the (0.001987 kcal/mol·K), and T is the absolute temperature. This relation, derived from the , directly links the to the conformer ratio and was used to determine preferences via low-temperature NMR spectroscopy. Representative A-values for other common substituents include 0.87 kcal/mol for the hydroxyl group (-OH) and 0.43 kcal/mol for the chloro group (-Cl), indicating progressively weaker equatorial biases compared to methyl but still favoring the equatorial position in the . These values were obtained through of separate axial and equatorial proton signals in NMR spectra at approximately -80°C in solvent. The rate of chair-chair inversion in monosubstituted remains largely unaffected by small substituents such as -CH₃, -OH, or -Cl, with activation energies close to that of unsubstituted (10.2 kcal/mol), as measured by NMR coalescence temperatures; for example, has a barrier of 10.8 kcal/mol. In contrast, bulky substituents like t-butyl raise the barrier further (11.0 kcal/mol for t-butylcyclohexane), slowing the inversion rate due to enhanced steric hindrance in the half-chair .

Disubstituted Derivatives

Disubstituted cyclohexanes exhibit , where the relative positions of the substituents influence the preferred chair conformations and overall stability. In these derivatives, the equatorial preference of substituents, as quantified by A-values from monosubstituted analogs, determines the dominant conformer, with diequatorial arrangements generally favored when possible. For 1,2-disubstituted cyclohexanes, the isomer adopts a diequatorial conformation in its most stable form, while the alternative diaxial conformer is less populated due to increased steric crowding. In contrast, the isomer features one axial and one equatorial substituent in both conformations, which are of equal and interconvert rapidly via ring flipping. A representative example is 1,2-dimethylcyclohexane, where the isomer exists as a pair of enantiomers—each with the diequatorial conformation as the predominant form—while the isomer is an achiral relative to the trans due to rapid interconversion of its enantiomeric conformations. In 1,3-disubstituted cyclohexanes, the isomer prefers the diequatorial conformation for stability, with the diaxial alternative being higher in energy. The isomer, however, has one axial and one equatorial substituent in both chair forms, resulting in equivalent conformers. For 1,4-disubstituted cyclohexanes, the isomer can adopt either a diequatorial (preferred) or diaxial conformation, with the former dominating due to minimized steric interactions. The isomer is restricted to one axial and one equatorial substituent in both chairs, which are equally stable when the substituents are identical.

Steric and Energetic Interactions

1,3-Diaxial Interactions

In the chair conformation of , 1,3-diaxial interactions arise from steric repulsions between axial s (or hydrogens) located at the 1 and 3 positions, as well as the 1 and 5 positions, on the same face of the ring. These pairs are oriented parallel and in close proximity, leading to unfavorable non-bonded contacts that destabilize the axial orientation relative to equatorial. The concept is rooted in early conformational studies, where such interactions were recognized as key to understanding preferences. The distance between the axial hydrogens in a 1,3-diaxial pair is approximately 2.5 , which is approximately equal to the sum of their van der Waals radii (about 2.4 ), resulting in repulsive steric strain. Each such contributes roughly 0.9 to the overall penalty, analogous to the gauche interaction in . For an axial substituent, the group experiences two such 1,3-diaxial interactions with the ring hydrogens—one at the 3-position and one at the 5-position—yielding a total strain of about 1.8 (2 × 0.9 ), which closely matches the observed for . This model highlights how the axial methyl's hydrogens mimic the gauche arrangement with the syn-axial C-H bonds. For larger substituents like tert-butyl, the 1,3-diaxial repulsions are amplified due to increased steric bulk. The axial tert-butyl group incurs severe interactions with the two syn-axial hydrogens, with the total cost approximating 4.9 kcal/, often conceptualized as four effective gauche-like interactions accounting for the branched structure's extended contacts. This substantial penalty locks the tert-butyl in the equatorial position, providing a rigid anchor for studying other substituents in disubstituted systems. Computational and experimental analyses confirm these values, emphasizing the role of van der Waals overlaps in driving conformational bias. To illustrate the geometry, consider the chair where axial positions align nearly parallel:
  • Axial H at C1 interacts with axial H at (distance ~2.5 ).
  • Similar for C1 and C5.
These close approaches underscore the repulsive nature, with energy scaling roughly with size but dominated by pairwise contacts in the standard model.

Gauche Butane Interactions

In n-butane, the gauche conformation incurs a of approximately 0.9 kcal/mol relative to the conformation, arising from the overlap of the methyl groups at a of 60° along the central C2–C3 bond. This interaction serves as a model for vicinal in larger systems, including derivatives. In 1,2-disubstituted cyclohexanes, the gauche interaction manifests in arrangements where the substituents on adjacent carbons adopt a 60° . For the 1,2-trans in its diaxial conformation, the substituents are antiperiplanar (180° ), incurring no such penalty between them. In contrast, the 1,2-cis in its axial-equatorial conformation features one gauche interaction between the substituents, contributing an energetic penalty of about 0.9 kcal/mol for methyl groups, as observed in cis-1,2-dimethylcyclohexane. For larger substituents, these gauche effects can be additive, increasing the overall strain beyond the simple methyl-methyl case, though the precise magnitude depends on the groups' sizes and the ring's constraints. Unlike acyclic alkanes, where rotation can minimize steric overlap by achieving an arrangement, the ring rigidly enforces angles near 60° in equatorial or mixed positions, thereby perpetuating the gauche penalty in preferred conformations.

Substituent Size Effects on Stability

The stability of conformations is significantly influenced by the size of substituents, as larger groups experience greater steric repulsion in axial positions, primarily through amplified 1,3-diaxial and gauche interactions. This effect is quantified by , which represent the difference (ΔG°) between axial and equatorial positions for a monosubstituted , measured in kcal/mol. Small substituents like exhibit minimal preference for the equatorial position, with an A-value of 0.15 kcal/mol, reflecting limited steric hindrance. In contrast, bulkier alkyl groups show progressively larger A-values, indicating stronger destabilization when axial: ethyl (1.75 kcal/mol), isopropyl (2.15 kcal/mol), and tert-butyl (4.9 kcal/mol). These trends arise because increasing substituent volume intensifies non-bonded repulsions with the ring hydrogens, favoring the equatorial orientation to minimize energy. For extremely bulky groups such as tert-butyl, the equatorial preference is nearly absolute (>99.9% equatorial at ), effectively locking the ring in one conformation and preventing observable chair inversion under typical conditions. The large results in the axial tert-butyl conformer being negligibly populated (<0.1%) at , despite the inversion barrier remaining ~11 kcal/mol.
SubstituentA-Value (kcal/mol)
F0.15
CH₃1.70
CH₂CH₃1.75
CH(CH₃)₂2.15
C(CH₃)₃4.9
This conformational rigidity imparted by large substituents has important implications in organic synthesis, where tert-butyl groups are often employed as protecting or directing moieties to fix the cyclohexane ring in a predictable chair form, facilitating stereoselective reactions and simplifying product analysis.

Equilibrium and Influences

Conformational Preferences

The conformational preferences of substituted cyclohexanes are governed by the relative stabilities of their chair conformers, with the population distribution at equilibrium determined by the Boltzmann distribution. For a monosubstituted cyclohexane, the percentage of the equatorial conformer is given by \%_{\text{eq}} = \frac{100}{1 + e^{\Delta G / RT}} where \Delta G is the free energy difference between the axial and equatorial forms (often denoted as the A-value), R is the gas constant, and T is the temperature in Kelvin. This equation arises from the equilibrium constant K = n_{\text{eq}} / n_{\text{ax}} = e^{-\Delta G / RT}, allowing direct prediction of conformer ratios from thermodynamic data. In disubstituted cyclohexanes, A-values are approximately additive for independent substituent positions, such as in 1,4-trans or 1,3-cis isomers, where the total \Delta G is the sum of individual A-values, leading to predictable population distributions via the Boltzmann relation. However, in geminal (1,1-disubstituted) cases, additivity breaks down due to direct interactions between the substituents on the same carbon, which alter the effective energy differences beyond simple summation. For example, in unsymmetrical geminal disubstitution like 1-ethyl-1-methylcyclohexane, the preference for the conformer with the smaller methyl group axial (larger ethyl equatorial) reflects this coupling, resulting in non-additive stabilization. Low-temperature NMR studies have confirmed these equilibria by slowing ring inversion to freeze individual conformers, enabling direct measurement of populations through signal integration. In derivatives like 1,1,4,4-tetramethylcyclohexane, such studies at reduced temperatures reveal a strong preference for chair over twist-boat forms, with axial methyl compressions still disfavoring the axial-rich conformer by observable ratios. These frozen-state observations validate Boltzmann predictions under standard conditions and highlight the dominance of steric factors in dictating conformer abundance. Overall, the major conformer in substituted cyclohexanes is predicted by minimizing the total \Delta G, calculated from summed A-values for independent cases or adjusted for interactions in coupled systems like geminal substitution, ensuring the lowest-energy arrangement predominates in the equilibrium mixture.

Solvent and Temperature Effects

Polar solvents can significantly influence the conformational equilibrium of monosubstituted cyclohexanes by stabilizing polar axial substituents through enhanced solvation interactions. For instance, in cyclohexanol, the A-value for the OH group shows solvent dependence, increasing from approximately 0.6 kcal/mol in nonpolar solvents or gas phase to 0.9 kcal/mol in polar protic solvents like water, due to better solvation of the equatorial OH through hydrogen bonding, which enhances the energy penalty for the axial orientation compared to non-polar environments. Temperature variations affect conformational equilibria by altering the population distribution according to the Boltzmann factor, with higher temperatures increasing the proportion of the higher-energy minor conformer. This temperature dependence can be analyzed using the van't Hoff equation, which relates the equilibrium constant K = \frac{[\text{equatorial}]}{[\text{axial}]} to temperature via \ln K = -\frac{\Delta H^\circ}{RT} + \frac{\Delta S^\circ}{R}, allowing determination of enthalpic (\Delta H^\circ) and entropic (\Delta S^\circ) contributions from plots of \ln K versus $1/T. In practice, for derivatives like chlorocyclohexane, such analyses reveal that the equatorial preference diminishes at elevated temperatures, as the thermal energy overcomes steric barriers. For bulky substituents, entropy plays a notable role in the conformational preference, often favoring the equatorial position due to greater rotational freedom and disorder in the surrounding molecular environment. In the axial orientation, large groups like tert-butyl experience restricted conformations, leading to a lower entropy state compared to the equatorial, where multiple rotameric forms are accessible; for example, the \Delta S^\circ for is approximately -0.44 cal/mol·K, contributing to the overall equatorial stabilization alongside enthalpic factors. This entropic contribution becomes more pronounced with increasing substituent size, enhancing the disorder in the preferred equatorial conformer.

Extensions and Applications

Heterocyclic Analogs

Heterocyclic analogs of cyclohexane, such as tetrahydropyran and piperidine, exhibit conformational behaviors that parallel the chair preference of the parent hydrocarbon but are modulated by the presence of heteroatoms, leading to alterations in bond angles, inversion barriers, and substituent preferences. In these six-membered rings, the chair conformation remains the dominant form at room temperature, akin to cyclohexane, but the electronegativity and size of the heteroatom introduce deviations that affect stability and dynamics. Tetrahydropyran, the oxygen analog of , strongly favors the chair conformation, with the ring oxygen's high electronegativity stabilizing orientations through reduced dipole interactions and better alignment with adjacent C-H bonds. This preference is evident in substituted derivatives, where the oxygen's electronegativity influences adjacent substituents. The C-O-C bond angle in tetrahydropyran is approximately 110°, narrower than the 111.4°-111.8° CCC angles in , due to the oxygen's sp³ hybridization and lone pair repulsion, which slightly puckers the ring and influences overall torsional strain. Piperidine, the nitrogen analog, also adopts a chair conformation but displays faster inversion dynamics compared to cyclohexane, with the nitrogen inversion barrier around 5-6 kcal/mol versus the 10-12 kcal/mol ring flip barrier in the hydrocarbon, allowing rapid interconversion between conformers even at low temperatures. The nitrogen lone pair shows a preference for the axial position in the predominant conformer (approximately 73% axial at 298 K, corresponding to equatorial N-H), but equatorial orientation occurs in a significant minority (about 27%), due to the lone pair's reduced 1,3-diaxial interactions relative to an axial hydrogen. This axial lone pair population is higher than in alkyl-substituted cyclohexanes, reflecting nitrogen's lower steric bulk and partial p-character in the orbital, which alters A-values for substituents (e.g., N-substituents have A-values of 0.5-1.0 kcal/mol, smaller than methyl's 1.7 kcal/mol). These heterocyclic systems highlight how heteroatom incorporation modifies cyclohexane-like geometry; for instance, the C-N-C bond angle in piperidine is about 111°, similar to CCC in cyclohexane but with greater flexibility due to the lone pair. An analogous deviation appears in cyclohexene, where the endocyclic double bond causes partial flattening of the half-chair conformation, reducing the pseudorotational amplitude with a barrier to pseudorotation of approximately 5-7 kcal/mol.

Role in Organic Synthesis and Spectroscopy

Cyclohexane's chair conformation plays a pivotal role in organic synthesis by guiding the design of protecting groups and enabling stereoselective transformations. In particular, protecting 1,2-diols as cyclohexylidene diacetals (CDAs) exploits the chair geometry to selectively shield vicinal hydroxyl groups in diequatorial positions, locking the ring and preventing unwanted reactivity during multi-step syntheses. This approach has been employed in the preparation of complex natural product fragments, where the CDA enforces conformational rigidity, ensuring high diastereoselectivity in subsequent functionalizations. For instance, in the synthesis of differentially protected saccharides, chiral phosphoric acid catalysts facilitate regioselective CDA formation on cyclohexane-derived diols, leveraging the chair's steric preferences to achieve >95% selectivity for the desired . Stereoselective reactions further highlight the chair's utility, as it dictates approach vectors for reagents in substituted cyclohexanes. In asymmetric epoxidations of allylic alcohols derived from cyclohexane scaffolds, the chair conformation positions bulky groups equatorially, directing nucleophiles to one face and yielding trans-diols with enantiomeric excesses exceeding 90%. Similarly, in aldol additions to enolates, the locked chair minimizes 1,3-diaxial interactions, favoring anti-products in up to 20:1 diastereomeric ratios, a principle central to total syntheses like that of zaragozic acid. In , (NMR) exploits conformational differences to assign axial and equatorial protons via vicinal coupling constants. In the chair form, axial-axial (ax-ax) couplings average ~12 Hz, while equatorial-equatorial (eq-eq) and axial-equatorial (ax-eq) values are ~4 Hz and ~2-5 Hz, respectively, allowing unambiguous stereochemical determination in disubstituted cyclohexanes. These J-values, derived from the Karplus relationship, enable low-temperature NMR studies to quantify conformational equilibria, such as in 4-tert-butylcyclohexanol, where the equatorial conformer predominates by 99% at 25°C. Infrared (IR) and distinguish chair and boat/twist-boat conformers through distinct vibrational modes. The chair's D_{3d} yields inactive symmetric C-H stretches (~2850-2950 cm^{-1}) in IR but active in Raman, while boat forms show additional IR bands near 1000 cm^{-1} due to ring-puckering modes, facilitating detection of minor conformers (<1%) at low temperatures. These spectral signatures confirmed the chair-boat energy gap as 5.5 kcal/mol via temperature-dependent IR measurements. Computational modeling of cyclohexane conformations relies on force fields like , which accurately predict the chair as the global minimum with a boat barrier of ~10.8 kcal/mol, matching experimental values within 0.2 kcal/mol. 's parameterization on ab initio data ensures reliable energy profiles for substituted systems, aiding virtual screening in drug design where equatorial preferences influence binding affinities.

Historical Context

Early Models and Discoveries

In the late 19th century, the understanding of cycloalkane structures, including , was dominated by the assumption of planarity. In 1885, proposed his to explain the relative stabilities of small-ring cycloalkanes, positing that their rings are flat like , resulting in angle strain due to bond angles deviating from the ideal tetrahedral value of 109.5° toward 60° in , 90° in , and 108° in , with experiencing minimal strain at 120° but still assumed planar. This planar model persisted into the early 20th century despite mathematical challenges to it. In 1890, Hermann Sachse demonstrated through geometric analysis that non-planar conformations of , specifically the chair and boat forms, could achieve strain-free tetrahedral geometry without angle distortion, introducing the concepts of axial and equatorial positions for substituents. However, Sachse's ideas were largely overlooked for decades, as chemists favored the simplicity of Baeyer's planar framework and lacked experimental evidence for puckered rings. In 1918, Ernst Mohr revived these concepts by applying them to the structure of diamond and fused ring systems like trans-decalin, demonstrating that non-planar arrangements better explained observed stabilities, though full experimental validation remained pending. Experimental confirmation of non-planar conformations emerged in the mid-20th century through physical methods. In 1947, Odd Hassel applied electron diffraction to cyclohexane vapor, providing the first direct structural evidence favoring the chair conformation over planar or boat forms due to minimized torsional strain and optimized bond distances. Building on this, Derek H. R. Barton in 1950 analyzed X-ray crystallographic data from steroid crystals, showing that the fused cyclohexane rings adopt chair conformations to accommodate observed bond lengths, angles, and substituent orientations without excessive strain. The visualization of these conformations advanced with the development of physical molecular models in the 1950s. Robert B. Corey and Linus Pauling introduced space-filling atomic models at Caltech, which accurately represented van der Waals radii and allowed construction of the chair form of cyclohexane, highlighting its stability relative to other puckered variants. These models, later refined by Walter Koltun into the widely used Corey-Pauling-Koltun (CPK) system in the early 1960s, facilitated broader acceptance of conformational analysis by enabling tangible demonstrations of ring puckering and substituent effects.

Key Contributors and Milestones

Odd Hassel pioneered the experimental confirmation of the chair conformation of cyclohexane through electron diffraction studies in the late 1940s. His work, including investigations of cyclohexane and its derivatives, demonstrated that the chair form is the preferred stable conformation due to minimized torsional strain, as evidenced by diffraction patterns of gaseous molecules. For these contributions to conformational analysis, Hassel shared the 1969 Nobel Prize in Chemistry with Derek Barton. In 1950, Derek Barton advanced conformational analysis by applying it to natural products, particularly steroids, in his seminal paper "The Conformation of the Steroid Nucleus." Barton illustrated how the chair conformation of cyclohexane rings influences the reactivity and biological activity of molecules like cholesterol and sex hormones, establishing a framework for predicting chemical behavior based on three-dimensional structure. This approach revolutionized organic chemistry and earned Barton a share of the 1969 Nobel Prize in Chemistry. The 1960s saw key developments in probing conformational dynamics using nuclear magnetic resonance (NMR) spectroscopy, notably by Frank A. L. Anet and A. J. R. Bourn. Their 1967 study on -d11 provided the first quantitative NMR evidence for ring inversion rates and activation energies, confirming the low barrier (approximately 10.8 kcal/mol) between chair conformers. Concurrently, Ernest L. Eliel's 1962 textbook Stereochemistry of Carbon Compounds systematized conformational principles, offering a comprehensive resource that integrated experimental data with theoretical insights for and related systems. Computational methods in the 1970s further validated cyclohexane conformations through early quantum mechanical calculations. John R. Hoyland's 1969 ab initio Hartree-Fock study, extended into the decade, calculated relative energies of chair and boat forms, predicting the chair as 5.5-6.0 kcal/mol more stable, aligning with experimental values and supporting the dominance of the chair in equilibrium. These quantum mechanical approaches marked a milestone in theoretically confirming structural preferences without reliance on empirical models.

References

  1. [1]
    Conformational analysis of cycloalkanes | ChemTexts
    Aug 12, 2015 · The two extreme conformations of cyclohexane are chair and boat. Chair conformation is the one without strain (Fig. 15b). In a chair ...
  2. [2]
    Cyclohexane Conformational Analysis
    The ground state conformation of cyclohexane is a fully staggered conformation which is shaped somewhat like a "chair". In this conformation there is no ...Missing: scholarly | Show results with:scholarly
  3. [3]
    Ab initio conformational analysis of cyclohexane - ACS Publications
    Comparison of ab initio, semiempirical, and molecular mechanics calculations for the conformational analysis of ring systems.
  4. [4]
    [PDF] The principles of conformational analysis - Nobel Prize
    It is obvious that in the chair conformation of cyclohexane two geometri- cally distinct types of carbon-hydrogen bonds are presents3,11. Six of the C-H bonds ...
  5. [5]
    Cyclohexane | C6H12 | CID 8078 - PubChem - NIH
    Cyclohexane is an alicyclic hydrocarbon comprising a ring of six carbon atoms; the cyclic form of hexane, used as a raw material in the manufacture of nylon.Missing: sp3 | Show results with:sp3
  6. [6]
    [PDF] Chapter 3 Alkanes & Cycloalkanes - SDSU Chemistry
    Feb 10, 2019 · All bond angles are approximately. 109.5°. Shapes of Alkanes. ❖ All carbon atoms in alkanes and cycloalkanes are sp3-hybridized, and.
  7. [7]
    Cycloalkanes and strain - Oregon State University
    ... bond lengths of 1.54 Å, bond angles of 109.5°. Examine the five simplest ... C-C: 1.508 Å C-C-C: 60° C-C-H: 118.2°. Cyclobutane Total strain: 26.3 kcal ...
  8. [8]
    4.1: Cyclohexane Ring Conformations - Chemistry LibreTexts
    Sep 21, 2022 · Because of the tetrahedral carbon atoms in the ring, each of the bond angles is 109.5°. This makes the cyclohexane ring pucker to give it three- ...
  9. [9]
    [PDF] Rings (Cycloalkanes)
    This geometry causes strain due to both small bond angles and torsional strain. The <C-C-C bond angle has the most strain (fixed at 60˚ in space) compared to ...
  10. [10]
    4.3: Conformation Analysis of Cyclohexane - Chemistry LibreTexts
    Dec 15, 2021 · These two bonds will be represented by two “Newman projections” that we are familiar with (two circle things) and each represents two carbons, ...Chair conformation of... · Properties of the chair... · Ring flipping
  11. [11]
  12. [12]
    The Energy Difference between the Chair and Boat Forms of ...
    The energy difference between the chair and boat forms of cyclohexane. The twist conformation of cyclohexane.
  13. [13]
    Cyclohexane (boat) C2v - ChemTube3D
    Click the Symmetry Operations above to view them in 3D C 2v point groups contain one C 2 rotation and 2σ v planes.
  14. [14]
    Spectroscopic detection of the twist-boat conformation of ...
    A simple method for the visualization of chair and twist-boat transition states in torsionally controlled addition reactions.Missing: seminal | Show results with:seminal
  15. [15]
    Conformation of cyclohexane - BYJU'S
    These twist-boat conformations of cyclohexane are much more stable than their boat-shaped counterparts. This conformation has a concentration of less than 1% ...
  16. [16]
    [PDF] Ab Initio Studies of Six-Membered Rings - SMU
    At the HF/4-21G level, the normal. C-C bond length of the cyclohexane chair is 1.54 A.For axial substituents C(1)-C(2) bond lengths are closer to this value ...
  17. [17]
    THE ENERGY BARRIER FOR THE CHAIR-CHAIR ...
    THE ENERGY BARRIER FOR THE CHAIR-CHAIR INTERCONVERSION OF CYCLOHEXANE. Click ... Half-Chair Inversion in Dioxene and 2,3-Dihydropyran. The Journal of ...
  18. [18]
    Dynamics and Entropy of Cyclohexane Rings Control pH ... - NIH
    The chair form is classically known as the thermodynamically favored conformation among the various conformations of six-membered ring compounds. The chair ...Missing: review paper
  19. [19]
  20. [20]
    The Rate of the Chair-Chair Interconversion of Cyclohexane
    ESR studies of barriers to ring inversion in cyclic monocarboxylic acid radicals. The Journal of Chemical Physics 1980, 72 (2) , 1325-1331. https://doi.org ...Missing: seminal mechanism
  21. [21]
    Molecular Geometry. I. Machine Computation of the Common Rings
    Type conformations and pseudorotation interconversion path-way between conformations ... Conformational pathways of simple six‐membered rings. Journal of Physical ...
  22. [22]
    Conformation of non-aromatic ring compounds, Part L boat/twist ...
    Theoretical calculations indicate a barrier opposing the B/TB pseudorotation of 0.8–1.7 kcal/mole for cyclohexane (16,17,19,20), 2.2 kcal/mole for cyclohexanone ...
  23. [23]
  24. [24]
  25. [25]
  26. [26]
    Axial and Equatiorial Bonds in Cyclohexane | MCC Organic Chemistry
    Since there are two equivalent chair conformations of cyclohexane in rapid equilibrium, all twelve hydrogens have 50% equatorial and 50% axial character.
  27. [27]
    Accurate Determination of the Structure of Cyclohexane by ...
    We combine femtosecond time-resolved rotational coherence spectroscopy with high-level ab initio theory to obtain accurate structural information for the ...
  28. [28]
    Conformational preferences in monosubstituted cyclohexanes ...
    Conformational preferences in monosubstituted cyclohexanes determined by nuclear magnetic resonance spectroscopy | Journal of the American Chemical Society.Missing: hydroxyl | Show results with:hydroxyl
  29. [29]
    [PDF] stereogenic center
    Cis-1,2-dimethylcyclohexane is an achiral molecule. Note that either of the cis-1,2-dimethylcyclohexane conformers is a configurational diastereomer (geometric.
  30. [30]
    [PDF] SUPPLEMENTARY NOTES FOR STEREOCHEMISTRY
    One chiral center renders the molecule chiral. H3C. CH3. H3C. CH3 cis-1,2-dimethylcyclohexane is an achiral molecule trans-1,2-dimethylcyclohexane is a chiral ...
  31. [31]
    Conformational Analysis. XVII.1 The 1,3-Diaxial Methyl-Methyl Interaction2
    **Key Findings on 1,3-Diaxial Methyl-Methyl Interaction in Cyclohexane:**
  32. [32]
    4.8 Conformations of Disubstituted Cyclohexanes - Organic Chemistry
    Sep 20, 2023 · In trans-1,2-dimethylcyclohexane, the two methyl groups are on opposite sides of the ring and the compound can exist in either of the two chair ...
  33. [33]
    trans-1,2-Dimethylcyclohexane
    View the gauche butane interaction (0.9 kcal/mol) as a Newman projection. The other chair conformation of trans-1,2-dimethylcyclohexane has the two methyl ...
  34. [34]
  35. [35]
    [PDF] Eliel E L. Conformational equilibria by nuclear magnetic resoisance ...
    Mar 19, 1982 · The conformational equilibrium between an equatorially and axially substituted cyclohexane. (here cyclohexyl bromide) is determined by ...
  36. [36]
    NMR-experiments on acetals—IVL : The conformational equilibrium ...
    NMR-experiments on acetals—IVL : The conformational equilibrium in some 4,4-disubstituted-1,1-dimethoxycyclohexanes. A challenge for the “additivity” principle☆.
  37. [37]
    nate stereochemistry as probed via 13C nmr chemical shifts and
    It has been shown, however (16), that additivity of. -AGO values breaks down in 1,l-disubstituted cyclohexanes, especially when hydrogen bonding can occur ...
  38. [38]
  39. [39]
    [PDF] Stereochemistry of Organic Compounds
    These are compounds confined to a single conformation in which the substituent X is either strictly equatorial or strictly axial. Such confinement may be ...
  40. [40]
    Conformational Properties of 1-Halogenated-1-Silacyclohexanes ...
    It was found by NBO analysis that the axial conformers are unfavorable in terms of steric energy and conjugation effects and that they are stabilized mainly by ...Missing: review | Show results with:review
  41. [41]
    [PDF] Redalyc.Enthalpic and Entropic Contributions to the Conformational ...
    A-values (differences in free energy, –∆G°, between the axial and equatorial conformations of monosubstituted cyclo- hexanes) are of great interest to chemists ...
  42. [42]
    Cycloalkanes with one heteroatom
    As one might expect, the tetrahydropyran ring adopts a chair conformation and is largely similar to cyclohexane; thus the energy barrier for ring flipping is ...Missing: value | Show results with:value
  43. [43]
    The Effect of Electrostatic Interactions on Conformational Equilibria ...
    The trisubstituted system preferred to adopt a half-chair (3H4) conformation with three axial substituents. These studies provided additional insight that ...<|separator|>
  44. [44]
    The Axial and Equatorial Hydrogen Bonds in the Tetrahydropyran ...
    Feb 16, 2001 · The axial form for THP⋅⋅⋅HCl and the equatorial one for PMS⋅⋅⋅HCl were found to be the most stable conformers. These conformational preferences ...
  45. [45]
    4.2: A-values and Equilibrium Ratios - Chemistry LibreTexts
    May 20, 2021 · A-values tell you the energetic preference for a substituent in the equatorial position. If we take a close look at these numbers, we see a few peculiarities.
  46. [46]
    From Cyclohexane to 2-Hydroxy-3-oxanone: A Conformation Study
    The C−C bond length in the chair conformer is found to be 1.551 Å, and the CCC angle calculated here, 111.8°, is very similar to the value found by electron ...
  47. [47]
    Experimental and Theoretical Study of the OH-Initiated Degradation ...
    Mar 29, 2024 · The conformational distribution in piperidine is calculated to be ∼82% eq and 18% ax at 200 K, ∼73% eq and 27% ax et 298 K, and ∼67% ax and 33% ...
  48. [48]
    The Conformational Preference of the Nonbonding Electron Pair in ...
    The Conformational Preference of the Nonbonding Electron Pair in Piperidine ... Steric manipulation of the lone pair in piperidine. Tetrahedron Letters 1970 ...
  49. [49]
    Cyclohexene - an overview | ScienceDirect Topics
    The half-chair conformation of cyclohexene possesses only a C2 axis bisecting the double bond and is chiral (point group C2). The two enantiomeric half-chairs ...
  50. [50]
    Unexpected Diastereomer Formation and Interconversions in ...
    Mar 27, 2023 · As part of this work, we evaluated use of cyclohexane-1,2-diacetal (CDA protection), aware of the regioisomeric protections (of bis-equatorial ...
  51. [51]
    Comparing the Performances of Force Fields in Conformational ...
    Apr 27, 2022 · Here, we compare the relative performances of different force fields for conformational searching of hydrogen-bond-donating catalyst-like molecules.
  52. [52]
    A vibrational study of cyclohexane and some of its isotopic ...
    Vibrational assignments are made with the aid of recent infrared solid state studies on cyclohexane. Suggestions are put forward for all active fundamentals in ...Missing: primary | Show results with:primary
  53. [53]
    Sachse
    In diamond Mohr also saw the fused chair cyclohexane structure of decalin: Surprisingly, full recognition of Sachse's premonitions would come only after half a ...
  54. [54]
    Historical Background to Conformational analysis - Ch.imperial
    ... Sachse in 1890 using only trigonometry, derived the non-planar chair and boat conformations of cyclohexane and the axial and equatorial positions of the ...
  55. [55]
    The conformation of the steroid nucleus | Cellular and Molecular Life ...
    Cite this article. Barton, D.H.R. The conformation of the steroid nucleus. Experientia 6, 316–320 (1950). https://doi.org/10.1007/BF02170915. Download ...
  56. [56]
    Precision space‐filling atomic models - Koltun - Wiley Online Library
    First published: December 1965 ; Citations · 151 ; The Corey-Pauling-Koltun Models to construct macromolecules soon will be available for research and teaching.
  57. [57]
    Nobel Prize in Chemistry 1969
    ### Summary of Odd Hassel's Contributions to Cyclohexane Conformation
  58. [58]
    [PDF] The Structure of 1,2-Epoxy-cyclohexane - Acta Chemica Scandinavica
    The epoxide was purified by distillation. ELECTRON DIFFRACTION-DIAGRAMS. A number of electron diffraction diagrams of 1,2-epoxy-cyclohexane were taken, using ...
  59. [59]
  60. [60]
    Derek Barton – Facts - NobelPrize.org
    In the 1950s Derek Barton charted conformations for a number of substances with biological importance, such as bile acids, sex hormones, cortisone and ...Missing: analysis | Show results with:analysis<|separator|>
  61. [61]
    Journal of the American Chemical Society
    Nuclear Magnetic Resonance Line-Shape and Double-Resonance Studies of Ring Inversion in Cyclohexane-d11. Click to copy article linkArticle link copied! F. A. L. ...
  62. [62]
    Stereochemistry of carbon compounds. -- : Eliel, Ernest Ludwig, 1921
    Jul 11, 2019 · Stereochemistry of carbon compounds. -- ; Publication date: 1962 ; Topics: Chemistry, Organic, Stereochemistry ; Publisher: New York : McGraw-Hill.