Fact-checked by Grok 2 weeks ago

Intramolecular force

Intramolecular forces are the chemical bonds that hold atoms together within a single or ionic , primarily consisting of covalent bonds, ionic bonds, and metallic bonds. These forces are significantly stronger than intermolecular forces, which operate between separate molecules, and they determine the structure, stability, and properties of individual chemical entities. Covalent bonds, a primary type of intramolecular force, involve the sharing of pairs between atoms, typically between nonmetals, resulting in stable molecular structures such as those in (H₂O) or (CH₄). These bonds can be polar, where electrons are shared unequally due to differences in , creating partial charges, or nonpolar, with equal sharing as in diatomic molecules like O₂. Ionic bonds, another key intramolecular force, form through the electrostatic attraction between oppositely charged ions, usually between metals and nonmetals, as seen in (NaCl), where electrons are transferred rather than shared. Metallic bonds occur in solid metals, where positively charged metal ions are surrounded by a "sea" of delocalized electrons, enabling properties like electrical conductivity and malleability in substances such as or iron. The strength and nature of intramolecular forces fundamentally influence chemical reactivity, , and the physical properties of substances, such as melting points and , by dictating how atoms are arranged and interact internally. In contrast to weaker intermolecular attractions like hydrogen bonding or van der Waals forces, intramolecular bonds require substantial to break, underscoring their role in maintaining the integrity of compounds under normal conditions. Understanding these forces is essential in fields ranging from , where they affect mechanical strength, to biochemistry, where they stabilize protein structures and molecules.

Introduction

Definition and Scope

Intramolecular forces are the attractive interactions that bind atoms together to form molecules or extended structures, such as crystal lattices, primarily through ionic, covalent, and metallic bonds, which dictate the and of these entities. These forces operate within a single chemical unit, contrasting with weaker interactions between separate units, and are fundamentally electrostatic in nature for ionic bonds or involve shared electrons for covalent and metallic bonds. The scope of intramolecular forces encompasses both discrete molecules, like (H₂O) where covalent bonds link and oxygen atoms, and infinite lattices, such as (NaCl) crystals where ionic attractions form a repeating network. In discrete cases, these forces create stable, finite arrangements, while in lattices, they extend throughout the material, providing rigidity and high melting points. This distinction highlights their role as primary bonds essential for material integrity, with covalent examples in diatomic molecules like (HCl) illustrating shared electron pairs, and ionic examples in NaCl demonstrating between atoms. The concept of intramolecular forces, particularly covalent bonding via electron-pair sharing or transfer, was first systematically recognized by in 1916, revolutionizing understanding of atomic interactions beyond mere electrostatic models. Lewis's framework emphasized how these forces achieve stable electron configurations, laying the groundwork for modern chemical bonding theory without delving into .

Distinction from Intermolecular Forces

Intramolecular forces are the attractive interactions that hold atoms together within a single , forming stable chemical bonds such as covalent, ionic, or metallic bonds, with typical energies ranging from 100 to 1000 kJ/. In contrast, intermolecular forces operate between separate molecules, influencing their aggregation and physical properties like and points, with energies generally much lower, between 1 and 50 kJ/. This fundamental difference in scope and strength distinguishes the two: intramolecular forces define the molecular structure and chemical identity, while intermolecular forces govern how molecules interact in bulk matter. The primary types of intermolecular forces include van der Waals forces (encompassing London dispersion forces and dipole-dipole interactions), and , which arise from temporary or permanent dipoles between molecules. These forces are significantly weaker than intramolecular bonds—often by a factor of 10 to 100—allowing molecules to separate during phase changes like or without disrupting the internal atomic connections. For instance, the required to vaporize (overcoming intermolecular hydrogen bonds at about 41 kJ/mol) is far less than that needed to break its covalent O-H bonds (927 kJ/mol), ensuring the molecule remains intact. This hierarchy of strengths has key implications for chemical behavior: intramolecular forces establish the fixed composition and reactivity of a substance, whereas intermolecular forces determine properties such as , , and phase transitions, enabling phenomena like in solvents without chemical alteration.

Types of Intramolecular Forces

Ionic Bonds

Ionic bonds form through the complete transfer of one or more electrons from a metal to a , resulting in the creation of positively charged cations and negatively charged anions that are held together by strong electrostatic attractions. This typically occurs between elements with large differences in , such as alkali metals and , leading to ions with electron configurations for stability. In ionic solids, these bonds manifest as extended structures rather than discrete molecules, where each cation is surrounded by multiple anions and vice versa in a repeating three-dimensional array. This arrangement contributes to characteristic properties, including high melting points due to the strong collective electrostatic forces required to disrupt the ; for instance, (NaCl) melts at 801°C. Ionic compounds are generally brittle solids at and exhibit poor electrical conductivity in the solid state but become conductors when molten or dissolved in . Common examples of ionic bonds include alkali halides such as (KBr), where the cation (K⁺) and anion (Br⁻) form a face-centered cubic similar to NaCl. Ionic bonds also occur in compounds involving polyatomic ions, such as the sulfate ion (SO₄²⁻) in salts like (Na₂SO₄), where the polyatomic anion consists of covalent bonds within the SO₄ group but interacts ionically with cations. The stability of ionic bonds in a lattice is quantified by lattice energy (U), which represents the energy released when gaseous ions combine to form the solid crystal and is given approximately by the formula U = -\frac{k Q_1 Q_2}{r} where k is Coulomb's constant ($8.99 \times 10^9 \, \mathrm{N \cdot m^2 / C^2}), Q_1 and Q_2 are the charges on the ions, and r is the distance between their centers. This expression derives from Coulomb's law applied to ion pairs, but for the full lattice, a Madelung constant accounts for the summation over all ion interactions in the crystal structure. Factors affecting lattice energy and thus bond stability include the magnitude of the ion charges (higher charges increase |U|), the ionic radii (smaller r yields higher |U| due to closer proximity), and the crystal geometry (e.g., rock salt vs. cesium chloride structures influence the effective coordination). Greater lattice energy corresponds to stronger ionic bonds and enhanced compound stability.

Covalent Bonds

Covalent bonds form through the sharing of one or more pairs of electrons between two atoms, typically nonmetals, to achieve stable electron configurations. These bonds can be (one shared pair), (two shared pairs), or triple (three shared pairs), with the multiplicity influencing bond strength and molecular properties. Covalent bonds are classified as nonpolar or polar based on the equality of electron sharing. In nonpolar covalent bonds, electrons are shared equally between atoms of identical or very similar , resulting in no charge separation; for example, the O=O in dioxygen (O₂) exemplifies this equal sharing. In contrast, polar covalent bonds involve unequal sharing, where the more electronegative atom attracts electrons more strongly, creating partial positive (δ⁺) and negative (δ⁻) charges; (HCl) illustrates this, with pulling electrons closer due to its higher electronegativity than . A key characteristic of covalent bonds is their directionality, arising from the overlap of atomic orbitals along specific axes, which dictates molecular geometries. This directionality leads to discrete molecular structures rather than extended lattices, making covalent bonds prevalent in molecular compounds like gases, liquids, and low-melting solids. For instance, in (CH₄), four directional C-H single bonds adopt a tetrahedral with bond angles of 109.5°, minimizing electron repulsion. Similarly, ethene (C₂H₄) features a planar structure with a C=C , where one and one pi component enforce 120° angles around each carbon. The degree of polarity in covalent bonds is determined by the difference in between bonded atoms, quantified on the Pauling scale, where has the highest value of 4.0 and carbon 2.5. A small difference (e.g., 0.4 or less) yields nonpolar bonds, while larger differences (0.4 to 1.7) produce polar bonds. This unequal sharing generates a bond moment, given by \mu = q \cdot d, where q is the magnitude of the separation and d is the distance between the charges. In polar bonds like C-F, the significant electronegativity difference (1.5) results in a substantial dipole, with bearing δ⁻ and carbon δ⁺.

Metallic Bonds

Metallic bonds arise from the electrostatic attraction between positively charged metal ions arranged in a and a surrounding "sea" of electrons that are free to move throughout the . In this model, metal atoms contribute their electrons to a shared pool, which binds the cations together nondirectionally, distinguishing from more localized interactions in other solids. This electron arrangement is unique to metals and certain alloys, enabling the observed in these materials. The characteristics of metallic bonds directly stem from the mobility of these delocalized electrons. Metals display high electrical conductivity because the free electrons can readily respond to an applied , facilitating current flow. Similarly, thermal conductivity is elevated as these electrons efficiently transfer heat energy across the . Malleability and arise from the ability of layers to slide over one another without breaking the network, as the electron sea maintains cohesion. Additionally, the luster of metals results from free electrons absorbing incident and re-emitting it across visible wavelengths, producing a shiny . Examples of metallic bonds include pure metals such as (Cu), where the single 4s per atom delocalizes to form the bonding sea. In alloys like , a of and (Cu-Zn), the persists through the combined valence electrons of both metals, often yielding enhanced strength and resistance compared to pure components. The band theory provides a qualitative framework for understanding this : atomic orbitals overlap extensively in the solid, forming broad energy bands that extend across the . In metals, these conduction bands are partially filled with electrons, permitting easy excitation and movement that underpins and other properties.

Bond Characteristics

Bond Energy and Strength

For covalent bonds, bond energy, also known as , is defined as the standard enthalpy change associated with the homolytic of a in the gas phase, corresponding to the AB(g) → A(g) + B(g) at 298 K. This quantity quantifies the energy required to break one of bonds, providing a measure of the bond's stability. For instance, the average for a C-H bond is 413 kJ/mol. For ionic bonds, strength is typically measured by , the change when gaseous ions form a solid ionic lattice, influenced by ion charges and sizes via the ; for example, NaCl has a of about 788 kJ/mol. Metallic bond strength is assessed by cohesive energy, the energy to separate the solid metal into gaseous atoms, depending on the number of delocalized electrons and atomic size; for , it is approximately 339 kJ/mol. Several factors influence bond strength across types. Higher bond order generally results in stronger bonds due to increased electron sharing; for example, the C≡C has a dissociation energy of approximately 839 kJ/mol, compared to 614 kJ/mol for C=C and 347 kJ/mol for C-C. Smaller atomic size enhances orbital overlap, leading to stronger bonds, as seen in the progression from H-F (565 kJ/mol) to H-I (298 kJ/mol). Differences in electronegativity between bonded atoms can affect bond polarity and thus stability, with greater differences promoting partial ionic character that may strengthen the bond in some cases. Bond energies for covalent bonds are experimentally determined using techniques such as , which measures heat changes in bond-breaking s, and , which analyzes energy levels from molecular vibrations or electronic transitions. bond energies, derived from multiple compounds, are used for estimating reaction enthalpies, whereas specific bond dissociation energies account for the unique molecular environment of a particular bond. energies for ionic compounds are calculated theoretically or measured indirectly via Born-Haber cycles, while metallic cohesive energies are determined from enthalpies and data.
Bond TypeTypical Energy Range (kJ/mol)Measurement Type
Ionic700–1100Lattice energy per (e.g., for halides)
Covalent100–1000Average dissociation energy per
Metallic100–400Cohesive energy per mole of atoms (e.g., for metals)
These ranges highlight the relative strengths across intramolecular bond types, with ionic and higher-order covalent bonds often exhibiting the greatest ; note that values are not directly comparable due to differing measurement contexts.

Bond Length and Order

Bond length refers to the distance between the nuclei of two bonded atoms, representing the average separation where the attractive and repulsive forces balance. For covalent bonds in molecules, for example, the carbon-carbon single bond (C-C) in measures approximately 154 pm, while the carbon-carbon (C=C) in ethene is about 134 pm, illustrating how multiple bonds result in shorter distances due to increased between the nuclei. In ionic compounds, interionic distances (e.g., 281 pm for Na-Cl in NaCl) reflect lattice packing and sizes, determined by the sum of ionic radii. For metallic bonds in lattices, the nearest-neighbor distance (e.g., 255 pm in ) indicates the spacing in the . These lengths are dynamic, with atoms vibrating around this position. Bond order quantifies the number of electron pairs shared between atoms and can be an integer for simple bonds (e.g., 1 for single, 2 for double) or fractional in cases involving delocalization, such as the 1.5 order for each carbon-carbon bond in due to . exhibits an inverse relationship with : higher orders lead to shorter bonds because greater orbital overlap concentrates and pulls nuclei closer together. This correlation also implies that shorter bonds tend to have higher energies, though the focus here is on structural dimensions. For ionic and metallic bonds, bond order is not typically defined, but effective coordination numbers describe bonding multiplicity. Several factors influence . Atomic radii play a key role, as larger atoms result in longer ; for instance, bond lengths decrease across a in the periodic due to shrinking atomic radii and increase down a group with expanding radii. Hybridization affects length through the s-character of orbitals: involving sp-hybridized atoms (50% s-character) are shorter than those with sp² (33% s-character) or sp³ (25% s-character), as higher s-character concentrates electron density closer to the nucleus, exemplified by shorter C-H bonds in alkynes (sp) compared to alkenes (sp²) or alkanes (sp³). delocalizes electrons, effectively increasing and shortening bonds, as seen in where all C-C bonds are 139 pm—intermediate between single and double bonds—rather than alternating lengths. For ionic bonds, lengths follow trends like smaller cations leading to shorter distances due to higher . Bond lengths are measured experimentally using techniques such as for solids, which determines atomic positions from patterns, and spectroscopy methods like for gases, which infers distances from rotational transitions. These measurements reveal , such as progressively shorter bonds for similar types (e.g., C-C) from left to right across periods due to increasing contracting orbitals. diffraction is also used for precise ionic and metallic lattice parameters.

Theories of Bond Formation

Valence Bond Theory

provides a localized description of chemical bonding, where intramolecular forces arise from the overlap of atomic orbitals on adjacent atoms, allowing shared electron pairs to form . Initially developed by and in 1927, the theory applied to explain the in the hydrogen molecule (H₂), demonstrating that bonding results from the symmetric combination of atomic wavefunctions, which lowers the system's energy compared to the separated atoms. This foundational work established the concept of , where electrons from different atoms occupy the same spatial region, stabilizing the bond through constructive interference of their wavefunctions. Linus Pauling significantly extended in the early 1930s by incorporating concepts like orbital hybridization and to better account for molecular geometries and electron delocalization in more complex systems. Hybridization involves the of atomic s, p, and d orbitals to form new hybrid orbitals with directional properties that match observed bond angles; for instance, in (CH₄), the carbon atom uses four equivalent sp³ hybrid orbitals arranged tetrahedrally, each overlapping with a 1s orbital to form bonds at 109.5° angles. Similarly, sp² hybridization in (C₂H₄) explains the planar trigonal geometry around each carbon, with three sp² orbitals forming bonds and one p orbital contributing to the . These models emphasize that bond strength increases with greater orbital overlap, prioritizing end-to-end alignment for bonds. To address cases where a single cannot fully describe bonding, Pauling introduced , representing the actual electronic structure as a superposition of multiple valence bond configurations with similar energies. In (O₃), for example, the molecule is depicted as a of two equivalent structures, each with a and a between oxygen atoms, resulting in equal bond lengths intermediate between single and double bonds (approximately 1.28 Å). This delocalization stabilizes the molecule beyond what a single structure predicts, distributing more evenly and lowering the overall energy. Despite its successes in describing localized bonds and molecular shapes, has limitations, particularly in handling extensive electron delocalization, such as in or conjugated systems, where it requires numerous structures to approximate the true wavefunction, making calculations cumbersome and less accurate qualitatively. It also struggles with quantitative predictions for systems involving ionic contributions or transition states, often overemphasizing localized pairs over the distributed orbitals better captured by alternative approaches.

Molecular Orbital Theory

Molecular orbital (MO) theory provides a quantum mechanical framework for understanding intramolecular bonding by describing electrons as occupying delocalized s formed from the (LCAO). Developed primarily by and Robert S. Mulliken in , this approach treats molecules as systems where atomic orbitals combine to produce molecular orbitals that extend over the entire molecule, rather than being localized between pairs of atoms. In the LCAO method, a molecular orbital wavefunction is expressed as a linear combination of atomic orbitals, such as \psi_{MO} = c_1 \phi_1 + c_2 \phi_2, where \phi_1 and \phi_2 are atomic orbitals from different atoms, and the coefficients c_1 and c_2 are determined by variational principles to minimize the energy. This combination typically yields two molecular orbitals: a bonding orbital of lower energy, where electron density increases between nuclei, and an antibonding orbital of higher energy, where density decreases, promoting nuclear repulsion. A key quantitative aspect of MO theory is the bond order, calculated as \frac{1}{2} (\text{number of bonding electrons} - \text{number of antibonding electrons}), which indicates the strength and multiplicity of the bond. For example, in the oxygen molecule (O₂), with 12 valence electrons filling the molecular orbitals, the bond order is 2, reflecting a . MO theory further explains the of O₂, as the two highest-energy π* antibonding orbitals each contain one , leading to a triplet with two unpaired spins. Beyond diatomic molecules, MO theory extends to extended systems, such as , where the numerous overlapping orbitals form continuous energy bands of molecular orbitals, known as the band structure. This delocalized model underlies the electronic properties of metals and semiconductors, where determine conductivity. Compared to , MO theory offers advantages in describing systems with delocalized electrons, such as conjugated π systems in organic molecules and , where localized orbital pairs fail to capture or band formation effectively.

Applications in Chemistry

In Inorganic Chemistry

In inorganic chemistry, intramolecular forces, particularly ionic bonds, play a crucial role in forming extended crystal s that define the structures and properties of many salts. For instance, (NaCl) adopts a rock salt structure, consisting of a face-centered cubic (FCC) arrangement of chloride ions with sodium ions occupying all octahedral holes, resulting in each ion being surrounded by six oppositely charged neighbors. This ionic maximizes electrostatic attractions while minimizing repulsions, leading to high stability and characteristic properties such as and in polar solvents. Covalent intramolecular forces contribute to robust network solids in inorganic materials, exemplified by (), which exhibits a diamond-like with alternating and carbon atoms in a tetrahedral arrangement, akin to the zinc blende lattice. In this covalent network, each atom forms four strong Si-C bonds, creating an extended three-dimensional framework that imparts exceptional and thermal stability to , making it suitable for applications like abrasives. Unlike ionic crystals, these covalent lattices resist deformation due to the directional nature of the bonds, highlighting the diversity of intramolecular forces in inorganic solids. Metallic bonds, another key intramolecular force, underpin the properties of alloys in , where delocalized electrons provide cohesion and enable desirable mechanical behaviors. In , an of iron and carbon, the in the iron lattice is modified by carbon atoms occupying sites, forming localized covalent interactions with surrounding iron atoms that enhance resistance to deformation and increase overall strength compared to pure iron. This combination of metallic delocalization and covalent reinforcement allows to balance with , essential for structural applications. Coordination compounds further illustrate the interplay of intramolecular forces, where metal-ligand bonds often exhibit coordinate covalent character. In hexaamminecobalt(III) , [Co(NH₃)₆]³⁺, the cobalt(III) center accepts electron pairs from six ligands, forming dative bonds that blend covalent sharing with electrostatic influences from the charged metal . These bonds stabilize the octahedral and influence its color and reactivity, demonstrating how intramolecular forces dictate the architecture of compounds in inorganic systems. Intramolecular covalent forces are central to the electronic properties of inorganic semiconductors, such as , where each atom forms four tetrahedral bonds in a lattice, resulting in a bandgap of 1.12 eV that separates . This bandgap governs silicon's semiconducting behavior, allowing controlled electrical conductivity through thermal excitation or doping, which underpins its use in electronic devices. The directional covalent bonding ensures structural integrity while enabling the tunable optoelectronic properties vital to modern inorganic materials.

In Organic and Biochemistry

In , covalent bonds form the backbone of hydrocarbons, which are fundamental to molecular structure and reactivity. In alkanes, such as (CH₄) and (C₂H₆), carbon atoms are connected by strong (σ) bonds formed through sp³ hybridization, where each carbon uses four equivalent sp³ orbitals to overlap with or other carbon atoms, resulting in tetrahedral and stable single bonds with bond energies around 348 kJ/mol for C-C and 413 kJ/mol for C-H. Aromatic hydrocarbons, like (C₆H₆), feature delocalized pi (π) bonds arising from sp² hybridization of carbon atoms, creating a resonant with alternating double bonds that enhances stability through electron delocalization, as evidenced by benzene's higher heat of compared to expected non-aromatic analogs. Hybridization also governs functional groups; for instance, the carbonyl (C=O) in aldehydes and ketones involves sp² hybridization at carbon, enabling planar and facilitating reactions due to the electrophilic nature of the carbon atom. In biochemistry, intramolecular covalent bonds are essential for the architecture and function of biomolecules. Peptide bonds, which are amide linkages (-CO-NH-), covalently connect amino acids in proteins via a condensation reaction between the carboxyl group of one residue and the amino group of another, forming the polypeptide backbone with partial double-bond character that restricts rotation and promotes planarity. In DNA, phosphodiester bonds covalently link nucleotides by connecting the 5'-phosphate of one deoxyribose sugar to the 3'-hydroxyl of the adjacent sugar, creating the sugar-phosphate backbone that provides structural integrity and directionality for replication and transcription processes. These bonds play a critical role in enzyme active sites, where covalent catalysis often occurs; for example, serine proteases like chymotrypsin form transient covalent intermediates with substrates via the nucleophilic attack of a serine hydroxyl group, lowering activation energies and accelerating hydrolysis reactions by factors exceeding 10⁶. The stability of biomolecules relies on a combination of covalent bonds and intramolecular bonds, though the primary framework is covalent. In protein alpha helices, the covalent backbone is stabilized by intramolecular bonds between the carbonyl oxygen of residue i and the of residue i+4, contributing approximately 4-8 / per to helical stability and enabling compact folding essential for function. Specific examples illustrate these principles: in , the ion in the group forms coordinate covalent bonds with four nitrogen atoms of the ring and a proximal residue, allowing reversible oxygen binding while maintaining . Similarly, in (ATP), the two phosphoanhydride bonds linking the phosphate groups are high-energy covalent linkages with hydrolysis free energies of about -30.5 / each, driving energy-requiring processes like and in cells.

References

  1. [1]
    Teacher Notes: Chemical Bonds and Forces - Sites@Duke Express
    Intramolecular bonds are the bonds that hold atoms to atoms and make compounds. There are 3 types of intramolecular bonds: covalent, ionic, and metallic.
  2. [2]
    [PDF] TutorTube: Intramolecular & Intermolecular Forces Spring 2020
    To start, let's define intramolecular and intermolecular. 'Intra' means within, so intramolecular forces occur within a molecule. 'Inter' means between, so ...
  3. [3]
    [PDF] Gilbert Newton Lewis - Chemistry
    Not until 1916 did Lewis publish his own model, which equated the classical chemical bond with the sharing of a pair of electrons between the two bonded atoms.
  4. [4]
  5. [5]
    Intermolecular Forces
    These forces can be divided into three categories: (1) dipole-dipole, (2) dipole-induced dipole, and (3) induced dipole-induced dipole.<|control11|><|separator|>
  6. [6]
    Ionic Bonding - Chemistry 301
    Ionic bonds are chemical bonds in which we assume the electrons have fully "moved" from one element in the compound to another.Missing: definition | Show results with:definition
  7. [7]
    Ionic Compounds | manoa.hawaii.edu/ExploringOurFluidEarth
    An ionic bond is a bond between ions where oppositely charged atoms attract each other and cancel their charges to produce neutral compounds. Information about ...
  8. [8]
    CH105: Chapter 3 - Ionic and Covelent Bonding - Chemistry
    With two oppositely charged ions, there is an electrostatic attraction between them because opposite charges attract. The resulting combination is the ionic ...
  9. [9]
    Lattice Energies in Ionic Solids
    They are not easily deformed, and they melt at relatively high temperatures. NaCl, for example, melts at 801°C. These properties result from the regular ...
  10. [10]
    5.3 Molecular and Ionic Compounds - Maricopa Open Digital Press
    Ionic compounds are solids that typically melt at high temperatures and boil at even higher temperatures. For example, sodium chloride melts at 801 °C and boils ...
  11. [11]
    Table of Polyatomic Ions - gchem
    Table of Polyatomic Ions ; sulfate, SO ; hydrogen sulfate (aka: bisulfate), HSO ; thiosulfate, S2O ; oxalate, C2O ...
  12. [12]
    Symbols and Names for Common Polyatomic Ions
    Symbols and Names of Some Common Polyatomic Ions and One Molecule. NH4+. ammonium ion. OH-. hydroxide ion. CN-. cyanide ion. SO42-. sulfate ion. O22-.
  13. [13]
    [PDF] Chapter 3- Atomic Structure (light interaction with e-)
    Lattice Energy. • Coulomb's Law (2 charged particles):. E = 2.31 x 10-19 J•nm(q. 1 q. 2. /d). • Lattice Energy U (crystal lattice of charged particles):. U=k(q.
  14. [14]
    Strength of an Ionic Bond - Lattice Energy
    Lattice Energy: Strengths of Ionic Bonds. Although there are several important factors which must be considered in determining the energy associated with ...Missing: affecting stability
  15. [15]
    Strengths of Ionic and Covalent Bonds – Chemistry - UH Pressbooks
    Ionic Bond Strength and Lattice Energy. An ionic compound is stable because of the electrostatic attraction between its positive and negative ions. The ...
  16. [16]
    5.2 Covalent Bonding – Chemistry Fundamentals - UCF Pressbooks
    Ionic bonding results from the electrostatic attraction of oppositely charged ions that are typically produced by the transfer of electrons between metallic ...
  17. [17]
    The Covalent Bond
    The term covalent bond is used to describe the bonds in compounds that result from the sharing of one or more pairs of electrons.
  18. [18]
    Molecular Structure & Bonding - MSU chemistry
    Thus, the four covalent bonds of methane consist of shared electron pairs with four hydrogen atoms in a tetrahedral configuration, as predicted by VSEPR theory.Missing: CH4 C2H4<|control11|><|separator|>
  19. [19]
  20. [20]
  21. [21]
    [PDF] Bonding in Metals, Semiconductors and Insulators - Band Structure
    Characteristic properties of metals include: (1) electrical conductivity, (2) opaqueness and (3) malleability. A very simple model in which the metallic crystal ...Missing: definition | Show results with:definition
  22. [22]
    Applying the soft sphere model to improve the understanding ... - NIH
    Dec 30, 2019 · In this model of metallic bonding, the metal atoms in a metal crystal lose their outer electrons to form a lattice of regularly spaced positive ...Missing: definition | Show results with:definition
  23. [23]
    Bonding in Metals and Semiconductors
    Band theory explains the correlation between the valence electron configuration of a metal and the strength of metallic bonding. The valence electrons of ...Missing: definition | Show results with:definition
  24. [24]
    Energy
    By definition, the bond-dissociation enthalpy for an X-Y bond is the enthalpy of the gas-phase reaction in which this bond is broken to give isolated X and Y ...
  25. [25]
    Bonding Frameset
    Since there are four bonds each bond contributes a certain amount of stability. 1652 kJ / 4 bonds = 413 kJ per CH bond.Missing: average | Show results with:average
  26. [26]
    Covalent Bond Energies - gchem
    Covalent bond energies are in kJ/mol. For example, H-H is 432, H-C is 413, C-C is 347, and C=C is 614.Missing: CH | Show results with:CH
  27. [27]
    [PDF] Models of Chemical Bonding - Web Pages
    bond order. :N≡N: 945 kJ/mol. :O=O: 498 kJ/mol. :F–F: 159 kJ/mol. • In general, bond strength (∆HB) increases with decreasing the size of the bonded atoms. H–F.Missing: affecting | Show results with:affecting
  28. [28]
    CHEM 125b - Lecture 4 - Electronegativity, Bond Strength ...
    Bonds involving atoms with lone-pair electrons are weakened by electron-pair repulsion. Electronegativity differences between atoms make ionic dissociation ( ...Missing: affecting | Show results with:affecting
  29. [29]
    [PDF] Pogil Bond Energy
    Bond energy, also known as bond dissociation energy, is the amount of energy required to break one mole of a specific chemical bond in a gaseous molecule, ...
  30. [30]
    6.4 Strengths of Ionic and Covalent Bonds – Chemistry Fundamentals
    Whereas lattice energies typically fall in the range of 600–4000 kJ/mol (some even higher), covalent bond dissociation energies are typically between 150–400 kJ ...
  31. [31]
    Bond Length Definition in Chemistry - ThoughtCo
    Jan 13, 2020 · In chemistry, bond length is the equilibrium distance between the nuclei of two groups or atoms that are bonded to each other.
  32. [32]
    15.2: Structure and Stability of Benzene - Chemistry LibreTexts
    Mar 17, 2024 · The 139 pm bond length is roughly in between those of a C=C double bond (134 pm) and a C-C single (154 pm) which agrees with the benzene ring ...
  33. [33]
    8.9: Covalent Bond Properties: Order, Length, and Energy
    May 7, 2022 · It needs to be understood that a bond is dynamic, with the atoms oscillating around an equilibrium distance, which we call the bond length.
  34. [34]
    Bond Order for Covalent Bonds | Organic Chemistry Tutorial
    The net bond order is 1.5 for one-and-a-half bonds. Another example is nitrate anion (NO3-), where nitrogen forms single and double covalent bonds with oxygen ...<|separator|>
  35. [35]
    Hybridization, Bond Lengths, Bond Strengths, and Bond Angles
    Jul 31, 2014 · Bond order and length are inversely proportional to each other: when bond order is increased, bond length is decreased.
  36. [36]
    Bond Parameters - BYJU'S
    For covalent bonds, the bond length is inversely proportional to the bond order – higher bond orders result in stronger bonds, which are accompanied by ...
  37. [37]
    Orbital Hybridization And Bond Strengths - Master Organic Chemistry
    Jan 19, 2018 · The trend for bond strengths, above, is sp > sp 2 > sp 3. In other words, the more s-character on carbon, the stronger the bond.1. A Quick Quiz · 5. Bonus Section: An Answer... · 6. Homolytic Versus...
  38. [38]
    Knowing more on Periodic Trends in Bond Length - Unacademy
    The bond length trends in the periodic table follow the same trends as atomic radii: bond length decreases across a period and increases down a group.
  39. [39]
    Valence Bond Theory—Its Birth, Struggles with Molecular Orbital ...
    Mar 15, 2021 · This essay describes the successive births of valence bond (VB) theory during 1916–1931. The alternative molecular orbital (MO) theory was born in the late ...
  40. [40]
    "The nature of the chemical bond. Application of results obtained ...
    Application of results obtained from the quantum mechanics and from a theory of paramagnetic susceptibility to the structure of molecules." February 17, 1931. J ...
  41. [41]
  42. [42]
    [PDF] Robert S. Mulliken - Nobel Lecture
    Molecular orbitals also began to appear in a fairly clear light as suitable homes for electrons in molecules in the same way as atomic orbitals for electrons in ...
  43. [43]
    [PDF] Friedrich Hund: A Pioneer of Quantum Chemistry
    Sep 9, 2022 · In view of this, the MO theory became known as the Mulliken-Hund. MO theory. Friedrich Hund was born on February 4, 1896 in Karlsruhe, Ger- many ...
  44. [44]
    1.17: Linear combination of atomic orbitals - Chemistry LibreTexts
    Mar 2, 2025 · Linear combination of atomic orbitals (LCAO) is a simple method of quantum chemistry that yields a qualitative picture of the molecular orbitals ...
  45. [45]
    9.3: Molecular Orbital Theory - Chemistry LibreTexts
    Jun 21, 2024 · LCAO-MO Introduction. This describes molecular orbitals that are the result of the linear combination of atomic orbitals. A standing wave ...Introduction · LCAO-MO Introduction · Bond Order · MO Electron Configurations
  46. [46]
    12.5: Molecular Orbital Theory - Chemistry LibreTexts
    Dec 14, 2018 · The number of electrons in an orbital is indicated by a superscript. In this case, the bond order is (1-0)/2=1/2 Because the bond order is ...Molecular Orbitals Involving... · Energy-Level Diagrams · Bond Order in Molecular...
  47. [47]
    9.10: Molecular Orbital Theory Predicts that Molecular Oxygen is ...
    Jun 30, 2023 · The molecular orbital configuration dictates the bond order of the bond. This in turns dictates the strength of the bond and the bond length ...
  48. [48]
    7.1: Molecular Orbitals and Band Structure - Chemistry LibreTexts
    Mar 9, 2023 · The band model (like MO theory) is based on a one-electron model. This was an approximation we made at the very beginning of our discussion of MO theory.
  49. [49]
    Valence Bond and Molecular Orbital: Two Powerful Theories that ...
    Nov 18, 2021 · However, resonance is an inherent part of VB theory and appeared in Pauling and Wheland's original VB description of benzene. (33) Resonance ...<|control11|><|separator|>
  50. [50]
    2.3: Molecular orbital theory- conjugation and aromaticity
    Jul 20, 2022 · Molecular orbital theory explains bonding in molecules, especially where valence bond theory fails, and is used to explain aromaticity, like in ...
  51. [51]
    6.11A: Structure - Rock Salt (NaCl) - Chemistry LibreTexts
    Feb 3, 2021 · The structure of NaCl is formed by repeating the face centered cubic unit cell. It has 1:1 stoichiometry ratio of Na:Cl with a molar mass of 58.4 g/mol.
  52. [52]
    Ionic Structures - Chemistry LibreTexts
    Jun 30, 2023 · This page explains the relationship between the arrangement of the ions in a typical ionic solid like sodium chloride and its physical properties
  53. [53]
    12.6: Network Covalent Atomic Solids- Carbon and Silicates
    Aug 10, 2022 · The unit cell of diamond can be described as an fcc array of carbon atoms with four additional carbon atoms inserted into four of the ...
  54. [54]
    Lesson 8: Metallic Bonding - Clackamas Community College
    This particular model of metallic bonding be used to explain the properties of metals (such as electrical conductivity, malleability, and thermal conductivity)
  55. [55]
    Scientific Principles
    To form the strongest metallic bonds, metals are packed together as closely as possible. Several packing arrangements are possible. Instead of atoms ...
  56. [56]
    9.9: Bonding in Coordination Complexes - Chemistry LibreTexts
    Nov 13, 2022 · An example of such an ion might be hexaaquotitanium(III), Ti(H2O)63+. The ligands (H2O in this example) are bound to the central ion by electron ...
  57. [57]
    [PDF] Electrons and Holes in Semiconductors
    In a silicon crystal, each Si atom forms covalent bonds with its four neighbors. In an intrinsic Si crystal, there are few mobile electrons and holes. Their ...
  58. [58]
    CH103 - Chapter 5: Covalent Bonds and Introduction to Organic ...
    For example, water (H2O) has a melting point of 4oC and a boiling point of 100oC compared with NaCl that has a melting point of 801oC and a boiling point of ...
  59. [59]
    Hydrocarbons – Chemistry - UH Pressbooks
    Alkanes, or saturated hydrocarbons, contain only single covalent bonds between carbon atoms. Each of the carbon atoms in an alkane has sp3 hybrid orbitals and ...
  60. [60]
    Alkanes & Cycloalkanes - MSU chemistry
    Alkanes and cycloalkanes are termed saturated, because they incorporate the maximum number of hydrogens possible without breaking any carbon-carbon bonds.
  61. [61]
    [PDF] TOPIC 2. ORGANIC COMPOUNDS
    1. Describe the structure of N, O and halogen containing functional groups: formal charges, hybridization, VSEPR theory, polarity, resonance.
  62. [62]
    [PDF] 1-3 The Peptide Bond
    The amino acids of a protein chain are covalently joined by amide bonds, often called peptide bonds: for this reason, proteins are also known as polypeptides.
  63. [63]
    Peptides and Proteins - MSU chemistry
    If the amine and carboxylic acid functional groups in amino acids join together to form amide bonds, a chain of amino acid units, called a peptide, is formed. A ...
  64. [64]
    Biochemistry, DNA Structure - StatPearls - NCBI Bookshelf - NIH
    This bond is known as a phosphodiester bond, and it forms via a condensation reaction during DNA synthesis. As a result, each strand of a DNA molecule has a ...
  65. [65]
    DNA Backbone
    Phosphodiester bonds are the covalent bonds that hold nucleotides together in a single strand of DNA. a. Find a phosphodiester bond in the three-dimensional ...
  66. [66]
    Chapter 7: Catalytic Mechanisms of Enzymes - Chemistry
    Covalent catalysis involves the formation of a covalent bond between the enzyme and at least one of the substrates involved in the reaction. Often times this ...
  67. [67]
    Analysis of forces that determine helix formation in α-proteins - NIH
    In the α-helix in its canonical Pauling form, all hydrogen bonds were formed by intramolecular interactions. The interactions between side chains were not ...
  68. [68]
    End-Capped α-Helices as Modulators of Protein Function - PMC
    A canonical α-helix is characterized by hydrogen bonds between the C=O of the ith amino acid and the NH of the i+4th amino acid of a peptide chain, which ...
  69. [69]
    Hemoglobin
    The Fe is anchored to a globin subunit by a coordinate, or dative, covalent bond to a histidine side chain, termed the F8 or proximal histidine. A coordinate ...
  70. [70]
    Myoglobin, Hemoglobin, and their Ligands - csbsju
    Mar 29, 2016 · The Fe2+ ion is coordinated to 4 N's on the 4 pyrrole rings, The 5th ligand is a supplied by proximal His (the 8th amino acid on helix F) of the ...
  71. [71]
    How Cells Obtain Energy from Food - NCBI - NIH
    Figure 2-75 compares the high-energy phosphoanhydride bonds in ATP with other phosphate bonds, several of which are generated during glycolysis.
  72. [72]
    [PDF] Bioenergetics - Rose-Hulman
    Although it contains high-energy phosphate bonds, the energy in each of the ATP phosphoanhydride bonds is somewhat lower than the energy of a few other ... As ...