Fact-checked by Grok 2 weeks ago

Pyramidal inversion

Pyramidal inversion is a stereochemical rearrangement in which a three-coordinate central atom exhibiting a pyramidal geometry interconverts between two equivalent enantiomeric configurations by traversing a planar transition state, effectively resembling the flipping of an umbrella. This process, also termed umbrella inversion, is a polytopal fluxional motion commonly observed in molecules with a lone pair on the central atom, such as nitrogen in ammonia (NH₃) or amines. The phenomenon was first elucidated through spectroscopic studies of in the early , revealing rapid inversion dynamics driven by quantum tunneling and low barriers. In , the inversion barrier is approximately 5.8 kcal/mol, enabling the molecule to invert billions of times per second at , which precludes the isolation of stable enantiomers. Factors influencing the barrier height include the of substituents, , and the size of the central atom; for instance, in phosphines like PH₃, the barrier increases to around 30–35 kcal/mol due to poorer overlap of 3p orbitals with orbitals, allowing for slower inversion and potential stereochemical stability at ambient conditions. This difference is pivotal in , as it explains why chiral amines typically racemize rapidly, whereas P-stereogenic phosphines can be configurationally stable and useful as s in asymmetric catalysis. Pyramidal inversion extends beyond nitrogen and phosphorus compounds to include arsines, stibines, sulfoxides, and even certain carbanions or under specific conditions, such as high pressure in solid-state iodates. Computational methods, including and calculations, have been instrumental in quantifying these barriers and elucidating transition states, aiding in the design of molecules with controlled stereodynamics. The process underscores fundamental principles of molecular and dynamics, with ongoing research exploring its role in biological systems and .

Fundamentals

Definition

Pyramidal inversion is a stereochemical process observed in trigonal pyramidal molecules, where the central atom—typically from group 15 of the periodic table (pnictogens)—is bonded to three substituents and possesses a of electrons, enabling rapid interconversion between two enantiomeric configurations. This dynamic fluxional motion occurs via passage through a planar , effectively exchanging the positions of the lone pair relative to the plane defined by the three substituents. The phenomenon is most famously exemplified by (NH₃), where the atom inverts its pyramidal geometry, resulting in the two mirror-image forms being indistinguishable at due to the low energy barrier. The structural prerequisite for pyramidal inversion is the inherent of the non-planar in such systems when the three substituents differ, creating two enantiomers that interconvert during inversion; however, if the substituents are identical, as in NH₃, the configurations are equivalent, and no net optical activity persists. This process is analogous to an flipping inside out, a that highlights the smooth transition from one pyramidal form to its without bond breaking. While primarily associated with pnictogen-centered molecules like (PH₃) and (AsH₃), similar inversion can occur in select group 16 () compounds, such as certain sulfoxides or selenoxides under specific conditions, though these are less common due to higher barriers. The pyramidal geometry of , foundational to understanding inversion, was first conceptualized by in his 1916 valence theory, which posited the tetrahedral arrangement of four electron pairs around nitrogen (three bonding and one lone pair), leading to a trigonal pyramidal molecular shape. The dynamic inversion itself was experimentally confirmed for in 1934 through by Cleeton and Williams, who observed spectral splitting attributable to the two inversion states. Further validation came in the via (NMR) spectroscopy, where studies on amines by Gutowsky and coworkers revealed line-broadening effects due to inversion rates, enabling quantitative assessment of the process in organic systems.

Molecular Geometry

Pyramidal inversion occurs in molecules exhibiting trigonal pyramidal geometry, a shape arising from the arrangement of three bonding pairs and one lone pair around a central atom with five valence electrons, such as nitrogen in ammonia (NH3). According to Valence Shell Electron Pair Repulsion (VSEPR) theory, these molecules are denoted as AX3E, where the central atom (A) is bonded to three substituents (X) and possesses one lone pair (E). The electron pairs adopt a tetrahedral arrangement to minimize repulsion, but the lone pair's greater spatial demand—due to its position in a hybrid orbital—displaces the bonding pairs, resulting in a trigonal pyramidal molecular geometry with compressed bond angles around 107° rather than the ideal tetrahedral 109.5°./03:_Simple_Bonding_Theory/3.02:_Valence_Shell_Electron-Pair_Repulsion/3.2.01:_Lone_Pair_Repulsion) The central atom's sp3 hybridization underpins this geometry. In NH3, the nitrogen atom hybridizes its 2s and three 2p orbitals to form four equivalent sp3 orbitals, three of which overlap with hydrogen 1s orbitals to create N-H sigma bonds, while the fourth holds the lone pair. This hybridization positions the lone pair in a manner that exerts stronger repulsion on the adjacent bonding pairs, causing the pyramidal distortion and reducing the H-N-H bond angle to 106.7° as determined by microwave spectroscopy. The lone pair's effective larger size stems from its higher electron density concentration near the central atom compared to bonding pairs, amplifying interpair repulsions in the tetrahedral framework./Advanced_Inorganic_Chemistry_(Wikibook)/01:_Chapters/1.08:_NH3_Molecular_Orbitals) Compared to tetrahedral geometry (Td point group symmetry, as in CH4), the lone pair in AX3E molecules lowers the symmetry to C3v, introducing a C3 rotation axis along the lone pair direction and three vertical mirror planes bisecting the bonding angles. This C3v symmetry reflects the molecule's threefold rotational equivalence and the absence of higher symmetry elements due to the apical lone pair. Key structural parameters for NH3 include an N-H bond length of approximately 1.012 Å and the aforementioned H-N-H angle of 106.7°, both experimentally verified. Substituents replacing hydrogen atoms can modulate planarity; bulky groups increase steric crowding, enhancing pyramidal height, while conjugating or electron-withdrawing substituents partially flatten the pyramid by stabilizing configurations closer to trigonal planar through orbital interactions./Advanced_Inorganic_Chemistry_(Wikibook)/01:_Chapters/1.08:_NH3_Molecular_Orbitals)

Mechanism of Inversion

Classical Description

Pyramidal inversion in the classical description refers to the stereochemical process by which a trigonal pyramidal molecule, such as (NH₃), interconverts between two enantiomeric configurations without breaking any bonds. This conformational change involves the central atom and its three substituents transitioning through a planar , where the on the central atom effectively migrates to the opposite side of the plane formed by the substituents. The process begins with the molecule in its ground-state pyramidal geometry, proceeds to a coplanar , and concludes with the reformed pyramid oriented as the of the initial structure. This geometric rearrangement is vividly analogous to the flipping of an inside out, where the "handle" (the central atom with its ) inverts relative to the "canopy" (the three substituents), allowing the to pass seamlessly from one pyramidal orientation to the other. In this classical view, the inversion is driven by overcoming the barrier, resulting in a continuous between the two equivalent pyramidal forms. Importantly, no covalent bonds are formed or broken during this motion; it is purely a reorganization of pairs and positions within the intact molecular framework. For specifically, the classical activation energy leads to an inversion rate on the order of 4 × 10¹⁰ s⁻¹ at , which is sufficiently rapid to render the enantiomeric pyramidal forms indistinguishable on typical experimental timescales. This high rate necessitates cryogenic conditions to observe or isolate the individual enantiomers, as the thermal motion averages the configurations at ambient temperatures.

Transition State

The transition state of pyramidal inversion features a planar with D_{3h} symmetry, where the central atom achieves sp² hybridization and the occupies a pure p-orbital oriented to the of the substituents. This configuration marks the midpoint of the inversion process, transforming the pyramidal into its through an umbrella-like flip. Accompanying this rehybridization, the bonds between the central atom and substituents undergo contraction due to increased s-character in the hybrid orbitals. In (NH₃), for instance, the N-H shortens from an value of approximately 1.012 Å to 0.997 Å at the . On the (PES), the corresponds to a , characterized by a single imaginary vibrational frequency that corresponds to the symmetric inversion mode along the . This connects the two equivalent pyramidal minima, facilitating the fluxional behavior observed in such molecules. Spectroscopic evidence for the planar transition state has been derived from matrix isolation studies, where IR and Raman spectra reveal vibrational signatures attributable to the predicted D_{3h} configuration, particularly through analysis of the umbrella bending mode and its overtone progressions in trapped species.

Energy Barriers

Calculation and Measurement

The energy barriers for pyramidal inversion are determined through a combination of experimental and computational techniques, focusing on measuring inversion rates or directly modeling the (PES) along the inversion coordinate. Experimental methods primarily include and (NMR) spectroscopy. has been used to probe the inversion in small molecules like (NH₃), where the barrier height is derived from the splitting of rotational levels due to the inversion motion. The first such measurement for NH₃ yielded a barrier of 2020 ± 12 cm⁻¹ (approximately 5.8 kcal/mol), obtained by fitting spectroscopic data to a potential function. This approach, pioneered by Swalen and Ibers in 1962, provided the initial quantitative insight into the inversion barrier for NH₃. For larger molecules where inversion rates are slower, temperature-dependent NMR , particularly dynamic NMR (DNMR), measures the barrier by observing signal coalescence or line-shape changes as the inversion matches the NMR timescale. The inversion k at the coalescence temperature T_c is approximated using the relation k_c \approx 2.22 \Delta \nu, where \Delta \nu is the ; the E_a is then from the temperature dependence via the : k = A \exp\left(-\frac{E_a}{RT}\right) where A is the pre-exponential factor, R is the gas constant, and T is the temperature. This method has been applied extensively to amines and related compounds to derive free energy barriers \Delta G^\ddagger typically in the range of 40–80 kJ/mol. Computational approaches model the PES to directly compute barrier heights, employing ab initio methods such as Hartree-Fock (HF), second-order Møller-Plesset perturbation theory (MP2), and coupled-cluster singles, doubles, and perturbative triples [CCSD(T)], as well as density functional theory (DFT). These techniques optimize geometries of pyramidal minima and planar transition states, yielding barrier heights from differences in electronic energies. For NH₃, early ab initio calculations at the HF level underestimated the barrier, while higher-level CCSD(T) methods with large basis sets reproduce experimental values closely, often within 1–2 kJ/mol. Modern DFT functionals, such as B3LYP, provide efficient PES modeling and barrier estimates for NH₃ close to experimental values.

Influencing Factors

Several factors influence the energy barrier to pyramidal inversion, modulating the stability of the pyramidal ground state relative to the planar transition state. Substituent effects play a significant role, with the direction depending on the nature of the substituents. For highly electronegative groups like fluorine, the barrier increases due to poorer p-orbital overlap in the planar transition state and stabilization of the pyramidal form. For example, in nitrogen compounds, the inversion barrier for NF₃ is approximately 78 kcal/mol, substantially higher than the 5.8 kcal/mol for NH₃. In contrast, conjugative electron-withdrawing groups can lower the barrier by stabilizing the planar transition state through resonance; for instance, trifluoromethyl substitution on nitrogen, as in CF₃NH₂, reduces the barrier compared to CH₃NH₂ by enhancing planarity in the transition state. Steric hindrance from bulky substituents can increase the inversion barrier by disfavoring the planar , where closer approach of groups heightens non-bonded repulsions. In sterically crowded systems like 1,3,4-oxadiazolidines with methyl groups at positions 3 and 5, inversion barriers exceed 30 kcal/, far higher than in unhindered analogs, owing to severe crowding in the flattened . The size of the central atom also affects the barrier, with heavier elements exhibiting higher barriers than lighter ones like . For instance, phosphines display inversion barriers around 31 kcal/ for (CH₃)₃P, compared to ~7 kcal/ for (CH₃)₃N, due to the larger s-p orbital separation in , which stabilizes the pyramidal with an s-rich orbital more than the planar p-lone pair configuration. Solvent effects arise from dielectric interactions that differentially stabilize the polar pyramidal state versus the less polar planar ; polar solvents often lower the barrier by better solvating the . In N-phenyloxaziridines, increasing solvent reduces the barrier by up to 2 kcal/mol through enhanced stabilization of the charge-separated . Temperature influences the observed inversion rate via thermal population of the , as described by the for the activation free energy: \Delta G^\ddagger = -RT \ln\left(\frac{k h}{k_B T}\right), where higher temperatures increase the rate constant k exponentially, effectively allowing faster inversion without altering the intrinsic barrier. For amines, rates double roughly every 10–15 K rise near . Isotopic substitution, such as deuterium for hydrogen, slightly increases the barrier due to differences in zero-point energies along the inversion coordinate. In ammonia, the effective barrier for ND₃ is ~0.2 kcal/mol higher than for NH₃, as the lower vibrational frequencies reduce the zero-point energy stabilization in the pyramidal ground state relative to the transition state.

Nitrogen Inversion

Characteristics

Pyramidal inversion in nitrogen-containing pyramidal molecules, such as amines, is distinguished by relatively low energy barriers, typically ranging from 23 to 35 kJ/mol for unstrained examples. These barriers permit exceptionally rapid inversion rates at room temperature, often exceeding 10^6 Hz, facilitating the interconversion of the two enantiomeric pyramidal forms on a timescale much faster than typical spectroscopic or chemical resolution methods. In (NMR) spectroscopy, this rapid dynamics results in signal averaging, where enantiotopic protons or groups appear as single, sharp peaks due to the time-averaged symmetric structure. At sufficiently low temperatures, for systems with higher barriers, the inversion rate can be slowed enough to enable dynamic NMR studies through techniques like line-shape analysis or coalescence, particularly when the invertomers are diastereomeric due to molecular asymmetry, allowing measurement of barriers. The stereochemical ramifications of such swift inversion preclude the isolation and stable existence of pyramidal enantiomers under ordinary conditions, in stark contrast to the persistent observed at tetrahedral carbon centers. This inherent renders most trivalent nitrogen compounds achiral despite their momentary pyramidal geometry. Relative to pyramidal phosphorus analogs, nitrogen inversion barriers are substantially lower, attributable in part to nitrogen's greater , which influences the energetic preference for the pyramidal over the planar in these lighter, more compact systems.

Quantum Effects

In pyramidal inversion of nitrogen compounds like , quantum tunneling plays a crucial role due to the relatively low energy barrier, enabling the molecule to transition between pyramidal configurations without classical surmounting of the barrier. This results in a double-minima , where the two equivalent minima correspond to the inverted geometries, leading to splitting of the vibrational energy levels into symmetric and antisymmetric states. The inversion in was first observed through in the , revealing the characteristic splitting due to tunneling. For (NH₃), the barrier height is approximately 2023 cm⁻¹, sufficiently low to make tunneling the dominant mechanism at low temperatures. The effects of tunneling are evident in the infrared spectrum through inversion doubling of the umbrella mode (ν₂), which corresponds to the out-of-plane bending vibration involved in inversion. In , this manifests as two closely spaced bands at 930 cm⁻¹ and 965 cm⁻¹, reflecting the energy splitting between the vibrational sublevels. Similar splittings are observed in deuterated (ND₃), where isotopic substitution reduces the frequencies to approximately 748 cm⁻¹ and 784 cm⁻¹, with the smaller enhancing the relative tunneling contribution. These features are captured via far-infrared , which probes the low-energy vibrational and rotational transitions sensitive to the quantum level structure. To quantify tunneling rates beyond simple , variational methods such as theory are employed, which model the imaginary-time trajectory across the barrier to compute splitting energies accurately. An approximate for the tunneling probability is given by \kappa \approx \exp\left(-\frac{2\pi E_a}{h \nu}\right), where E_a is the , h is Planck's constant, and \nu is the of the imaginary mode at the ; for NH₃, this yields tunneling frequencies on the order of 10¹⁰ s⁻¹. The (ZPE) associated with the vibrations further modulates the effective inversion barrier by lowering it relative to the classical value, as the ZPE occupies part of the barrier region in the . Computational vibrational analyses, including anharmonic corrections, incorporate this ZPE to refine barrier estimates; for NH₃, the ground-state barrier is about 2020 cm⁻¹, requiring ZPE back-correction of roughly 500–600 cm⁻¹ to obtain the classical height. This quantum correction is essential for matching experimental splittings and highlights the deviation from classical predictions in light-element systems.

Examples in Amines

One of the most well-studied examples of pyramidal inversion occurs in (NH₃), where the atom rapidly interconverts between two enantiomeric pyramidal configurations through a planar . The energy barrier for this process is 24.2 kJ/mol, enabling an inversion frequency on the order of 10¹⁰ s⁻¹, as evidenced by the spectral splitting at 24.87 GHz. This high rate renders the hydrogens magnetically equivalent on the NMR timescale even at very low temperatures, precluding observation of coalescence. In primary amines like (CH₃NH₂), pyramidal inversion proceeds with a comparable energy barrier of 23.2 kJ/mol, slightly lower than in due to the electron-donating effect of the methyl that stabilizes the pyramidal relative to the . The process remains rapid at ambient conditions, with influences such as steric hindrance playing a minor role in this case, maintaining fast exchange on typical spectroscopic timescales. Cyclic amines illustrate how structural constraints can elevate inversion barriers. In , the three-membered ring imposes significant angle strain on the planar , raising the barrier to approximately 81.7 kJ/mol (19.5 kcal/mol). This increased barrier permits partial resolution of enantiomers in chiral substituted and facilitates dynamic NMR studies, where coalescence of signals for invertomers can be observed at elevated temperatures around 300–350 K. Bridgehead nitrogen compounds, such as those in the 7-azabicyclo[2.2.1]heptane framework, represent cases of geometrically restricted inversion. For 7-methyl-7-azabicyclo[2.2.1]heptane, the bicyclic structure enforces a higher barrier of about 59 kJ/mol (14.1 kcal/mol), slowing the inversion sufficiently for measurement of the barrier, though the process is not completely suppressed as once hypothesized for such rigid systems. This allows assessment of invertomer populations and exchange rates, highlighting the role of skeletal strain in modulating inversion dynamics.

Inversion in Heavier Elements

Phosphorus and Arsenic Compounds

Phosphines of the general formula PR₃ exhibit pyramidal inversion through a planar , with energy barriers typically in the range of 120–140 kJ/mol, significantly higher than those observed in amines and enabling the existence of stable chiral configurations at . For (PH₃), the experimental inversion barrier is estimated at 132 kJ/mol (31.5 kcal/mol), determined from and theoretical corrections for tunneling effects. This results in an inversion rate of approximately 10^{-10} s^{-1} at 298 K, far slower than the rapid interconversion in (NH₃). The elevated barriers in phosphines arise primarily from the longer P–C or P–H bond lengths (compared to N analogs), which increase the angular strain in the pyramidal ground state while providing less stabilization in the planar form due to poorer orbital overlap; involvement of d-orbitals in has been suggested to contribute to planarity in the but remains a subject of debate in computational analyses. Arsines (AsR₃) display similar pyramidal geometries but with inversion barriers that are generally comparable to those of phosphines, facilitating access to planar configurations yet still allowing for chiral stability under appropriate conditions. For (AsH₃), spectroscopic estimates place the barrier at 134 kJ/mol (32.1 kcal/mol), reflecting the even larger of and reduced lone-pair s-character, which diminish the energy difference between pyramidal and planar states. In substituted arsines, such as ethylmethylphenylarsine, the barrier is measured at 125 kJ/mol (29.8 kcal/mol) via dynamic NMR studies of kinetics. These values indicate that arsenic compounds invert at rates similar to their phosphorus counterparts (~10^{-10} s^{-1}), permitting enantiomeric resolution in some cases. Stibines (SbR₃) follow the trend of increasing barriers down Group 15, with even higher energy requirements for inversion due to the larger atomic size and greater s-p separation. For (SbH₃), calculations estimate the barrier at approximately 193 kJ/mol (46.2 kcal/mol), resulting in extremely slow inversion rates and enhanced configurational stability compared to lighter analogs. This allows chiral stibines to maintain under ambient conditions, though practical applications are limited by the toxicity and reactivity of compounds. Synthetic chiral phosphines, exemplified by ethylmethylphenylphosphine (PPhMeEt), have inversion barriers around 130–145 kJ/mol (31–35 kcal/mol), as determined from rates in polar solvents, allowing their into enantiomers through diastereomeric formation with chiral resolving agents like . Bulky substituents, such as in menthyl-based phosphines, can elevate these barriers to 140–160 kJ/mol, enhancing configurational stability for prolonged storage and use. Optically active P-chiral phosphines serve as ligands in transition-metal , where the stereogenic center at phosphorus imparts high enantioselectivity; notable examples include TangPhos and DuanPhos in Pd- and Rh-catalyzed asymmetric allylic substitutions and hydrogenations, achieving ee values up to 99% in the synthesis of pharmaceuticals and fine chemicals.

Differences from Nitrogen

Pyramidal inversion barriers in group 15 compounds increase down the group from to and , reflecting a trend where the (Ea) rises due to greater stabilization of the pyramidal relative to the planar . Representative values include approximately 5.8 kcal/mol (24 kJ/mol) for (NH3), 32 kcal/mol (134 kJ/mol) for (PH3), and 32 kcal/mol (134 kJ/mol) for (AsH3). This progression arises primarily from the larger s-p orbital energy separation in heavier pnictogens, which favors a with high s-character in the pyramidal configuration, making the transition to a pure p-character in the planar state more energetically costly. In , the pyramidal aligns closely with sp³ hybridization, featuring bond angles near 107° and a orbital with roughly 25% s-character. In contrast, and compounds exhibit bond angles closer to 93° and 91.8°, respectively, indicative of bonds with predominantly p-character and a orbital possessing significantly higher s-character (up to ~84% in PH3). This hybridization shift enhances the of the pyramidal form for heavier elements, as the contracted s-orbitals lower the energy of the , thereby elevating the inversion barrier compared to analogs. Quantum tunneling plays a prominent role in nitrogen inversion, particularly in , where the low barrier and light result in observable splitting of vibrational levels (e.g., 23.8 cm⁻¹ inversion doubling), facilitating rapid effective inversion even below classical rates. For and compounds, however, the substantially higher barriers and increased atomic masses suppress tunneling effects, rendering the process classically activated with negligible quantum contributions at typical temperatures. The elevated inversion barriers in and compounds enable greater stereochemical stability, allowing enantiomers of trivalent derivatives to be resolved and maintained at without rapid , in stark contrast to the fleeting of analogous species due to their fast inversion. This difference underpins the utility of chiral phosphorus compounds in stereoselective applications, where configurational integrity persists under ambient conditions.

Exceptions

In small bicyclic heterocycles with bridgehead phosphorus atoms, pyramidal inversion is significantly hindered by geometric strain, analogous to for bridgehead double bonds in carbon systems. The planar required for inversion imposes trans-like geometry in the rings, which is energetically unfavorable in strained structures like bicyclo[2.2.1] systems. For instance, in 1-phosphabicyclo[2.2.1]heptane, the inversion barrier exceeds 100 kJ/mol, preventing facile interconversion at and allowing isolation of configurationally stable isomers. In metal-coordinated phosphines, ligation to the fixes the pyramidal geometry, effectively blocking inversion without dissociation from the metal center. This is particularly evident in complexes like (RhCl(PPh3)3), where the ligands maintain their throughout the , as the coordinated phosphorus cannot achieve the planar without weakening the metal-phosphorus bond. Such coordination-induced stability is crucial for using chiral phosphines, where is suppressed even at elevated temperatures. Bulky substituents, such as in pentaarylphosphorus compounds (e.g., those with multiple ortho-substituted aryl groups on a trivalent ), raise the inversion barrier above 150 kJ/mol through steric crowding that disfavors the planar . This allows stable at , enabling the isolation and use of enantiopure forms in stereoselective reactions. For example, triarylphosphanes with highly congested positions exhibit barriers that ensure configurational integrity over extended periods, contrasting with less substituted analogs.

References

  1. [1]
    pyramidal inversion (P04956) - IUPAC
    A polytopal rearrangement in which the change in bond directions to a three-coordinate central atom having a pyramidal arrangement of bonds.
  2. [2]
    Pyramidal Inversion - Chemistry LibreTexts
    Feb 28, 2022 · 2a and 2b, however, have no independent existence because they rapidly interconvert. The process by which 2a and 2b interconvert is called pyramidal inversion.
  3. [3]
    a fresh perspective on the ammonia pyramidal inversion and bond ...
    Oct 4, 2023 · This work offers a comprehensive and fresh perspective on the bonding evolution theory (BET) framework, originally proposed by Silvi and collaborators.
  4. [4]
    Factors Affecting Energy Barriers for Pyramidal Inversion in Amines ...
    Apr 26, 2013 · An undergraduate exercise in computational chemistry that investigates the energy barrier for pyramidal inversion of amines and phosphines is presented.
  5. [5]
    Pyramidal inversion in the solid state - RSC Publishing
    Pyramidal inversion is a stereochemical phenomenon that describes the interconversion between two equivalent pyramidal configurations of the same chemical ...
  6. [6]
    Nitrogen Inversion - an overview | ScienceDirect Topics
    Nitrogen inversion is a rapid interconversion of enantiomers in tertiary amines, where the stereogenic nitrogen atom's configuration changes.
  7. [7]
    Gilbert Newton Lewis | Science History Institute
    In 1916 Gilbert Newton Lewis (1875–1946) published his seminal paper suggesting that a chemical bond is a pair of electrons shared by two atoms. Once ...Missing: pyramidal inversion ammonia
  8. [8]
    Electromagnetic Waves of 1.1 cm Wave-Length and the Absorption ...
    Electromagnetic Waves of 1.1 cm Wave-Length and the Absorption Spectrum of Ammonia. C. E. Cleeton and N. H. Williams. University of Michigan. PDF Share. X ...Missing: inversion | Show results with:inversion
  9. [9]
    VSEPR NH3 Ammonia - ChemTube3D
    These are arranged in a tetrahedral shape. The resulting molecular shape is trigonal pyramidal with H-N-H angles of 106.7°. Ammonia adopts a trigonal pyramidal ...
  10. [10]
    VSEPR calculation for ammonia, NH 3 - University of Sheffield
    The consequence of this for ammonia is that the lone pair makes room for itself by pushing the three hydrogen atoms together a little and the H-N-H bond angles ...
  11. [11]
  12. [12]
    [PDF] Resource-Efficient Quantum Circuits for Molecular Simulations - arXiv
    Dec 7, 2023 · The transition state, with D3h symmetry, occurs when the N atom becomes coplanar to the H atoms. (δ = 0). The geometrical parameters and ...
  13. [13]
    Potential Function for the Inversion of Ammonia - AIP Publishing
    The data on the inversion of NH 3 and ND 3 , summarized by Benedict and Plyler, have been fit in terms of the angular coordinate with a potential function.
  14. [14]
    A dynamic proton NMR and ab initio MO investigation of the barrier ...
    A dynamic proton NMR and ab initio MO investigation of the barrier to pyramidal inversion in azetidine ... methods describe inversion and racemization barriers?.
  15. [15]
    Electronic Structure and Inversion Barrier of Ammonia - AIP Publishing
    Apr 15, 1970 · Ab initio molecular orbital wavefunctions have been constructed for the planar and pyramidal conformations of ammonia in its ground electronic state.
  16. [16]
    The inversion potential for NH3 using a DFT approach - ScienceDirect
    Our results indicate that the approach with DFT and the Becke3LYP method give us the most realistic description of the potential function for the vibrational ...
  17. [17]
    inversion of NH3, NH2F, NHF 2' NF 3 and PH3, PH2F, PHF2, PF3
    A recent study1 of the inversion of NHF 2 has shown marked changes in bond lengths in the transition state compared with the equilibrium structure, and ...
  18. [18]
    Pronounced Steric Hindrance for Nitrogen Inversion in 1,3,4 ...
    Aug 18, 2000 · Very high inversion barriers for a pyramidal nitrogen atom have been found for the sterically hindered 1,2,4‐oxadiazolidines 1 (R1, R2=CH3) ...
  19. [19]
    Pyramidal inversion - Wikipedia
    The general phenomenon of pyramidal inversion applies to many types of molecules, including carbanions, amines, phosphines, arsines, stibines, and sulfoxides.
  20. [20]
    N‐Inversion in N‐phenyloxaziridine: substituent and solvent effects ...
    Mar 12, 2019 · The incorporation of an oxygen atom in the aziridine ring strongly weakens the N-inversion process. In addition, while both t-butyl substituent ...
  21. [21]
    Strong inverse kinetic isotope effect observed in ammonia charge ...
    Jan 10, 2020 · Isotopic substitution has long been used to understand the detailed mechanisms of chemical reactions; normally the substitution of hydrogen ...
  22. [22]
    3.3.2: Structure and Properties of Amines
    ### Summary of Pyramidal Inversion in Amines
  23. [23]
    Nuclear Magnetic Resonance Spectra and Nitrogen Inversion in 1 ...
    N.m.r. spectra of cyclic amines. II—Factors influencing the chemical shifts of α‐protons in aziridines. Organic Magnetic Resonance 1977, 9 (6) , 328-332 ...
  24. [24]
  25. [25]
    Review Large amplitude inversion tunneling motion in ammonia ...
    Jun 1, 2021 · This review summarizes four most classic examples of molecules featuring inversion tunneling motion and their coordination complexes.
  26. [26]
    Fourier transform infrared absorption spectroscopy characterization ...
    Oct 29, 2012 · The ammonia gave rise to components at 930 and 965 cm−1, assigned ... (inversion doubling).33 The feature observed at 1621 cm−1 was ...
  27. [27]
    Vibrational relaxation of ND3 trapped in a rare gas matrix
    Aug 6, 2025 · The ν2 umbrella mode of vibration–inversion of ammonia trapped in condensed media exhibits typical tunneling slow-down due to the statics, the ...
  28. [28]
    Analytical approach for the tunneling process in double well ...
    Oct 1, 2020 · For NH3, the tunneling frequencies were 2.56 , 3.67 and 1.12 × 10 10 s−1, using the APS function, WKB and instanton approaches respectively, ...
  29. [29]
    [PDF] the inversion barrier in ammonia - University of Regina
    Hence, in the present work, barriers were computed generally at three levels of theory: Hartree-Fock (HF), MP2 with nitrogen ls core electrons frozen (MP2(fc)).
  30. [30]
    Ammonia Inversion: Classical and Quantum Models
    The ammonia molecule. NH. 3. has a trigonal pyramidal configuration, with the nitrogen atom connected to three hydrogen atoms. The molecule readily undergoes ...
  31. [31]
    [PDF] 1.01 Aziridines and Azirines: Monocyclic - Elsevier
    The barrier for pyramidal inversion in aziridine itself is 19.5 kcal mol 1. Aziridines are less basic than acyclic amines due to the increased s character ...
  32. [32]
    Control of Pyramidal Inversion Rates by Redox Switching
    The dynamics of pyramidal nitrogen inversion can be controlled by reversible redox switching in trans-2,3-diphenylaziridines bearing a suitable 1,4- ...<|separator|>
  33. [33]
  34. [34]
    Why is the inversion barrier larger in PH3 than it is in NH3?
    Oct 7, 2015 · The gist of it is that because nitrogen is smaller and lighter than phosphorus, the rate of tunnelling and hence inversion of chirality is much faster.
  35. [35]
    Barrier to pyramidal inversion in ethylmethylphenylarsine
    PYRAMIDAL INVERSION BARRIERS OF PHOSPHINES AND ARSINES*,†. Transactions of the New York Academy of Sciences 1973, 35 (3 Series II) , 227-242. https://doi ...
  36. [36]
    Synthesis and absolute configuration of optically active phosphine ...
    The stereoselective conversion of epimerized alkoxyl phosphine–borane ... Organophosphorus Catalysis to Bypass Phosphine Oxide Waste. ChemSusChem 2013 ...
  37. [37]
    Synthesis and applications of high-performance P-chiral phosphine ...
    This review article describes the synthesis and applications of P-chiral phosphine ligands possessing chiral centers at the phosphorus atoms.
  38. [38]
    Effect of ligand electronegativity on the inversion barrier of phosphines
    Factors Affecting Energy Barriers for Pyramidal Inversion in Amines and ... PYRAMIDAL INVERSION BARRIERS OF PHOSPHINES AND ARSINES*,†. Transactions of ...
  39. [39]
    Low barrier to pyramidal inversion in phospholes. Measure of ...
    Kurt Mislow. PYRAMIDAL INVERSION BARRIERS OF PHOSPHINES AND ARSINES*,†. Transactions of the New York Academy of Sciences 1973, 35 (3 Series II) , 227-242 ...
  40. [40]
    [PDF] Product Class 6: Cyclic Phosphines - Thieme Connect
    This section concerns secondary and tertiary phosphines with a phosphorus atom that is part of a saturated or at most partially unsaturated carbocyclic ring ...
  41. [41]
    Characterization and interconversion of metal-phosphorus single ...
    Glueck. Copper-Catalyzed Asymmetric Alkylation of Secondary Phosphines via Rapid Pyramidal Inversion in P-Stereogenic Cu–Phosphido Intermediates.
  42. [42]
    The chemistry of phosphines in constrained, well-defined ...
    Feb 19, 2021 · The binding of these phosphines to metal centres in the presence of these hosts alters the equilibria between species with differing numbers ...
  43. [43]
    About the Inversion Barriers of P‐Chirogenic Triaryl‐Substituted ...
    Jun 14, 2018 · About the Inversion Barriers of P-Chirogenic Triaryl-Substituted Phosphanes. Jens Holz,. Corresponding Author. Jens Holz. jens.holz@catalysis.de ...
  44. [44]
    Taming PH3: State of the Art and Future Directions in Synthesis
    Sep 7, 2022 · In this perspective we highlight current trends in forming new P–C/P–OC bonds with PH 3 and discuss the challenges involved with selectivity and product ...
  45. [45]
    Nitrogen Inversion Barrier of 2-Methyl-2-azabicyclo[2.2.1]heptane ...
    In this paper, we report the dynamic NMR (DNMR) analysis of 13C spectra of an isomeric bicyclic amine, 2-methyl-2-azabicyclo[2.2.1]heptane (2). A relatively low ...
  46. [46]
    Nitrogen Inversion in Cyclic Amines and the Bicyclic Effect
    Aug 6, 2025 · Chemoenzymatic enantioselective synthesis of 7-azabicyclo[2.2.1]heptane derivatives ... nitrogen inversion barrier according to the ...