Fact-checked by Grok 2 weeks ago

Hyperconjugation

Hyperconjugation is a fundamental concept in describing the delocalization of electrons from filled σ orbitals, typically C–H or C–C bonds, into adjacent empty or partially filled p or π* orbitals, which stabilizes molecular structures such as carbocations, free radicals, and unsaturated systems through partial orbital overlap and charge redistribution. This interaction, often visualized as "no-bond" resonance structures in or as second-order orbital mixing in , enhances at electron-deficient centers and influences bond lengths, vibrational frequencies, and conformational preferences. The phenomenon was first observed experimentally in the Baker–Nathan effect of 1935, where alkyl substituents showed an unexpected order of stabilizing influence (methyl > ethyl > isopropyl > tert-butyl) in solvolysis rates of diphenylmethyl halides, attributed to varying degrees of σ-π hyperconjugation. Robert S. Mulliken formalized the term "hyperconjugation" in 1939, linking it to electron release from saturated bonds to adjacent unsaturated or deficient sites, a concept later validated by isotope effects (e.g., secondary k_H/k_D effects in relevant reactions) and spectroscopic data. By the 1960s, its role gained widespread recognition, including in work on reaction mechanisms, underscoring its ubiquity beyond simple alkenes to include applications in stereoelectronic effects and stabilization. In modern understanding, hyperconjugation provides significant stabilization, often comparable in certain contexts to traditional π-conjugation, driving key structural features like the staggered conformation of (total stabilization ≈ 3–4 kcal/mol from hyperconjugation) and the in carbohydrates, where lone-pair-to-σ* donation favors axial substituents. It also governs reactivity patterns, such as enhanced stability of secondary over primary carbocations via multiple C–H σ donations and regioselectivity in electrophilic additions to alkenes, where hyperconjugative assistance lowers barriers. Computational tools like (NBO) analysis have quantified these effects, revealing hyperconjugation's contributions to bond weakening (e.g., allylic C–H bonds) and its extension to negative hyperconjugation in hypervalent species with electronegative atoms.

Fundamentals

Definition and Concept

Hyperconjugation is a fundamental concept in involving the delocalization of electrons from filled σ-orbitals, typically those of C-H or C-C bonds, into adjacent unoccupied orbitals such as empty p-orbitals or π* antibonding orbitals. This interaction stabilizes molecules by distributing and often imparts partial double-bond character to the participating σ-bonds, influencing and reactivity. Unlike traditional π-conjugation, hyperconjugation bridges saturated and unsaturated systems through σ-π or σ-p interactions. To understand hyperconjugation, it is essential to recall the basics of σ- and π-bonding. σ-bonds form from the end-on overlap of orbitals along the internuclear , creating a symmetric distribution, while π-bonds arise from lateral overlap perpendicular to this , resulting in nodal planes. These bonding types provide the orbital framework for hyperconjugation, where the higher-energy filled σ-orbitals donate density to lower-energy acceptor orbitals. Key characteristics of hyperconjugation include its permanent nature, as it occurs continuously in molecules possessing suitably aligned adjacent bonds, without requiring specific excitation or reaction conditions. It is commonly depicted using no-bond structures in , which illustrate the delocalization by showing formal "breaking" of a σ-bond to form a π-bond, though no actual bond cleavage occurs. Effective overlap demands geometric alignment, such as antiperiplanar orientation of the donating σ-bond relative to the acceptor orbital, often involving bonds in the α-position to unsaturated or electron-deficient centers. A simple illustration of hyperconjugation is found in the ethyl cation (CH₃CH₂⁺), where the vacant p-orbital on the positively charged carbon accepts from the σ C-H bonds of the adjacent . This leads to hybrids: the primary structure features the empty p-orbital perpendicular to the C-C bond, while contributing no-bond resonance forms show donation from a C-H σ-orbital, creating partial C=C double-bond character and distributing the positive charge toward the hydrogens. Such delocalization significantly stabilizes the .

Historical Development

The concept of hyperconjugation originated in the late as an extension of theory to explain electron delocalization involving σ bonds adjacent to π systems or empty p orbitals. Although G. N. Lewis's 1916 introduction of the shared laid groundwork for ideas of partial character, the term "hyperconjugation" was formally coined by Robert S. Mulliken in to describe the interaction of σ from C-H bonds with π in cyclic dienes, based on observed intensities in spectra. Earlier observations, such as the anomalous ordering of alkyl effects in rates (methyl > ethyl > isopropyl > tert-butyl), reported by J. W. and W. S. Nathan in 1935, were retrospectively attributed to hyperconjugative stabilization and became known as the Baker-Nathan effect. During the , hyperconjugation gained traction as an explanation for electron donation by alkyl groups. By 1944, George W. Wheland integrated hyperconjugation into in his seminal book The Theory of Resonance and a companion paper on hydrocarbons, treating it as a phenomenon akin to "no-bond doubles" that contributes to molecular . Mulliken further developed the orbital delocalization model between and 1942, emphasizing its role in UV and conjugation effects. The saw debates over hyperconjugation's relative importance versus inductive effects in alkyl influences, with some chemists questioning its magnitude in saturated systems and favoring or steric arguments. Mulliken's comprehensive 1949 review on conjugation and hyperconjugation helped solidify its , highlighting quantitative estimates from data. Experimental milestones in the included NMR studies revealing spin-density distributions consistent with hyperconjugative delocalization in conjugated ligands. By the 1970s, using methods confirmed hyperconjugative energies, while photoelectron provided direct validation of orbital mixing, countering early dismissals of "no-bond resonance" as artifactual. These advances led to widespread acceptance of hyperconjugation as a in .

Mechanism and Types

Orbital Interactions

Hyperconjugation arises from the delocalization of s through the partial overlap of a filled σ bonding orbital, typically a C-H or C-C , with an adjacent empty or low-lying acceptor orbital such as a p orbital on a center or the antibonding π* orbital of a . This interaction requires proper geometric alignment, where the donor σ orbital is oriented parallel to the acceptor orbital to enable (lateral) overlap, maximizing the interaction efficiency. Additionally, effective hyperconjugation demands close matching of the orbital energies between the donor and acceptor to facilitate significant electron delocalization without excessive energetic penalty. In molecular orbital theory, the interaction between the filled σ donor and empty acceptor orbital results in the formation of new delocalized molecular orbitals: a bonding combination that is stabilized by lowering the overall energy of the system and an antibonding counterpart that is raised in energy. The stabilization energy can be quantified using natural bond orbital (NBO) analysis, which decomposes the interaction into second-order perturbation terms. The hyperconjugative stabilization is given by the expression E^{(2)} = \frac{q_i F_{ji}^2}{\epsilon_j - \epsilon_i} where q_i is the occupancy of the donor orbital, F_{ji} is the Fock matrix element between the donor (i) and acceptor (j) orbitals, and \epsilon_j - \epsilon_i is the energy difference between the acceptor and donor orbitals; larger E^{(2)} values indicate stronger interactions. A representative visualization of this σ-π* overlap occurs in propene (CH₃-CH=CH₂), where the in-plane C-H σ orbitals of the methyl group align with the π* antibonding orbital of the C=C double bond, leading to delocalization that mixes the σ bonding character into the π system and stabilizes the molecule. This can be depicted schematically as the sideways overlap between a filled σ_{C-H} orbital and the empty π* orbital, forming a four-center delocalized system; accompanying energy diagrams show the lowering of the bonding MO energy relative to the isolated orbitals, with the extent of mixing dependent on the overlap integral. The strength of hyperconjugative interactions is modulated by several factors, including bond angle and torsional conformation, where eclipsed or bisected arrangements position the σ donor orbital for optimal parallel overlap with the acceptor, enhancing delocalization compared to staggered geometries. Electronegativity differences also play a role, as evidenced by the isotope effect: C-H bonds exhibit stronger hyperconjugation than C-D bonds due to the slightly higher σ orbital energy and better donor ability of C-H, arising from differences in zero-point vibrational energies that affect orbital overlap and . Furthermore, the vicinal (adjacent) positioning of the donor and acceptor orbitals is essential, as hyperconjugation primarily involves interactions between bonds separated by one intervening atom, limiting its range to neighboring functional groups.

Positive and Negative Hyperconjugation

Hyperconjugation manifests in two primary modes—positive and negative—distinguished by the direction of electron delocalization and the nature of donor-acceptor interactions. Positive hyperconjugation involves the of from a filled σ orbital, typically a C-H or C-C bond, to an adjacent empty acceptor orbital, such as an unfilled p orbital in carbocations or a π* orbital in alkenes. This process stabilizes electron-deficient systems by dispersing positive charge or reducing the energy of the lowest unoccupied (LUMO). A classic example is the isopropyl cation ((CH₃)₂CH⁺), where the empty p orbital on the central carbon accepts density from the six equivalent C-H σ bonds of the methyl groups, leading to structures that depict partial double-bond character in the C-C linkages and hydrogen migration. This delocalization contributes significantly to the relative stability of secondary carbocations over primary ones. In contrast, negative hyperconjugation entails the donation of electron density from a filled π or lone-pair (n) orbital to an adjacent σ* antibonding orbital, often resulting in the weakening or elongation of the donor bond while stabilizing the acceptor. This mode is prominent in anions or molecules with electronegative substituents, where the interaction transfers density away from a high-electron-density region. For instance, in the fluoromethyl anion (F-CH₂⁻), fluorine's lone pairs donate into the σ* orbital of the adjacent C-F bond, enhancing stability without net charge buildup on carbon. Similarly, in difluoromethane (F-CH₂-F), mutual n_F → σ_C-F interactions lead to elongation of the C-F bonds compared to monofluoromethane, as the delocalization populates the antibonding orbital. This effect underlies the anomeric effect in carbohydrates and related systems, where axial electronegative substituents are preferred due to enhanced lone-pair donation to σ orbitals. Positive hyperconjugation predominates in carbocations and radicals, where it disperses positive charge through σ → p or σ → π* interactions, whereas negative hyperconjugation is more relevant in anions and hypervalent molecules, facilitating charge accommodation via n → σ* or π → σ* pathways. Both modes are quantified through delocalization energies derived from (NBO) analysis or block-localized wavefunction (BLW) methods, typically ranging from 5–15 kcal/mol per interaction, though they produce opposing charge effects: positive hyperconjugation increases electron density at the acceptor site, while negative hyperconjugation depletes it from the donor. These energies highlight hyperconjugation's role as a pervasive stabilizing force, comparable in magnitude to traditional conjugation in many cases. Rare variants include homo-hyperconjugation, where σ orbitals interact with adjacent σ* orbitals in saturated alkanes, contributing to conformational preferences like the staggered structure through bidirectional delocalization.

Structural and Energetic Effects

Influence on Bond Lengths and Geometry

Hyperconjugation imparts partial double-bond character to adjacent σ bonds, resulting in shortened bond lengths compared to non-hyperconjugated systems. In propene, the C-C bond between the sp²-hybridized carbon and the measures approximately 1.50 , significantly shorter than the 1.54 C-C bond in , where no such delocalization occurs. This contraction arises from the overlap of the π* orbital of the C=C bond with the σ orbitals of the methyl C-H bonds, effectively increasing the of the C-C linkage. The donating σ C-H bonds in hyperconjugating methyl groups experience slight elongation due to partial depopulation of the σ orbital, weakening the bond. In the α-methyl group of propene, these C-H bonds are marginally longer than those in isolated methyl groups, as evidenced by (NBO) analyses that quantify the donor-acceptor interaction energies. This effect is more pronounced in systems with multiple α-hydrogens, correlating with the degree of delocalization. Hyperconjugation also induces geometric distortions by altering hybridization and torsional preferences. In carbocations, the interaction between adjacent C-H σ orbitals and the empty p orbital promotes pyramidal flattening at the cationic center, approaching planarity to maximize overlap and increase s-character in the affected bonds. Staggered conformations in alkanes and alkenes minimize torsional while optimizing hyperconjugative alignment, as the anti-periplanar orientation of C-H bonds to adjacent π systems or empty orbitals reduces overall molecular energy. Experimental observations confirm these structural changes. Microwave spectroscopy of propene reveals the contracted C-C bond length, while in conjugated hydrocarbons like 1,3-butadiene shows bond alternation patterns attributable to hyperconjugative contributions. detects shifts in C-H frequencies, with hyperconjugating bonds exhibiting lower wavenumbers due to elongation and reduced force constants. NBO calculations further link these geometries to hyperconjugative stabilization energies of 2-5 kcal/mol per interaction in systems.

Role in Molecular Stability

Hyperconjugation contributes to molecular stability by delocalizing from adjacent σ bonds into empty or antibonding orbitals, lowering the overall of the system. The stabilization per hyperconjugative typically ranges from 5 to 15 kcal/, depending on the system and the quality of orbital overlap. This effect scales directly with the number of available alpha hydrogens; for instance, in alkyl , a benefits from approximately nine such hydrogens, providing up to three times the stabilization of a primary carbocation with three alpha hydrogens, as evidenced by affinity measurements and computational analyses. The magnitude of hyperconjugative stabilization exhibits notable trends across conditions. It is generally greater in the gas phase than in , where solvent molecules can solvate the charged center and reduce the relative importance of intramolecular delocalization. Temperature influences the effect through conformational dynamics, as rotational averaging at higher temperatures may misalign σ bonds, diminishing optimal hyperconjugation in flexible molecules. In comparison to stabilization, which often exceeds 20 kcal/mol in π-conjugated systems like allyl cations, hyperconjugation offers weaker individual contributions but operates pervasively in non-conjugated hydrocarbons, cumulatively enhancing thermodynamic stability. Isotopic substitution provides direct evidence for hyperconjugation's role in stability. Replacing alpha hydrogens with weakens the σ donation due to the higher vibrational of C-D bonds, resulting in reduced stabilization. This manifests as normal secondary kinetic isotope effects in reactions involving intermediates, such as solvolysis of alkyl tosylates, where rate ratios k_H/k_D exceed 1 (often 1.05-1.15 per ), reflecting greater reactivity for protio compounds. Computational studies using (DFT) methods, such as B3LYP with basis sets like 6-31G* or 6-311G(d), alongside approaches, confirm hyperconjugation as the primary driver of stability in these systems. These calculations decompose energies to show that hyperconjugative delocalization accounts for 70-80% of the total stabilization in simple alkyl s, with the remainder attributable to inductive effects from alkyl substituents. (NBO) analyses in such computations quantify individual σ → p* interactions at 10-20 kcal/mol each in carbocation systems, underscoring their dominance over polar contributions.

Specific Applications

In Carbocations and Radicals

Hyperconjugation plays a crucial role in stabilizing by delocalizing the positive charge through interactions between the empty p-orbital and adjacent C-H σ-bonds. The increases with the degree of alkyl , following the order tertiary > secondary > primary, as more alpha hydrogens are available for hyperconjugative donation. For instance, the ethyl carbocation (CH₃CH₂⁺) has three alpha hydrogens, the isopropyl carbocation ((CH₃)₂CH⁺) has six, and the tert-butyl carbocation ((CH₃)₃C⁺) has nine equivalent alpha hydrogens, each contributing to structures that distribute the charge over the molecule. In the tert-butyl cation, these nine hyperconjugative structures form a hybrid, significantly lowering the energy compared to less substituted analogs. Computational studies quantify this stabilization, showing that the ethyl carbocation benefits from approximately 9 kcal/mol of hyperconjugative stabilization due to the three C-H bonds. This effect scales with substitution, making carbocations the most stable. The enhanced stability directly influences reactivity, as evidenced by SN1 solvolysis , where tertiary alkyl halides react orders of magnitude faster than secondary (relative ~30:1) and primary (~1000:1) counterparts, reflecting the lower for carbocation formation. In free radicals, hyperconjugation similarly stabilizes the species by delocalizing the into adjacent σ C-H bonds, though the effect is weaker than in carbocations due to the half-filled interacting less strongly with the filled σ-orbitals. The tert-butyl radical ((CH₃)₃C•), for example, benefits from nine alpha C-H bonds, following the same substitution trend as carbocations but with reduced overall stabilization energy. Electron spin resonance (ESR) confirms this delocalization, revealing large hyperfine splitting constants (typically 20-25 G) from alpha hydrogens in alkyl like ethyl and isopropyl, indicating significant spin density transfer via hyperconjugation.

In Alkenes and Conjugated Systems

Hyperconjugation plays a key role in stabilizing alkenes by allowing delocalization of from adjacent σ bonds, primarily C-H or C-C, into the antibonding π* orbital of the C=C . This σ → π* donation increases with the degree of substitution at the carbons, as more alkyl groups provide additional donor σ bonds. Consequently, tetrasubstituted alkenes exhibit greater than monosubstituted ones, with each additional alkyl enhancing the number of hyperconjugative interactions by up to six for a tetrasubstituted case compared to zero in ethene. This trend is quantitatively supported by heats of hydrogenation, which measure the energy released upon addition of H₂ across the double bond and inversely reflect alkene stability. For instance, the heat of hydrogenation of ethene is -32.8 kcal/mol, while that of trans-2-butene (a 1,2-disubstituted ) is -27.6 kcal/mol, demonstrating the ~5.2 kcal/mol stabilization from two methyl groups via hyperconjugation. Similarly, trisubstituted and tetrasubstituted alkenes show progressively lower (less exothermic) values, confirming the cumulative effect of multiple σ → π* interactions. In conjugated systems like dienes and polyenes, hyperconjugation complements π-π conjugation by extending electron delocalization through σ bonds adjacent to the unsaturated framework. In 1,3-butadiene, hyperconjugative interactions contribute to the overall stabilization energy of approximately 3.5 kcal/mol, as evidenced by its heat of of -57.1 kcal/mol compared to -60.6 kcal/mol expected for two isolated double bonds. The s-cis conformer benefits particularly from vicinal σ → π* donations that reinforce the conjugated array, leading to bond length alternation where the central C2-C3 bond shortens to ~1.48 due to partial double-bond character. Analogous effects occur in 1,3-butadiyne, where hyperconjugation contributes to delocalization across the cumulated triple bonds. Vicinal hyperconjugation is especially relevant in allenes and cumulenes, where the orthogonal π bonds of the cumulated double bonds interact with adjacent σ orbitals to favor linear geometries. In (H₂C=C=CH₂), hyperconjugative interactions help maintain the structure's rigidity. This effect extends to longer cumulenes like 1,2,3-butatriene, where hyperconjugation reinforces the cumulative linear chain, contributing to their characteristic rigidity and planarity in unsubstituted cases. Evidence for these hyperconjugative effects in alkenes comes from both spectroscopic and computational methods. UV spectroscopy reveals bathochromic shifts in absorption maxima (λ_max) for more substituted alkenes; for example, propene absorbs at ~188 nm compared to ethene's ~175 nm, reflecting a reduced -LUMO gap due to σ → π* donation raising the π energy. (NBO) analyses of propene confirm three primary C-H σ → π* interactions that weaken the adjacent C-H bonds and strengthen the C=C bond by delocalizing into the π* acceptor. These findings underscore hyperconjugation's role in modulating electronic properties without altering the basic π framework.

Rotational Barriers in Alkanes

In , the torsional barrier to rotation about the C–C bond is approximately 2.9 kcal/mol, as determined by of isotopic variants, with the staggered conformation preferred over the eclipsed one. This energetic preference arises primarily from hyperconjugative stabilization through six antiperiplanar σ C–H / σ* C–H interactions in the staggered conformer, compared to zero such interactions in the eclipsed form. A longstanding debate regarding whether steric repulsion or hyperconjugation dominates the barrier was addressed through computational analysis, which demonstrated that disabling hyperconjugative interactions reverses the conformational preference, favoring hyperconjugation as the key factor over Pauli repulsion. This hyperconjugative mechanism extends to larger alkanes, where rotational barriers exhibit similar σ-donation influences. In , the torsional barrier is about 3.4 kcal/, reflecting contributions from both methyl-hydrogen and methyl-methyl interactions modulated by hyperconjugation. For n-butane, the barrier between the and eclipsed conformations is approximately 3.8 kcal/, with hyperconjugative effects accounting for roughly one-third of the total, as shown by energy decomposition in calculations that isolate delocalization from steric components. Experimental evidence for the role of hyperconjugation in these barriers comes from , which provides precise barrier heights for and its isotopologues. Computational studies at the level further confirm that hyperconjugative stabilization outweighs Pauli repulsion in determining staggered preferences across alkanes. Isotopic substitution supports this view: the rotational barrier in CH₃CD₃ (2.87 kcal/mol) is lower than in CH₃CH₃ (2.90 kcal/mol), attributable to the reduced hyperconjugative donor ability of C–D bonds compared to C–H bonds.

References

  1. [1]
    Hyperconjugation: A More Coherent Approach - ACS Publications
    May 15, 2012 · Curiously, although the term 'hyperconjugation', meaning very large conjugation, was a misnomer in its original application to C–H and C–C–x ...Missing: credible | Show results with:credible
  2. [2]
    [PDF] Hyperconjugation
    This review outlines the ubiquitous nature of hyperconjugative interactions and their role in the structure and reactivity of organic molecules.Missing: article | Show results with:article
  3. [3]
    Intensities of Electronic Transitions in Molecular Spectra IV. Cyclic ...
    Mulliken; Intensities of Electronic Transitions in Molecular Spectra IV. Cyclic Dienes and Hyperconjugation. J. Chem. Phys. 1 May 1939; 7 (5): 339–352.
  4. [4]
    John William Baker and the origin of the Baker-Nathan effect
    This paper will first deal with the life of John William Baker, then examine how Baker and his collaborator, Wilfred Samuel Nathan, discovered what was ...
  5. [5]
    Conjugation and Hyperconjugation | PDF | Molecular Orbital - Scribd
    Rating 4.0 (1) and was introduced in the late 1930s ... lowing are exemplary: (1) The acidity of carboxylic See VALENCE. Matthew F. Schlecht acids stems from the conjugation ...
  6. [6]
    Hyperconjugation in Paraffin Hydrocarbons - AIP Publishing
    G. W. Wheland, J. T. Pinkston, Jr.; Hyperconjugation in Paraffin Hydrocarbons, The Journal of Chemical Physics, Volume 12, Issue 2, 1 February 1944, Pages 69,
  7. [7]
    [PDF] Robert S. Mulliken - Nobel Lecture
    Mulliken's Nobel work was for "fundamental work concerning chemical bonds and the electronic structure of molecules by the molecular-orbital method". A ...
  8. [8]
    Hyperconjugation effects of para-alkyl groups - ScienceDirect.com
    In the present paper the inductive effects of para-alkyl groups (Me, Et, Pri, But) in diverse reactivities of benzene derivatives are evaluated in this manner ...Missing: debates 1950s
  9. [9]
    Robert Sanderson Mulliken | Biographical Memoirs: Volume 78
    Robert S. Mulliken was a quiet, soft-spoken man, yet so single-minded and determined in his devotion to understanding molecules that he came to be called “Mr. ...
  10. [10]
    Spin-Density Distributions in Conjugated Ligands of ... - AIP Publishing
    Effects attributable to hyperconjugation have been observed and the conjugating abilities of linking groups C=C,. N =N, NH, 0, S, and S02 have been assessed. It ...
  11. [11]
    XVI. Photoelectron spectroscopy and molecular conformations: Ge ...
    The hyperconjugative ability of the C M(M = Si, C, Ge and Sn) bonding MO's with the ethylene π-MO is explained by simple perturbation theory. This result ...
  12. [12]
    Hyperconjugation* | Journal of the American Chemical Society
    Cross‐Hyperconjugation: An Unexplored Orbital Interaction between π‐Conjugated and Saturated Molecular Segments. Angewandte Chemie 2013, 125 (3) , 1017-1021 ...
  13. [13]
    A natural bond orbital analysis of carbanions - ScienceDirect.com
    ... perturbation energies: E(2) = ΔEij = qi(Fi,j)2/(ei − ej) where qi is the donor orbital occupancy; ei and ej are orbital energies; and Fi,j is the off ...
  14. [14]
    [PDF] Kinetic Isotope Effects in Organic Chemistry - Macmillan Group
    Sep 14, 2005 · C-H bonds vibrate with greater frequency and with greater amplitudethan do C-D bonds due to their higher. ZPE. This asymmetry in the vibrational ...
  15. [15]
    Anomeric effect, hyperconjugation and electrostatics: lessons from ...
    Aug 24, 2021 · We show that the complete hyperconjugative model remains superior in explaining the interplay between structure and reactivity. We will use ...
  16. [16]
    Hyperconjugation - 2019 - WIREs Computational Molecular Science
    Sep 6, 2018 · This review outlines the role of hyperconjugative interactions in the structure and reactivity of organic molecules.
  17. [17]
    Origin of β-agostic interaction in d0 transition metal alkyl complexes
    Jun 15, 2018 · An explanation based on negative hyperconjugation is put forward to explain the special case of C-H activation using d0-metal centers, which is ...
  18. [18]
    The C-C bond length in propene is little shorter 149 pm than the C-C ...
    Jan 21, 2020 · The C-C bond length in propene is little shorter 149 pm than the C-C bond length 154 pm in ethane. This is due to · ← Prev Question Next ...
  19. [19]
    Experimental data for C 2 H 6 (Ethane) - CCCBDB
    Bond length · Rotational Constant · Moment of Inertia · Dipole and ... H8, 108.000, H7, C2, H8, 108.000. Bond descriptions. Examples: C-C single bond, C=C ...
  20. [20]
    Bond lengths and bond energies in conjugation and hyperconjugation
    It is concluded that Dewar is partly right in supposing that the lengths of conjugated and hyperconjugated CC single bonds are determined, to a greater extent ...Missing: elongation | Show results with:elongation
  21. [21]
    Implications of hyperconjugative effects on bond lengths of allylic ...
    Therefore, we have applied natural bond orbital (NBO) calculations to identify the main orbital interactions in the ground state structures. Additionally, we ...Missing: paper | Show results with:paper
  22. [22]
    Hyperconjugation in Carbocations, a BLW Study with DFT ...
    Conjugation involves an interaction between π orbitals. It reputedly implies large stabilization energy (or resonance energy) and the effect extends across ...<|separator|>
  23. [23]
    Theoretical study of conjugation, hyperconjugation, and steric effect ...
    A pio- neering work dealing with the limited expansion of MOs has been made by Mulliken and Parr.6 To study the electronic delocalization versus localization, ...<|control11|><|separator|>
  24. [24]
    Microwave spectra for the three 13C1 isotopologues of propene and ...
    New measurements of microwave lines (A and E) of propene and its three 13C1 isotopologues have been made in the 10–22 GHz region with FT accuracy.
  25. [25]
    Chemical origin of blue and red shifts of C–H stretching vibrations in ...
    The NBO results suggest that both the red and blue shifts are attributed to changes of electron density in σ∗CH and s-character of carbon in C–H bond. Our ...
  26. [26]
    Exploring the Forces That Control the P−C Bond Length in ...
    The high n(O) → σ*(C−C) hyperconjugative interactions are reflected in the long R(C−C) bond length of 1.554 Å. Effect of Coordination with a Metal Center on the ...
  27. [27]
    Stabilization of carbocations CH 3 + , C 2 H 5 + , iC 3 H 7 + , tert-Bu ...
    Feb 13, 2017 · The isolated carbocations are stabilized due to the intramolecular charge distribution under the influence of hyperconjugation.
  28. [28]
    Hyperconjugation in Carbocations, a BLW Study with DFT ... - Frontiers
    The geometry of ethyl cation is discussed, and the hyperconjugation effect in carbocations is evaluated at the B3LYP/6-311G(d) level.
  29. [29]
    Theoretical Studies of Alkyl Radicals in the NaY and HY Zeolites
    The main focus is on the hyperfine interaction of alkyl radicals in the NaY and HY zeolites. The hyperfine splitting for neutral free radicals and free radical ...
  30. [30]
    Alkene Heats of Hydrogenation - Oregon State University
    Dec 6, 2021 · Ranking of alkene stability based on enthalpy of hydrogenation. Values (in kcal/mol) taken from the NIST Webbook and reflect more current values.
  31. [31]
    MOs of the 1,3-Butadiene System - Oregon State University
    Dec 15, 2019 · propene gives ΔH° = -60.6 kcal/mol; for 1,3-butadiene ΔH° = -57.1 kcal/mol. Somewhat restricted rotation about C2-C3.Missing: delocalization | Show results with:delocalization
  32. [32]
    Coarctate and Möbius: The Helical Orbitals of Allene and Other ...
    Apr 25, 2018 · We trace the origins and detailed composition of the helical orbitals of cumulenes, which emerge in the simplest Hückel model and are not much modified in ...
  33. [33]
    Barrier to internal rotation in ethane from the microwave spectrum of ...
    Aug 1, 1979 · Rotational spectra of CH3CHD2 were observed by using a source‐modulation microwave spectrometer. Of 26 observed transitions, 16 showed ...
  34. [34]
    Coupled cluster calculations of equilibrium geometries, harmonic ...
    At the QCISD(T)/cc-pVQZ//MP2/cc-pVQZ level, the rotational barrier is 2.789 kcal/mol, while the experimental value is 2.882±0.010 kcal/mol [11]. In his study, ...