Fact-checked by Grok 2 weeks ago

Lone pair

A lone pair is a pair of valence electrons occupying an orbital on an atom that is not involved in covalent bonding and thus remains localized on that single atom, often represented by two dots in Lewis structures. These nonbonding electron pairs play a crucial role in determining molecular geometry through the valence shell electron pair repulsion (VSEPR) theory, where they exert stronger repulsive forces on surrounding bonding pairs due to their higher electron density and occupation of larger spatial regions compared to shared bonding electrons. In acid-base chemistry, lone pairs enable atoms to act as Lewis bases by donating electrons to electron-deficient species, forming coordinate covalent bonds essential for coordination compounds and reactions involving metal ions. This electron donation capacity also contributes to intermolecular forces such as hydrogen bonding, where lone pairs on electronegative atoms like oxygen or interact with hydrogen atoms bound to similar electronegative atoms, influencing properties like boiling points and in molecules such as or . Furthermore, the presence and orientation of lone pairs affect molecular , as they create regions of high that can lead to moments even in otherwise symmetric structures. Beyond basic molecular structure, lone pairs are integral to advanced chemical phenomena, including their stereochemical activity in solid-state materials where they can distort coordination geometries to enhance properties like ion conductivity or in compounds containing elements such as lead or . In and inorganic synthesis, lone pairs facilitate nucleophilic attacks and stabilization, underscoring their fundamental importance across chemical disciplines.

Fundamentals

Definition and Characteristics

A lone pair, also known as an unshared or nonbonding pair, consists of two electrons localized on a single atom that are not involved in covalent bonding with another atom. These electrons occupy a specific orbital in the atom's valence shell, contributing to the fulfillment of the , which posits that atoms tend to achieve a configuration with eight valence electrons. The concept of lone pairs was introduced by in his 1916 seminal paper, where he described the and covalent bonding as involving shared electron pairs, with unshared pairs completing the octet on atoms such as oxygen and . Lone pairs exhibit characteristics similar to bonding pairs in that they occupy or orbitals, but they generate stronger repulsive forces in due to their higher localized compared to the more delocalized density in bonding pairs. This increased repulsion arises because lone pair electrons are held closer to the of the central atom, without being shared across an interatomic bond. Lone pairs play a key role in calculating an atom's , a measure used to assess the electron distribution in Lewis structures. The is given by the formula: \text{Formal charge} = (\text{valence electrons}) - (\text{nonbonding electrons}) - \frac{1}{2} (\text{bonding electrons}) where nonbonding electrons include those in lone pairs. This calculation, formalized by , helps identify the most stable structures by minimizing formal charges. Representative examples illustrate these properties: in the water molecule (H₂O), the oxygen atom possesses two lone pairs, completing its octet alongside two bonding pairs to the atoms; similarly, in (NH₃), the atom has one lone pair, achieving its octet with three bonding pairs to atoms.

Representation in Lewis Structures

In Lewis dot structures, lone pairs are depicted as pairs of dots positioned adjacent to the symbol of the atom possessing them, distinguishing these non-bonding electrons from bonding pairs, which are represented by lines connecting atoms. This convention, introduced by in 1916, facilitates the visualization of distribution in molecules and ions. The construction of Lewis structures follows the , particularly for elements in the second period (such as carbon, , oxygen, and ), where atoms seek to surround themselves with eight electrons to achieve stability akin to . After placing atoms and drawing bonds to connect them, remaining electrons are distributed as lone pairs to fulfill this octet for each atom, starting with terminal atoms. To calculate the number of lone pairs on a central , subtract the number of electrons committed to bonding from the 's electrons and divide by 2, as each lone pair consists of two electrons. For a central with V electrons forming B single bonds, the formula is \frac{V - B}{2}. This method assumes standard single bonds; adjustments apply for multiple bonds or charged species. Notation for lone pairs can vary slightly; while dots are standard, some representations use short dashes or lines for lone pairs to streamline sketches, especially in skeletal formulas where hydrogens and some lone pairs are implied. For elements beyond the second period (like or ), the can be exceeded, permitting expanded shells with 10, 12, or more electrons, which may accommodate additional bonding electrons and fewer or no lone pairs on the central . Consider (NH₃) as an example: , with 5 valence electrons, forms three single bonds to atoms (using 3 electrons from nitrogen), leaving 2 electrons as one lone pair, depicted as two dots above the nitrogen symbol in the structure H–N–H with the third H below and :: on N. (H₂O) follows similarly: Oxygen, with 6 valence electrons, forms two single bonds (using 2 electrons), resulting in 4 electrons forming two lone pairs, shown as :: on top and bottom of the O in H–O–H. In contrast, (SF₆) exhibits an expanded octet: , with 6 valence electrons, forms six single bonds to atoms (using all 6 plus additional from fluorines, totaling 12 electrons around sulfur), yielding no lone pairs on sulfur, while each fluorine has three lone pairs to complete its octet. Formal charge, computed as valence electrons minus lone pair electrons minus half the bonding electrons, serves as a tool to verify structure accuracy by minimizing charges on atoms.

Influence on Molecular Geometry

Valence Shell Electron Pair Repulsion Theory

The Valence Shell Electron Pair Repulsion (, introduced by Ronald J. Gillespie and Ronald S. Nyholm in , serves as a foundational model for predicting molecular geometries by considering the spatial arrangement of pairs around a central . The core principle is that pairs in the shell—encompassing both bonding pairs (shared between atoms) and lone pairs (unshared on the central )—exert repulsive forces on one another, positioning themselves to achieve the minimum overall repulsion and thus the lowest energy configuration. This repulsion arises from the and electrostatic interactions among the negatively charged electron domains. In VSEPR, an electron domain is defined as either a lone pair or a bonding pair (including single, double, or triple bonds, each counted as one domain). The total number of domains around the central atom, termed the steric number, dictates the basic geometry, independent of whether the domains are bonding or lone pairs. Lewis structures provide the starting point for identifying these domains. The standard geometries for different steric numbers are outlined below:
Steric Number GeometryExample Central Atom Configuration
2LinearBe in BeCl₂ (AX₂)
3Trigonal planarB in BF₃ (AX₃)
4TetrahedralC in CH₄ (AX₄)
5Trigonal bipyramidalP in PCl₅ (AX₅)
6OctahedralS in SF₆ (AX₆)
These geometries represent the ideal arrangements that minimize repulsion among equivalent domains. A key aspect of VSEPR is the differential space occupation and repulsion strengths: lone pairs, being confined to the central atom, repel more strongly and occupy larger regions than bonding pairs, which are delocalized between atoms. The theory establishes a clear of pairwise repulsions—lone pair–lone pair > lone pair–bonding pair > bonding pair–bonding pair—which guides the placement of lone pairs in positions that further minimize energy, often in equatorial or axial sites depending on the . This hierarchy ensures that lone pairs are positioned to avoid maximum repulsion with other lone pairs or bonding pairs. To illustrate, molecules are classified using the AX_mE_n notation, where A is the central atom, X denotes each bonding domain (m total), and E denotes each lone pair (n total), with m + n equal to the steric number. For ammonia (NH₃), nitrogen has a steric number of 4 (three bonding domains to H and one lone pair, AX₃E), adopting a trigonal pyramidal molecular shape as the lone pair occupies one vertex of the tetrahedron. For water (H₂O), oxygen has a steric number of 4 (two bonding domains to H and two lone pairs, AX₂E₂), resulting in a bent molecular shape with the lone pairs positioned to minimize their mutual repulsion. These examples highlight how lone pairs distort the arrangement from the ideal electron pair geometry without altering the basic framework.

Bond Angle Changes

In molecules exhibiting tetrahedral electron-pair geometry, such as those with four electron domains around the central atom, the ideal bond angle between bonding pairs is 109.5°, as observed in (CH₄) where all domains are bonding pairs. However, the presence of one or more lone pairs leads to compression of the bond angles, as seen in ammonia (NH₃), which has AX₃E geometry and an H–N–H angle of 107.3°, and water (H₂O), with AX₂E₂ geometry and an H–O–H angle of 104.5°./09%3A_Molecular_Geometry_and_Covalent_Bonding_Models/9.02%3A_VSEPR_-Molecular_Geometry)/12%3A_Liquids_Solids_and_Intermolecular_Forces/12.08%3A_Water-_An_Extraordinary_Substance) This deviation arises because lone pairs occupy a larger effective volume than bonding pairs, exerting stronger repulsive forces on adjacent bonding pairs and pushing them closer together. The magnitude of angle compression increases with the number of lone pairs, as multiple lone pairs amplify the overall repulsion within the electron domain arrangement. For instance, in tin(II) chloride (SnCl₂), which adopts AX₂E geometry similar to H₂O, the Cl–Sn–Cl bond angle is approximately 95°, reflecting significant compression due to the lone pair on the larger tin atom, where orbital overlap and steric effects further reduce the angle beyond typical second-period expectations./02%3A_Molecules/2.05%3A_Valence_Shell_Electron-Pair_Repulsion/2.5.01%3A_Lone_Pair_Repulsion) In contrast, for trigonal bipyramidal electron geometry in xenon difluoride (XeF₂, AX₂E₃), the three lone pairs preferentially occupy the equatorial positions to minimize repulsion, leaving the two bonding pairs in axial positions and resulting in an unperturbed F–Xe–F bond angle of 180° with no observable compression in the molecular structure./Descriptive_Chemistry/Elements_Organized_by_Block/2_p-Block_Elements/Group_18%3A_The_Noble_Gases/Xenon_Compounds/XeF_2) Additional factors, such as the of the surrounding atoms, can modulate lone pair effects on bond angles. In (OF₂, AX₂E₂), the F–O–F bond angle is 103.1°, slightly smaller than that of H₂O, because the highly electronegative fluorine atoms withdraw from the bonding pairs, reducing bond pair–bond pair repulsions and allowing the lone pairs to compress the angle further./16%3A_The_Group_16_Elements/16.09%3A_Oxoacids_and_their_Salts) This electronegativity influence highlights how properties interact with lone pair repulsions to fine-tune molecular geometries.

Physical and Chemical Effects

Dipole Moments

Lone pairs contribute to molecular by localizing regions of negative charge on the atom bearing them, which enhances the overall in asymmetric molecules through uneven distribution. This effect arises because lone pairs, being non-bonding pairs, exert a partial negative charge that does not participate in bonding but influences the charge separation within the . In symmetric arrangements, such contributions may cancel out, but in asymmetric cases, they add to the vector sum of bond , leading to a net . The magnitude of the dipole moment, denoted as \mu, is fundamentally given by the equation \mu = q \times d, where q is the magnitude of the partial charges and d is the distance between their centers of positive and negative charge. Lone pairs increase the effective q on the central atom by concentrating electron density, thereby amplifying the charge separation and contributing to a larger \mu when the molecular geometry prevents cancellation. This is particularly evident in molecules where the lone pair distorts the structure away from perfect symmetry, aligning the dipole vectors in a non-canceling manner./Physical_Properties_of_Matter/Atomic_and_Molecular_Properties/Dipole_Moments) For instance, in (NH₃), the single lone pair on the atom results in a trigonal pyramidal , producing a net of 1.47 D primarily directed toward the lone pair region. In (H₂O), the two lone pairs on oxygen create a bent structure with a of 1.85 D, where the lone pairs significantly amplify the beyond what the O-H dipoles alone would provide. By comparison, (CO₂) exhibits a of 0 D due to its linear symmetry and absence of lone pairs on the central carbon atom, allowing the C=O dipoles to cancel perfectly. A useful comparison is between (BF₃), which has no lone pairs on and a trigonal planar leading to a of 0 D, and (NF₃), where the lone pair on induces asymmetry in the otherwise similar trigonal pyramidal shape, yielding a of 0.23 D. In NF₃, the lone pair's contribution is partially offset by the strong of atoms, which reverse the bond directions relative to NH₃, but the net effect still results in measurable . These examples illustrate how lone pairs are key to determining whether a molecule's moments reinforce or cancel, establishing its overall electrical character.

Role in Reactivity and Intermolecular Forces

Lone pairs play a crucial role in chemical reactivity by serving as sites for donation, enabling molecules to act as nucleophiles or bases. In reactions, such as SN2 mechanisms, the lone pair on the nitrogen atom of amines attacks the electrophilic carbon center of alkyl halides, displacing the and forming a new C-N . Similarly, the lone pair on allows it to function as a base, accepting a proton from acids to form ions, which underlies its basicity in aqueous solutions. This donation can also lead to the formation of coordinate covalent bonds, where the lone pair coordinates directly with metal cations or other electrophiles, stabilizing complexes in coordination chemistry. In intermolecular forces, lone pairs contribute significantly to non-covalent interactions, particularly as acceptors in bonding. For instance, the oxygen lone pairs in molecules accept bonds from the O-H groups of neighboring molecules, forming an extensive network that elevates 's to 100°C—substantially higher than expected for a of its size compared to (–60°C), which lacks such effective bonding due to sulfur's weaker lone pair basicity. This lone pair involvement enhances the overall polarity and cohesive forces in polar solvents. Additionally, the localized electron density from lone pairs can amplify van der Waals interactions, as seen in lone pair-π attractions where oxygen lone pairs engage with aromatic rings at van der Waals distances, providing extra stabilization in molecular assemblies. Lone pairs also influence solvent effects by facilitating cation solvation through donation to metal ions. In ethers like diethyl ether (R₂O), the oxygen lone pairs act as Lewis bases, coordinating with cations such as Li⁺ or Na⁺ to form solvated ion pairs, which enhances the solubility of salts in non-aqueous media and stabilizes reactive intermediates in organometallic reactions. This coordination is particularly evident in cyclic ethers like crown ethers, where multiple lone pairs create a cavity that selectively binds alkali metal cations, but the principle extends to simple dialkyl ethers in general solvation processes.

Special Cases

Stereogenic Lone Pairs

A stereogenic lone pair refers to a non-bonding on a central atom that functions as one of four distinct substituents in a tetrahedral arrangement, thereby creating a chiral center capable of existing as stable enantiomers. This phenomenon occurs in molecules where the central atom, typically a second- or third-row element like or , adopts an AX3E VSEPR , with the lone pair occupying one vertex of the and three different ligands attached to the others. For to persist, the pyramidal inversion barrier must be sufficiently high to prevent rapid at ambient temperatures; for instance, sulfoxides exhibit barriers around 40-50 kcal/mol due to the strong S=O bond, ensuring configurational , while phosphines have barriers of approximately 29-35 kcal/mol influenced by substituent sterics and electronic effects. Sulfoxides represent a classic example, where the sulfur atom bears a lone pair, an oxygen, and two dissimilar carbon groups (e.g., methyl phenyl ), rendering the sulfur stereogenic. The first optically active was reported in through resolution of menthyl p-tolyl sulfinate derivatives, confirming the tetrahedral geometry and slow inversion. In the , advancements enabled the practical isolation and application of enantiopure sulfoxides, such as those used in drugs like and , where the chiral sulfur enhances . Similarly, tertiary phosphines (PRR'R'') with three distinct R groups and a stereogenic phosphorus lone pair exhibit , as seen in P-chiral ligands like those derived from o-phenylene phosphoramidites, which maintain configuration due to higher inversion barriers compared to amines. These phosphines have been pivotal since the in transition-metal for enantioselective reactions. Such stereogenic lone pairs find extensive use in asymmetric synthesis, serving as chiral auxiliaries to induce in reactions like aldol additions or Diels-Alder cycloadditions, with sulfoxides often providing up to 99% enantiomeric excess in product formation. For instance, chiral sulfoxides direct the of adjacent carbon centers in syntheses, leveraging the lone pair's influence on molecular and coordination. Resolution of racemic mixtures typically involves diastereoselective with chiral auxiliaries, such as the Andersen method using (-)- for sulfoxides, or enzymatic kinetic resolution with reductases like methionine sulfoxide reductase in non-aqueous media, achieving high enantioselectivity since the 1980s. These methods underscore the lone pair's role in enabling isolable without carbon-based stereocenters, expanding the toolkit for stereocontrolled organic transformations.

Unusual Lone Pairs

In hypervalent molecules, central atoms exceed the octet rule by accommodating more than eight valence electrons, often involving lone pairs that contribute to expanded coordination spheres. For instance, chlorine trifluoride (ClF₃) features a central chlorine atom with three bonding pairs and two lone pairs, resulting in a trigonal bipyramidal electron geometry but a T-shaped molecular geometry where the lone pairs occupy equatorial positions to minimize repulsion. Similarly, xenon tetrafluoride (XeF₄) exhibits an octahedral electron geometry with four bonding pairs and two lone pairs positioned axially, yielding a square planar molecular structure. Certain hypervalent bonding scenarios incorporate three-center-four-electron (3c-4e) interactions, where lone pairs from surrounding atoms delocalize into the central atom's space, effectively mimicking or supplementing traditional lone pair roles without invoking d-orbital participation. These 3c-4e bonds stabilize structures in compounds like interhalogens and , distributing over three nuclei with four electrons, as seen in models for molecules such as SF₄. The manifests in heavier p-block elements, where the ns² becomes increasingly reluctant to participate in bonding due to poor overlap with orbitals and relativistic stabilization of the s orbital, favoring lower s. In (I) compounds like Tl⁺, the 6s² lone pair remains unshared, directing a preference for the +1 over +3 and inducing structural distortions in coordination polyhedra, as evidenced by elongated Tl-O bonds in TlAlSiO₄. Carbenes represent another atypical case, where a divalent carbon bears a lone pair alongside two substituents, leading to distinct and triplet ground states depending on . In carbenes like methylene (CH₂), the lone pair occupies a σ orbital (sp² hybridized), enabling nucleophilic behavior, while triplet carbenes feature two unpaired electrons in orthogonal p orbitals, conferring character. Spectroscopic techniques provide direct evidence for these unusual lone pair orbitals. Photoelectron spectroscopy of carbenes reveals ionization potentials from the lone pair orbital, spanning approximately 6 eV across related low-valent carbon species, confirming the energetic distinction between σ-lone pair and π-vacant configurations. In hypervalent systems like XeF₄, (NMR) studies show that lone pair electrons on the central atom influence paramagnetic spin-orbit contributions, supporting their role in overall electronic structure.

Theoretical Descriptions

Multiple Lone Pairs in Different Models

In the Valence Shell Electron Pair Repulsion (, modeling molecules with multiple lone pairs introduces challenges related to whether the lone pairs are equivalent or non-equivalent in their spatial arrangement and repulsive interactions. Equivalent lone pairs, as seen in XeF4, occupy trans positions in an octahedral electron geometry (AX4E2), leading to a square planar molecular shape where the two lone pairs are symmetrically identical and exert balanced repulsion on the bonding pairs. Non-equivalent lone pairs arise in cases where their positions differ due to varying hybridization influences, complicating predictions of bond angles and geometry distortions. provides an approximation here by suggesting that lone pairs, behaving like highly electronegative substituents, receive greater s-character in the hybrid orbitals, which can lead to more p-character in bonding orbitals and altered repulsions in multi-lone pair systems. Descriptive models like the Gillespie-Nyholm approach refine VSEPR by quantifying repulsion strengths, positing that lone pair-lone pair repulsions are stronger than lone pair- pair repulsions, which in turn exceed pair- pair repulsions. This hierarchy is primarily qualitative in standard VSEPR applications but allows for more nuanced treatments in multi-lone pair systems, where quantitative adjustments account for angular distortions beyond ideal geometries. For instance, in OF2 (AX2E2), the two equivalent lone pairs a bent molecular shape with a bond angle of approximately 103°, smaller than the 104.5° in H2O due to stronger effective lone pair repulsions influenced by the higher of , which concentrates on the lone pairs. Similarly, ClF3 (AX3E2) adopts a T-shaped based on trigonal bipyramidal electron arrangement, with the two lone pairs positioned equatorially to minimize 90° interactions, though the description highlights one axial and two equatorial bonding pairs, resulting in compressed angles around 87° for axial-equatorial bonds. These examples illustrate how VSEPR handles multiple lone pairs through prioritized positioning to reduce overall repulsion, yet the model remains largely qualitative for precise energy calculations in complex systems. VSEPR encounters limitations when applied to molecules with multiple lone pairs in transition metal compounds or molecular clusters, where d-orbital participation and multicenter bonding disrupt the simple electron pair repulsion assumptions. For transition metals, geometries like square planar Ni(II) complexes (e.g., Ni(CN)42-) defy tetrahedral predictions for AX4E0, as crystal field effects dominate over lone pair repulsions. In clusters such as B5H9, delocalized electrons and three-center bonds invalidate pairwise repulsion models, requiring more advanced quantum descriptions.

Valence Bond and Molecular Orbital Views

In valence bond (VB) theory, lone pairs are conceptualized as localized pairs of electrons occupying non-bonding hybrid atomic orbitals on the central atom. This approach emphasizes the formation of hybrid orbitals to achieve optimal overlap for bonding while accommodating lone pairs in the remaining hybrids. For instance, in (NH₃), the atom hybridizes its 2s and three 2p orbitals to form four equivalent sp³ hybrid orbitals, with three participating in σ-bonds to hydrogen atoms and the fourth holding the lone pair. This localization aligns with the directional properties of hybrid orbitals, providing a framework for predicting through hybridization schemes. In contrast, molecular orbital (MO) theory describes lone pairs as electrons in non-bonding molecular orbitals, which are linear combinations of atomic orbitals that do not contribute significantly to bonding but may have higher energy relative to bonding orbitals. These non-bonding orbitals, often denoted as n-orbitals, can participate in delocalization, particularly in conjugated systems. For example, in pyridine, the nitrogen lone pair resides in a non-bonding σ-type orbital (7a₁ symmetry) in the plane of the ring, distinct from the π-system and available for interactions like protonation. In carbonyl compounds, one of the oxygen lone pairs occupies a non-bonding orbital that can conjugate with the C=O π* antibonding orbital, contributing to resonance stabilization and influencing reactivity, as seen in the delocalized electron density in amides. The two theories differ fundamentally in their treatment of lone pairs: VB theory prioritizes localized hybridization to explain bond angles and steric effects, while MO theory highlights delocalization and orbital symmetries, often revealing non-equivalent lone pairs. A illustrative comparison is found in water (H₂O), where VB theory posits two equivalent lone pairs in sp³ hybrid orbitals on oxygen, predicting a tetrahedral arrangement. In MO theory, however, the lone pairs occupy distinct non-bonding orbitals: the 1b₁ (pure p-like, out-of-plane/perpendicular to the molecular plane) and 3a₁ (with mixed s-p character, in-plane/within the molecular plane), leading to non-equivalence. Photoelectron spectroscopy supports the MO perspective, showing distinct ionization energies for these orbitals (approximately 12.6 eV for 1b₁ and 14.7 eV for 3a₁), confirming their differing energies and compositions rather than the VB equivalence. Both views build upon Lewis structures as a starting point for electron pair assignments.

References

  1. [1]
    CH150: Chapter 4 - Covalent Bonds and Molecular Compounds
    Rather than being shared, they are considered to belong to a single atom. These are called nonbonding pairs (or lone pairs) of electrons.
  2. [2]
    Molecular Geometry
    Lone pairs of electrons are assumed to have a greater repulsive effect than bonding pairs. Because of the nonbonding pairs of electrons are spread over a ...
  3. [3]
    6.2 Using the VSEPR Model to Determine Molecular Geometry
    Because a lone pair is not shared by two nuclei, it occupies more space near the central atom than a bonding pair (Figure 6.2(d) "The Difference in the Space ...
  4. [4]
    Lewis Acid/Base theory
    The Lewis theory deals with the way in which a substance accepts or donates a lone pair of electrons in an acid-base type of reaction.
  5. [5]
    [PDF] Lecture B2 VSEPR Theory - UCI Department of Chemistry
    Valence shell electron pair repulsion (VSEPR) theory is a model in chemistry used to predict the shape of individual molecules based upon the extent of ...
  6. [6]
    [PDF] Review of VSEPR and Polarity - De Anza College
    1) Write the Lewis structure of the molecule or polyatomic ion. 2) Determine the number of bonding pairs and lone pairs around the central atom, then assume ...
  7. [7]
    [PDF] Chemistry, Structure, and Function of Lone Pairs in Extended Solids
    Jan 1, 2022 · Important properties associated with s2 electron-derived lone pairs include their role in creating conditions favorable for ion transport, in ...
  8. [8]
    [PDF] INTRODUCTION TO LEWIS ACID-BASE CHEMISTRY
    They frequently contain atoms that have nonbonding electrons, or lone pairs. On the other hand, Lewis acids frequently contain atoms with an incomplete octet, ...
  9. [9]
    lone pair (L03618) - IUPAC Gold Book
    A lone pair, also called an electron pair, is two paired electrons localized in the valence shell on a single atom. It is designated with two dots.Missing: definition | Show results with:definition
  10. [10]
    THE ATOM AND THE MOLECULE. - ACS Publications
    Chemistry, Structure, and Function of Lone Pairs in Extended Solids. ... Pair Repulsions in the Cation Affinity and Lewis Acid/Lewis Base Interactions. ACS ...Missing: definition | Show results with:definition
  11. [11]
  12. [12]
    [PDF] Lecture B1 Lewis Dot Structures and Covalent Bonding
    Covalent bonds are drawn as lines. Lone pairs are drawn as pairs of dots. I2 molecule Page 9 Lewis dot structures provide a simple, but extremely powerful, ...
  13. [13]
    Lewis Structures Part 1 - EdTech Books
    Most structures—especially those containing second row elements—obey the octet rule, in which every atom (except H) is surrounded by eight electrons.
  14. [14]
    Lewis Dot Structures - Chemistry 301
    So for example if we look at CO2 each oxygen has two lone pairs (4 electrons) and 2 bonds (double bond). Oxygen has 6 valence electrons. So the formal ...
  15. [15]
    [PDF] Writing Lewis Structures Using the NASL Method.
    Lstands for the number of electrons that are left over. ⇒ L = A − S Divide this number by 2 to get the number of lone pairs in the structure. When using ...
  16. [16]
    10.3: VSPER Theory- The Effect of Lone Pairs - Chemistry LibreTexts
    Sep 9, 2018 · Lone pairs are shown using a dashed line. (CC BY-NC-SA; anonymous) ... represented as. Figure 10 . 3 . 2 ). The three oxygens are ...
  17. [17]
    Drawing the Lewis Structure for SF 6
    ... expanded octet and is able to have more than 8 valence electrons. For the SF6 Lewis structure there are a total of 12 valence electrons on the Sulfur (S) atom.Missing: lone pairs
  18. [18]
    [PDF] Drawing Lewis Dot Structures - East Central College
    The lines each represent a shared pair of electrons. Page 6. Step 5: Add lone pairs so that each atom has its “octet”. H−. O−H. Page 7. Example II: Carbon ...
  19. [19]
    Basic Lewis Structure "Rules"
    Where FC is the formal charge, V is the number of valence electrons, L is the number of lone pair electrons, S is the number of shared electrons. For more on ...
  20. [20]
    Inorganic stereochemistry - Quarterly Reviews, Chemical Society ...
    Inorganic stereochemistry. R. J. Gillespie and R. S. Nyholm, Q. Rev. Chem. Soc., 1957, 11, 339 DOI: 10.1039/QR9571100339. To request permission to reproduce ...
  21. [21]
  22. [22]
    [PDF] Lone pairs affect polarity
    Oct 20, 2019 · Lone pairs contribute to localize negative charge on an atom. The effect of lone pairs alone is why the T‐shaped molecule I4 is polar, even.
  23. [23]
    Molecular Dipole - The Overall Polarity of the Molecule
    It is important to mention that lone pairs of electrons also affect the magnitude and direction of the molecular dipole. For example, the water molecule has two ...
  24. [24]
    Ammonia | NH3 | CID 222 - PubChem - NIH
    Online Posting Date: April 16, 2010. Hazardous Substances Data Bank (HSDB). Dipole moment, gas: 4.9x10-30 C m; 1.47 D.
  25. [25]
    Water | H2O | CID 962 - PubChem - NIH
    Apparent dipole moment: 6.24X10-30 C.m. Morgan JJ et al; Kirk-Othmer Encyclopedia of Chemical Technology. (1999-2014). New York, NY: John Wiley & Sons ...
  26. [26]
    Experimental data for BF 3 (Borane, trifluoro-)
    Calculated electric quadrupole moments for BF3 (Borane, trifluoro-). Electric dipole polarizability (Å3) polarizability. alpha, unc. Reference. 2.420, 1998Gus ...
  27. [27]
    Dipole Moment of NF3 - AIP Publishing - American Institute of Physics
    M. Mashima; Dipole Moment of NF3, The Journal of Chemical Physics, Volume 24, Issue 2, 1 February 1956, Pages 489, https://doi.org/10.1063/1.1742528.
  28. [28]
    Amines as Nucleophiles - Chemistry LibreTexts
    Jan 22, 2023 · All amines contain an active lone pair of electrons on the very electronegative nitrogen atom. It is these electrons which are attracted to ...
  29. [29]
    16.9: Lewis Acids and Bases - Chemistry LibreTexts
    Jul 12, 2023 · Ammonia is both a Brønsted and a Lewis base, owing to the unshared electron pair on the nitrogen. ... lone pair of electrons. Lewis's ...
  30. [30]
    13.4: Hydrogen Bonding - Chemistry LibreTexts
    Jan 20, 2025 · + hydrogens and lone pairs so that every one of them can be involved in hydrogen bonding. This is why the boiling point of water is higher than ...Water as a "perfect" example... · Hydrogen bonding in alcohols
  31. [31]
    A Review of Solvate Ionic Liquids: Physical Parameters ... - Frontiers
    Apr 17, 2019 · The solvation occurs when the lone pairs on the oxygen atoms of the ether moieties act as a Lewis base, donating electrons to the Lewis acid ...
  32. [32]
    15.10: Crown Ethers - Chemistry LibreTexts
    May 30, 2020 · The cation is stabilized by interacting with lone pairs of electrons on the surrounding oxygen atoms. Thus crown ethers solvate cations inside a ...
  33. [33]
    Modern Stereoselective Synthesis of Chiral Sulfinyl Compounds
    Apr 29, 2020 · Among their counterparts in which the place of lone electron pair is taken by a ═NR′ fragment, sulfoximines require two different carbon groups ...
  34. [34]
    [PDF] The Synthesis of P-Stereogenic Phosphorus Compounds
    • “Chiral phosphines bearing stereogenic phosphorus atoms are prone to ... – Delocalization of the lone pair lowers this barrier. • In practical terms ...
  35. [35]
  36. [36]
    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.
  37. [37]
    [PDF] 1 Asymmetric Synthesis of Chiral Sulfoxides - Wiley-VCH
    The first example of an optically active sulfoxide was described in 1926 [1]. This discovery was helpful for discussions regarding the nature of the SaO bond ...Missing: history | Show results with:history
  38. [38]
    Unconventional Biocatalytic Approaches to the Synthesis of Chiral ...
    Sep 11, 2020 · Enantiopure sulfoxides are also used in chemistry as chiral ... In the 1980s, Zaks and Klibanov demonstrated how the enzyme activity in ...
  39. [39]
    Theory of Hypervalency: Recoupled Pair Bonding in SFn (n = 1−6)
    Hypervalent bonding occurs when it is “energetically favorable” (9) to decouple a pair of electrons in order to form a new bond. Much of the insight gained by ...
  40. [40]
    Application of three-center-four-electron bonding for structural and ...
    We show that the structures and stabilities of main group closed-shell hypervalent molecules with five and six valence electron pairs around the central atom
  41. [41]
    [PDF] The role of inert pairs in exclusion of Tl from silicate minerals - RRuff
    The geometrical data indicate that the inert pair causes distortion of the Tl-polyhedron.
  42. [42]
    A ∼6 eV Range of Ionization Potentials among Carbenes, Ylides ...
    Dec 2, 2022 · Unlike its cousin, the conventional carbene, the carbone possesses two free electron lone pairs. In combination with their unique allene ...
  43. [43]
    VSEPR calculation for water, OH 2 - University of Sheffield
    Since the lone pairs are not 'seen', the shape of water is bent. The two lone pairs compress the H-O-H bond angle below the ideal tetrahedral angle to 104.5°.
  44. [44]
    ClF 3 - VSEPR Theory - University of Bristol
    The lone pairs occupy equatorial positions on the trigonal bipyramid. · The equatorial-axial F–Cl–F bond angles are less than 90° due to lone-pair:bonding-pair ...
  45. [45]
    Valence Bond and Molecular Orbital: Two Powerful Theories that ...
    Nov 18, 2021 · As the names suggest, valence bond theory considers a molecule as a collection of bonds, whereas molecular orbital theory views the molecule as ...
  46. [46]
    We Need To Update the Teaching of Valence Theory
    Jan 13, 2012 · According to Lennard-Jones, (16) these two kinds of electron pairs (lone pairs and bond pairs) are localized and mutually repulsive; then, with ...Why It Took So Long · Atomic Orbitals · Molecular Orbitals
  47. [47]
    Ionization of pyridine: Interplay of orbital relaxation and electron ...
    Jun 26, 2017 · A characteristic feature of the pyridine molecule is the non-bonding σ-type lone-pair orbital of the nitrogen atom (7a1). This orbital ...
  48. [48]
    Lone pair orbitals and their interactions studied by photoelectron ...
    Lone pair orbitals and their interactions studied by photoelectron spectroscopy. ... MO method. Journal of Molecular Structure: THEOCHEM 1990, 208 (1-2) ...<|control11|><|separator|>