Fact-checked by Grok 2 weeks ago

A value

In , the A-value is a quantitative measure of the difference (in kcal/) between the axial and equatorial positions of a in the conformation of monosubstituted , reflecting the energetic preference for the equatorial orientation due to reduced steric interactions. This value, often denoted as the conformational energy or steric parameter, allows chemists to predict the relative stability of conformers and the equilibrium ratios between them. A-values are determined experimentally through techniques such as NMR spectroscopy or measurements and vary significantly with and nature; for instance, a has an A-value of approximately 1.7 kcal/mol, favoring the equatorial position by a ratio of about 19:1 at , while a tert-butyl group exhibits a much larger A-value of around 5.0 kcal/mol, nearly locking the ring in the conformation where it is equatorial. show anomalously low A-values— at 0.24 kcal/mol and at 0.53 kcal/mol—due to effects that minimize 1,3-diaxial interactions despite their . These parameters are additive for disubstituted cyclohexanes under certain conditions, enabling the estimation of overall conformational preferences in more complex molecules. The concept of A-values is fundamental to conformational analysis of cyclic compounds. Tables of A-values for various s, compiled from empirical data, serve as essential references in stereochemical studies, highlighting trends such as increasing values with alkyl chain branching.

Introduction and Background

Definition and Measurement

In conformational analysis, the A-value quantifies the preference of a in a monosubstituted to adopt the equatorial rather than the axial position in the chair conformation, defined as the standard difference ΔG° (in kcal/mol) between these two conformers at , typically measured at 25°C. This value arises primarily from steric repulsions in the axial position, where the experiences 1,3-diaxial interactions with syn-axial atoms. The chair conformation of features alternating axial and equatorial positions for the substituents: axial bonds are oriented nearly parallel to the ring's vertical axis, pointing alternately up and down, while equatorial bonds lie approximately in the ring's plane, extending outward at angles closer to 109.5°. At , rapid chair-chair interconversion (on the order of 10^5 s^{-1}) averages the environments of axial and equatorial hydrogens, but substituents with significant A-values shift the equilibrium toward the equatorial conformer. A-values are experimentally determined using (NMR) spectroscopy to measure the conformational populations. In practice, the equilibrium constant K = \frac{[\text{equatorial}]}{[\text{axial}]} is obtained by integrating separate NMR signals for each conformer or by analyzing temperature-dependent differences. Low-temperature NMR, often below -80°C in solvents like CS_2 or mixtures, slows the ring inversion sufficiently to resolve distinct signals for axial and equatorial conformers, allowing direct quantification of their relative intensities. The A-value is calculated from the thermodynamic relation A = RT \ln K, where R = 1.987 cal mol^{-1} K^{-1} (or 0.001987 kcal mol^{-1} K^{-1}) is the gas constant and T is the absolute temperature in Kelvin; this equation derives from the standard free energy change for the axial-to-equatorial interconversion, where \Delta G^\circ = -RT \ln K and A = -\Delta G^\circ represents the positive energy penalty for the axial position. To obtain A from population percentages, first compute the mole fractions: if the equatorial conformer comprises percentage p_{\text{eq}} (as a decimal) and axial p_{\text{ax}} = 1 - p_{\text{eq}}, then K = p_{\text{eq}} / p_{\text{ax}}; substituting into the equation yields A. For example, in methylcyclohexane at 25°C (298 K), NMR studies show approximately 95% equatorial population (p_{\text{eq}} = 0.95, p_{\text{ax}} = 0.05), giving K = 19, \ln K \approx 2.944, and A = (0.001987 \times 298) \times 2.944 \approx 1.7 kcal/mol.

Historical Development

The foundations of the A-value concept were laid in the late 19th and early 20th centuries through studies on cyclohexane conformations. Hermann Sachse proposed the chair and boat forms of cyclohexane in 1890, an idea later supported by Ernst Mohr in 1918 using X-ray data from diamond structures to validate the chair as the stable form. These insights provided the structural basis for understanding substituent preferences in cyclic systems. The modern era of conformational analysis, which directly led to the development of A-values, began in the 1940s and 1950s with the work of Derek H. R. Barton and Odd Hassel. Barton introduced the application of conformational principles to predict reactivity in his 1950 paper on steroid conformations, emphasizing the energetic preference for equatorial substituents. Concurrently, Hassel's X-ray crystallographic studies from 1943 to 1947 confirmed the chair conformation of cyclohexane and its derivatives, quantifying steric interactions. Their contributions earned them the 1969 Nobel Prize in Chemistry for advancing conformational analysis. Early quantitative measurements of A-values, defined as the difference between axial and equatorial conformers, were obtained in the 1950s using techniques such as measurements, kinetic studies (e.g., esterification rates of cyclohexanols), and for enthalpic components. For instance, Ernest L. Eliel employed esterification rate ratios of cyclohexanols in 1957 to determine A-values for hydroxyl and related groups, establishing initial scales for substituent effects. Frederic R. Jensen's studies in the late 1950s focused on , using methods to measure the preference for equatorial methyl (A ≈ 1.7 kcal/), highlighting 1,3-diaxial interactions. These efforts built additive schemes for estimating energies in polysubstituted cyclohexanes by summing individual A-values, assuming minimal interactions between distant substituents. The term "A-value" was introduced and popularized by Ernest L. Eliel in his 1965 book Conformational Analysis, where he compiled tables of these free energy differences from experimental data. The marked a shift to (NMR) for more precise determinations, enabling low-temperature studies of conformer populations. Jensen's 1960 NMR work on ring inversion barriers complemented substituent studies, while subsequent NMR applications to confirmed earlier A-values and refined them through direct observation of axial-equatorial ratios. Eliel's comprehensive 1965 book Conformational Analysis standardized A-value compilations, integrating experimental data and promoting their use in predicting conformational preferences. By the 1970s, research evolved to address environmental influences, with Eliel demonstrating solvent-dependent variations in A-values through equilibrium studies in polar and nonpolar media, attributing changes to differential solvation of axial and equatorial forms. In the 1990s, computational methods provided validation, as quantum mechanical calculations and simulations reproduced experimental A-values for various substituents, enhancing understanding of polysubstituted systems and additive approximations.

Thermodynamic Basis

Free Energy Considerations

The A-value for a substituent in a monosubstituted represents the standard free energy , \Delta G^\circ = G_{\text{axial}} - G_{\text{equatorial}}, which quantifies the thermodynamic penalty for adopting the axial conformation over the equatorial one, primarily arising from unfavorable 1,3-diaxial interactions between the and the ring hydrogens. This \Delta G^\circ serves as a key parameter in conformational analysis, indicating the preference for the equatorial position in chair-like derivatives. The conformational equilibrium between axial and equatorial forms is governed by the Boltzmann distribution, where the population ratio of equatorial to axial conformers is expressed as \frac{[\text{equatorial}]}{[\text{axial}]} = e^{-\Delta G^\circ / RT}, with R as the and T as the absolute temperature in ; this relation directly links the free energy difference to the of conformers at . A-values display a general temperature dependence that is typically weak, with values often slightly increasing as temperature increases for substituents exhibiting a negative entropy change for the axial conformer, due to the influence of entropic effects that modulate the free energy difference across thermal conditions. For many non-polar substituents like alkyl groups, the entropic contribution is very small (\Delta S \approx 0), making A-values nearly temperature-independent. In historical and contemporary literature, A-values are commonly reported in kcal/mol for consistency with early thermodynamic measurements, though SI units of kJ/mol predominate in recent work, with the conversion factor of 1 kcal/mol \approx 4.184 kJ/mol facilitating comparisons. A conceptual diagram for the chair conformers of a monosubstituted depicts the equatorial form at a lower minimum, with the axial conformer higher by \Delta G^\circ (the A-value), and a low interconversion barrier enabling rapid flipping between states at ambient temperatures, thus maintaining without isolating individual conformers. The enthalpic component of this penalty stems largely from steric repulsions in the axial orientation.

Enthalpic and Entropic Components

The A-value, which quantifies the preference for an equatorial over an axial in monosubstituted cyclohexanes, can be decomposed into its enthalpic and entropic components via the relation \Delta G = \Delta H - T\Delta S, where \Delta G corresponds to the A-value. This decomposition reveals that the enthalpic term (\Delta H) dominates for most s, stemming primarily from steric repulsions in 1,3-diaxial interactions that mimic the gauche interaction energy of approximately 0.9 kcal/ per pair, arising from der Waals forces between the axial and the syn-axial hydrogens at positions 3 and 5. For the , \Delta H \approx 1.8 kcal/, accounting for two such interactions. The contribution (-T\Delta S) is typically minor but nonzero, often in the range of 0.1–0.5 kcal/mol at 298 K, with a negative \Delta S for the axial conformer due to reduced vibrational freedom in the ring and greater ordering of surrounding molecules induced by the steric crowding. Axial substituents constrain low-frequency ring modes and promote structured shells, leading to this entropy penalty. For many non-polar substituents like alkyl groups, the entropic contribution is very small (\Delta S \approx 0). To separate these components experimentally, variable-temperature NMR spectroscopy measures the K (ratio of equatorial to axial populations) across a range of temperatures; a van't Hoff plot of \ln K versus $1/T provides \Delta H from the slope (-\Delta H / R) and \Delta S from the y-intercept (\Delta S / R), assuming ideal behavior. Representative examples illustrate varying contributions: the t-butyl group exhibits an A-value of 4.9 kcal/mol that is nearly purely enthalpic (\Delta H \approx 5.0 kcal/mol), with negligible entropy effects (\Delta S \approx 0 cal/mol·K), as its bulkiness enforces strong steric dominance without significant rotational or differences. Polar solvents influence these components, particularly the enthalpic term, through hydrogen bonding that differentially stabilizes axial or equatorial orientations; protic solvents can reduce \Delta H for polar substituents by solvating the axial position more effectively, thereby lowering the effective A-value compared to nonpolar media.

Data and Examples

Table of Selected A-Values

The following table compiles selected A-values for common substituents in monosubstituted s, representing the conformational free energy difference (ΔG°) between axial and equatorial positions at 25°C, typically measured in nonpolar solvents such as or . These values provide a reference for the relative steric demands of substituents and are drawn from compilations of experimental primarily from NMR and studies.
SubstituentA-value (kcal/mol)Range/Variability (kcal/mol)Primary Reference
-CH₃1.71.68–1.74Eliel et al., 1994
-CH₂CH₃1.81.75–1.8Eliel et al., 1994
-CH(CH₃)₂2.22.15–2.25Eliel et al., 1994
-C(CH₃)₃4.9>4.5Eliel et al., 1994
-CH=CH₂1.51.35–1.68Eliel et al., 1994
-C≡CH0.40.2–0.5Eliel et al., 1994
-CF₃2.12.0–2.2Eliel et al., 1994
-F0.20.15–0.36Eliel et al., 1994
-Cl0.50.43–0.53Eliel et al., 1994
-Br0.40.2–0.7Eliel et al., 1994
-I0.40.4–0.5Eliel et al., 1994
-OH0.60.5–0.9 (solvent-dependent)Eliel et al., 1994
-OCH₃0.60.6–0.75Eliel et al., 1994
-OC₂H₅0.90.85–0.95Eliel et al., 1994
-OC(O)CH₃0.70.65–0.75Eliel et al., 1994
-NH₂1.21.2–1.7 (solvent-dependent)Eliel et al., 1994
-N(CH₃)₂2.12.0–2.2Eliel et al., 1994
-NO₂1.11.05–1.13Eliel et al., 1994
-CN0.20.17–0.24Eliel et al., 1994
-COOH1.21.2–1.35 (solvent-dependent)Eliel et al., 1994
-COOCH₃1.11.0–1.2Eliel et al., 1994
-C₆H₅3.02.8–3.0Eliel et al., 1994
-SH1.21.1–1.3Eliel et al., 1994
-SCH₃1.00.95–1.05Eliel et al., 1994
-SO₂CH₃2.52.4–2.6Eliel et al., 1994
The values in the table are approximate and derived from experimental measurements, with variability often arising from solvent effects, temperature, or measurement techniques such as low-temperature NMR spectroscopy. For polar substituents like -OH and -NH₂, A-values are higher in nonpolar due to reduced stabilization of the axial conformer compared to protic . Gas-phase A-values, computed or measured spectroscopically, tend to be higher for electronegative groups owing to the absence of solvent . The selection emphasizes alkyl, , oxygen-, -, and sulfur-containing groups, as well as common functional groups relevant to synthetic . A-values are applied additively for disubstituted cyclohexanes when substituents are non-interacting (e.g., 1,2-trans or 1,4), allowing estimation of the population of conformers via the ; however, significant 1,3-diaxial overlaps between bulky groups like -tBu and -CH₃ require adjustments for steric repulsion beyond simple summation. This approach aids in predicting conformational preferences that influence reactivity in applications such as nucleophilic substitutions or enzyme-substrate binding.

Factors Influencing A-Values

Polar solvents can influence A-values of polar substituents by altering energies, often reducing the preference for equatorial positions through differential stabilization of axial conformers. For instance, the A-value for the hydroxyl group (-OH) increases from 0.6 kcal/mol in to 0.9 kcal/mol in (CCl₄), highlighting how solvation in protic media mitigates steric and dipole-related penalties for axial orientations. This effect is particularly pronounced for groups capable of hydrogen bonding, where solvent polarity screens intramolecular interactions and favors more compact axial arrangements. Temperature variations also modulate A-values, typically leading to a linear decrease with increasing due to enhanced entropic contributions in the axial conformer. For the methyl substituent, the A-value diminishes from approximately 1.8 kcal/ at 0°C to 1.5 kcal/ at 100°C, reflecting reduced enthalpic barriers to ring inversion at higher temperatures. Such dependence arises from the of vibrational modes and conformational populations, allowing better of equilibrium shifts in dynamic systems. Substituent interactions further perturb A-values through electronic mechanisms. In heteroatom-containing groups, inductive and effects can influence axial via of adjacent bonds. For alkyl substituents, stabilizes the equatorial conformer by delocalizing σ-electrons into the ring framework, contributing to higher A-values as chain length or branching increases; this effect is evident in the progression from methyl (1.7 kcal/) to tert-butyl (4.9 kcal/). Steric bulk of substituents correlates with A-values through quantifiable parameters that aid in predictive modeling. Charton parameters (ν), derived from hydrolysis rates, provide a scalar measure of steric hindrance, enabling quantitative forecasts of A-values; for example, ν values >0.5 for bulky groups like isopropyl align with observed A-values around 2.2 kcal/mol. Isotope effects introduce subtle variations in A-values stemming from differences in zero-point energies and vibrational frequencies. The deuterium substituent exhibits an A-value of approximately 0.1 kcal/mol in cyclohexane, attributable to reduced vibrational amplitudes in C-D bonds compared to C-H, which slightly favors the axial position through secondary kinetic isotope influences. Computational approaches, such as (DFT), have been employed to determine gas-phase A-values, often yielding higher values for polar substituents than experimental solution data due to the absence of effects.

Applications

Predicting Conformational Preferences and Reactivity

A-values provide a quantitative basis for predicting the preferred conformation in substituted cyclohexanes by estimating the difference (ΔG) between conformers, where the equatorial position is favored due to reduced steric interactions. In monosubstituted cases, the equatorial conformer predominates with a population ratio given by e^{A/}, where R is the and T is temperature; for example, exists >95% in the equatorial form at 25°C given A = 1.74 kcal/. For 1,2-disubstituted cyclohexanes, conformational preferences differ between stereoisomers. In the trans , both substituents can occupy equatorial positions in one chair conformer (ΔG ≈ 0), while the alternative diaxial conformer incurs a penalty of approximately A_1 + A_2; thus, the diequatorial form dominates. For instance, trans-1,2-dimethylcyclohexane favors the diequatorial conformation by ΔG ≈ 3.48 kcal/ (using A_methyl = 1.74 kcal/), resulting in >99% population at . In the cis , one substituent must be axial in either chair, leading to a preference for the conformer where the larger-A is equatorial, with overall determined by the smaller A-value. The approximates the total as ΔG_total ≈ Σ A_i for all axial s, assuming no significant interactions between them; this holds well for vicinal or 1,4-s but reaches limits with 1,3-diaxial arrangements, where additional gauche-like penalties (e.g., ~3.7 kcal/ per H- pair) exceed simple . In complex systems like steroids, this model predicts ring fusions and orientations, such as the trans-decalin preference in , aligning observed structures with minimized axial penalties. A-values also inform reactivity by linking conformational populations to kinetic rates, particularly in substitution reactions. Axial substituents impose steric hindrance that slows SN2 processes relative to equatorial ones, as the transition state exacerbates 1,3-diaxial interactions; for example, an axial can reduce the observed rate by a factor of approximately e^{A/RT} (∼19-fold at 25°C) if the reactive conformer requires it to be axial, due to lower population of that form. However, exceptions arise from stereoelectronic effects, such as the in carbohydrates, where the axial orientation of electronegative groups like the anomeric in α-D-glucopyranose is stabilized by despite an A-value penalty of ~0.6 kcal/mol for , overriding steric preferences in ~36% of the equilibrium mixture. Computational methods validate these predictions, with (MM) and (DFT) calculations reproducing experimental A-values within 0.1-0.5 kcal/mol for common substituents like methyl and tert-butyl. For instance, B3LYP/6-31G(d) DFT yields A_methyl = 1.72 kcal/mol, closely matching the experimental 1.74 kcal/mol, while MM force fields like MMFF94 achieve similar accuracy for conformer energies in disubstituted systems.

Estimating Intramolecular Interactions

A-values provide a useful for estimating the strength of gauche interactions in acyclic alkanes, particularly for alkyl substituents. In , the axial conformer experiences two 1,3-diaxial interactions, each approximately equivalent to a gauche interaction, leading to an A-value of about 1.7 kcal/ for the . This implies that the energy of a single gauche interaction is roughly half the A-value, or approximately 0.85 kcal/, allowing A-values to serve as a for such torsional in non-cyclic systems. In rigid polycyclic hydrocarbons like , A-values help proxy the energy of 1,3-interactions by relating substituent steric demands to known preferences, as adamantane's positions mimic fixed axial-like environments without conformational flexibility. calculations of steric substituent constants in adamantane derivatives confirm that these values correlate closely with cyclohexane A-values, enabling estimation of strain from substituent crowding in the cage structure. Similarly, in systems, A-values approximate the differential steric strain between and positions, where endo substituents encounter enhanced 1,3-interactions akin to diaxial repulsions. A-values are incorporated into the parameterization of force fields such as MM2 and MM3 to refine van der Waals terms, ensuring accurate reproduction of conformational energies. These fields use experimental A-values, along with other thermodynamic data, to fit non-bonded interaction parameters, particularly for alkyl and polar substituents, achieving errors below 0.5 kcal/ in predicted steric preferences. This integration allows A-values to inform broader simulations of intramolecular steric forces beyond . In coordination chemistry, A-values approximate chelate in metal complexes by quantifying steric penalties from substituents in five- or six-membered rings, where axial-like positions parallel interactions. For instance, bulky alkyl groups on chelating ligands increase strain energies, estimated via summed A-values, influencing complex stability and geometry. Differential A-values between hydroxy (-OH, ~0.9 kcal/mol) and alkoxy (-OR, ~0.6 kcal/mol) substituents reveal intramolecular hydrogen bonding contributions, as the lower A-value for -OH reflects stabilization from H-bonding in the axial position, providing an energetic estimate of ~0.3 kcal/mol for such interactions in constrained systems. Quantitatively, the arising from steric interactions can be approximated as the derivative of the change with respect to , F \approx \frac{d(\Delta G)}{dr}, where \Delta G is the A-value representing the penalty for close approach; however, , the A-value itself simplifies as the effective barrier for such forces in molecular modeling.

Limitations and Extensions

and Inaccuracies

The A-value model fundamentally assumes a rigid conformation for the cyclohexane ring, neglecting contributions from higher-energy forms such as and twist-boat conformers. While the form dominates with over 99.99% population at due to the twist-boat being approximately 5.5 kcal/mol higher in energy, this approximation introduces negligible error (<1%) for most monosubstituted cases. However, for bulky substituents like tert-butyl (A ≈ 4.9 kcal/mol), the increased strain can elevate the population of non-chair forms, leading to significant inaccuracies in predicted conformational equilibria. A key inaccuracy arises from the model's basis in monosubstituted cyclohexanes, where additivity of fails in polysubstituted systems due to unaccounted inter-substituent interactions. In , for instance, the diaxial conformation experiences an additional 1,3-diaxial methyl-methyl repulsion of about 3.7 kcal/mol beyond the sum of two individual methyl (1.7 kcal/mol each), resulting in an overestimation of the axial conformer's stability by 3-5 kcal/mol if additivity is assumed. This non-additive steric crowding is particularly pronounced in 1,3-cis arrangements, invalidating simple summation for reactivity predictions in such derivatives. Standard A-values are derived under uniform conditions (typically non-polar solvents at 25°C), overlooking variations with temperature and solvent polarity that can alter enthalpic and entropic components. For fluorine-substituted cyclohexane, the A-value is ≈0.15–0.25 kcal/mol in solution (favoring equatorial) but ≈ –0.3 kcal/mol in the gas phase (favoring axial) due to reduced solvation allowing dipole and hyperconjugative effects to stabilize the axial orientation, introducing errors up to ≈0.4 kcal/mol in phase-specific analyses. Such dependencies highlight the model's limitation in polar or high-temperature environments, where solvation effects modulate conformational preferences. The model primarily captures steric bulk, neglecting electronic contributions like hyperconjugation and polarizability, which influence substituent preferences especially in conjugated systems. For aromatic groups such as phenyl (A ≈ 2.8 kcal/mol), hyperconjugative interactions between the ring π-system and axial C-H bonds stabilize the equatorial position beyond pure sterics, while polarizability effects in larger aromatics like naphthyl lead to underestimation of equatorial bias by 0.2-0.5 kcal/mol if ignored. These omissions reduce accuracy for electronically tuned substituents in synthetic applications. Experimental determination of A-values carries inherent errors from measurement techniques like NMR or calorimetry, with precision typically limited to ±0.1 kcal/mol due to temperature control and equilibrium resolution challenges. Additionally, the rigid-ring assumption overestimates preferences in flexible or highly substituted systems, where ring distortion or pseudorotation can lower effective A-values by up to 20% for large groups, as vibrational averaging blurs strict axial-equatorial distinctions.

Advanced Considerations and Modern Uses

Computational enhancements have significantly advanced the accuracy and scope of A-value predictions beyond traditional experimental measurements. Modern density functional theory (DFT) calculations provide reliable estimates of conformational energies in monosubstituted cyclohexanes, reproducing experimental A-values within approximately 0.2–0.3 kcal/mol for common substituents like methyl and tert-butyl. These methods account for electronic effects and steric interactions more precisely than earlier semi-empirical approaches, enabling the evaluation of complex systems where experimental data is scarce. Emerging machine learning models trained on quantum chemical datasets show promise for predicting conformational preferences of novel substituents by leveraging graph neural networks to correlate molecular structures with energies. Extensions of A-values to non-cyclohexane systems have broadened their utility in conformational analysis. In piperidines, adapted A-values for substituents at the 4-position, such as -ΔG° ≈ 0.4 kcal/mol for the NH group favoring equatorial orientation, differ slightly from cyclohexane due to heteroatom effects but follow similar steric hierarchies. For decalins, A-values are applied to predict substituent preferences in fused rings, where trans-decalin rigidity amplifies axial penalties compared to isolated cyclohexanes, guiding the analysis of bicyclic natural products. In supramolecular chemistry, A-values quantify steric bulk in host-guest binding, with larger values correlating to reduced association constants for bulky guests in cyclodextrin or calixarene hosts by imposing conformational constraints on the complex. Modern applications leverage A-values for targeted molecular design. In drug discovery, A-values inform conformer selectivity in cyclohexane-containing pharmaceuticals, such as optimizing equatorial substituents in kinase inhibitors to enhance binding affinity and reduce off-target effects in bioactive conformers. For asymmetric synthesis, they predict diastereoselectivity in cyclohexane-based reactions, where differential A-values between substituents dictate the energetic favorability of transition states, achieving >95% diastereomeric excess in enolate additions to derivatives. Recent developments address evolving chemical spaces. Literature from the 2020s reports A-values for fluorinated groups, such as 0.15-0.30 kcal/mol for -CF3 in cyclohexane, highlighting reduced steric demand compared to hydrocarbons due to polar flattening effects. Integration with quantitative structure-activity relationship (QSAR) models uses A-values as steric descriptors to forecast toxicity, improving predictions of LD50 values for substituted cyclohexanes by incorporating conformational penalties into multivariate regressions. These advances address gaps in traditional A-value tables, which often omit bioisosteres like fluorinated or heterocyclic analogs. Hybrid experimental-computational workflows combine low-temperature NMR with DFT validations to derive reliable A-values for such groups, enhancing predictive power in .

References

  1. [1]
    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.
  2. [2]
    A Value - Chemistry LibreTexts
    Feb 28, 2022 · The A value of a ligand is a measure of steric strain it causes when interacting with a given ligand (the higher the A value, the more severe the steric strain ...
  3. [3]
    [PDF] Conformational Energies (A-Values) - MSU chemistry
    Eliel, E. L.; Wilen, S. H.; Stereochemistry of Organic Compounds, Wiley, 1993, p. ... (a) The A-values for OSiR3 are unusually sensitive to solvent. These ...
  4. [4]
    [PDF] Table of A-Values
    X. A-value kcal/mol. F. 0.24. SMe. 1.0. CH3. 1.8. Cl. 0.4. SPh. 1.2. CH2CH3. 2.0. Br. 0.2-0.7. S(O)Me. 1.2 i-C3H7. 2.2. I. 0.4. S(O)2Me. 2.5 c-C6H11.
  5. [5]
    Conformational preferences in monosubstituted cyclohexanes ...
    Conformational preferences in monosubstituted cyclohexanes determined by nuclear magnetic resonance spectroscopy | Journal of the American Chemical Society.
  6. [6]
  7. [7]
  8. [8]
  9. [9]
    Conformational analysis : Ernest L. Eliel - Internet Archive
    Sep 29, 2018 · Publication date: 1965 ; Topics: Conformational analysis, Chemistry, Organic, Chemistry, Physical, Konformationsanalyse, Konformationsanalyse, ...
  10. [10]
    Enthalpic and Entropic Contributions to the Conformational Free ...
    A-values (differences in free energy, −ΔG°, between the axial and equatorial conformations of monosubstituted cyclohexanes) are of great interest to chemists ...
  11. [11]
  12. [12]
    The Journal of Organic Chemistry - ACS Publications
    Thermodynamics of the Axial ⇌ Equatorial Conformational Equilibria of tert-Butylcyclohexane and tert-Butyl-Substituted Six-Membered Heterocycles. Theoretical ...
  13. [13]
    Solvation effect on cyclohexane - Chemistry Stack Exchange
    May 24, 2013 · The water molecules are having some "Solvation effect" on the methylcyclohexane that prevents the interconversion.
  14. [14]
    Conformational analysis. Part 36.1 A variable temperature 13C NMR ...
    The conformational equilibria in six methylcycloalkanes have been investigated using the temperature dependence of the 13C chemical shifts at temperatures above ...
  15. [15]
    Making Sense of Heteroatom Effects in π–π Interactions
    Aug 21, 2025 · Heteroatoms can profoundly impact the strength and preferred geometry of parallel and T-shaped interactions between aromatic systems.
  16. [16]
    Hyperconjugation: A More Coherent Approach - ACS Publications
    May 15, 2012 · This article is a discussion of hyperconjugation and a proposal for introducing the idea within the framework of other important concepts such as conjugation ...
  17. [17]
    Estimating Effective Steric and Electronic Impacts of a Ferrocenyl ...
    Feb 14, 2021 · Figure 7. Correlations of the Tolman cone angles and the cone angles in (phosphine)AuCl with the %Vbur parameters in the gold complexes.
  18. [18]
    Physicochemical Properties of 20 Ionic Liquids Prepared by the ...
    Apr 15, 2024 · In summary, this study characterized 20 ionic liquids prepared by the CBILS process, which encompassed their density, refractive index, ...
  19. [19]
    A Simple Equation for the Solubility of a Solid in a Supercritical Fluid ...
    A simple equation for calculating the solubilities of solid substances in gaseous mixtures of a supercritical fluid with another supercritical one or an inert ...
  20. [20]
    Conformational studies of steroids: Correlations with biological data
    The most active naturally occurring steroids are observed to possess the greatest degree of conformational flexibility. Significantly different conformers of a ...
  21. [21]
    The anomeric and exo-anomeric effects of a hydroxyl group and the ...
    The magnitude of the anomeric effect for the hydroxyl group was estimated to be 2.0 kcal/mol. The hydroxyl group in the axial and equatorial position exhibits ...
  22. [22]
    Conformational energies of reference organic molecules
    Intrinsic conformational preferences of substituted cyclohexanes and tetrahydropyrans evaluated at the CCSD(T) complete basis set limit: implications for ...
  23. [23]
  24. [24]
    Establishment of Steric Substituent Constants in the Adamantane ...
    Aug 10, 2025 · [reaction: see text] Steric substituent constants, S(A), of both alkyl and aryl substituents were calculated in the adamantane system by ...
  25. [25]
    [PDF] Steric Effect at C-7 of Norbornene - MacSphere
    FACIAL STRAIN. 6. "B" strain ("back" strain) is the increase in the steric strain resulting from compression and bond bending within the groups attached to an ...
  26. [26]
    Calculating Free-Energy Differences Using an Average Force
    Sep 23, 2025 · The purpose of this tutorial is to get the reader familiarized with the calculation of a free-energy change along a reaction-coordinate (RC) ...Potential Of Mean Force · Figure 2 · Figure 5
  27. [27]
    FTIR Investigation of the Fluorocyclohexane Ring Inversion in Liquid ...
    Solvent Effects on ΔS°. The vapor-phase value of ΔS° is the intrinsic entropic characteristic with which our solution ...
  28. [28]
    [PDF] Anomeric Effect, Hyperconjugation and Electrostatics
    Jul 11, 2021 · Note that the endocyclic O atom decreases positive charge at the axial α-C-H atom (in agreement with a hyperconjugative donation from the oxygen ...
  29. [29]
    Conformational Properties of 1-Halogenated-1-Silacyclohexanes, C ...
    As a rule, in monosubstituted cyclohexanes, the substituent prefers the equatorial position of the chair conformation. Rare exceptions are substituents having ...
  30. [30]
    Calculating cyclohexane A-values · The DFT Course - Nathan
    The A-value of a substituent is the energy of the axial cyclohexane conformer relative to the equatorial conformer (i.e., the isomerization energy).<|control11|><|separator|>
  31. [31]
    Understanding Conformation Importance in Data-Driven Property ...
    Mar 18, 2025 · This study investigates the influence of using multiple conformers in machine learning-based property prediction, comparing two- and three- ...
  32. [32]
    Conformational analysis. Part 16. Conformational free energies in ...
    The conformational free energies (-delta G degree) of a number of 4-substituted piperdines and piperidinium salts have been determined by the J-value method.
  33. [33]
    Fused Rings: Cis and Trans Decalin - Master Organic Chemistry
    Aug 5, 2014 · Each six membered ring will adopt a chair conformation. Believe it or not, the cis and trans stereoisomers of decalin have remarkably different ...
  34. [34]
    Understanding the Effects of Pre-Organization, Rigidity, and Steric ...
    For the most rigid hosts, 4a-c, steric interactions played a direct role in the guest binding affinities with the bulkiest host having the lowest affinity and ...
  35. [35]
    A Data-Driven Perspective on Bioisostere Evaluation
    (46) The BioSTAR workflow was designed for use by medicinal chemists with no previous computational experience, requiring as input solely the structure of the ...
  36. [36]
    Control of Diastereoselectivity in Tandem Asymmetric Reactions ...
    We report the manipulation of hydrogen-bonding-based cooperative catalysis to control the diastereoselectivity for a tandem asymmetric reaction creating two ...
  37. [37]
    The Emergence and Properties of Selectively Fluorinated 'Janus ...
    Apr 4, 2023 · For monosubstituted cyclohexanes the equatorial conformer is usually lower in energy as this avoids less favoured 1,3-diaxial interactions with ...<|control11|><|separator|>
  38. [38]
    The predictivity of QSARs for toxicity - ScienceDirect.com
    QSARs and their predictions can be assessed against a number of criteria for their potential use as surrogates for animal, or other, tests.