Fact-checked by Grok 2 weeks ago

Bond length

Bond length is the between the nuclei of two atoms joined by a , representing the most probable separation in a where the attractive and repulsive forces balance. This is typically measured in picometers (pm, where 1 pm = 10^{-12} m) or angstroms (Å, where 1 Å = 100 pm = 10^{-10} m). Bond lengths are determined experimentally through techniques such as diffraction for solids and for gases, providing essential data for molecular structure analysis. In covalent bonds, bond length varies primarily with bond order, where single bonds are longer than double bonds, and double bonds are longer than triple bonds, due to greater orbital overlap and in higher-order bonds. For instance, the C-C is approximately 154 pm, while the C=C is about 134 pm, and the C≡C is around 120 pm. The atomic radii of the bonded atoms also play a key role, as bonds between larger atoms, such as those in later periods of the periodic table, tend to be longer. Bond length is inversely related to bond strength, with shorter bonds generally exhibiting higher bond dissociation energies because the electrons are held more tightly between the nuclei. This relationship underscores bond length's importance in predicting molecular reactivity, stability, and spectroscopic properties, such as vibrational frequencies in . In ionic compounds, interatomic distances approximate bond lengths but are influenced more by lattice energies and ion sizes rather than shared electron pairs. Overall, bond lengths provide a quantitative measure of bonding interactions across diverse chemical systems.

Fundamentals

Definition and Basic Concepts

Bond length refers to the distance between the nuclei of two atoms involved in a , representing the average separation at which the attractive and repulsive forces between the atoms are balanced. This distance is a fundamental property of , providing insight into the stability and nature of molecular structures. In the context of covalent bonds, bond length arises from the overlap of atomic orbitals, where shared pulls the nuclei closer together while electron-electron and nucleus-nucleus repulsions prevent complete coalescence. The concept applies across different bond types, though its interpretation varies. In covalent bonds, it directly corresponds to the extent of orbital overlap in diatomic or polyatomic molecules. For ionic bonds in crystalline solids, bond length denotes the internuclear distance between oppositely charged ions, governed primarily by electrostatic interactions within the . In metallic bonds, it describes the average spacing between metal atoms in a , facilitated by the delocalized sea of electrons that allows for variable distances without discrete pairing. These variations highlight how bond length encapsulates the electronic and structural characteristics unique to each . It is important to distinguish bond length from related internuclear distances, such as the van der Waals distance, which measures the closest approach between non-bonded atoms or molecules without forming a ; the latter is typically longer due to weaker intermolecular forces. Bond lengths are usually expressed in picometers () or angstroms (), with 1 equaling 100 , facilitating comparisons across compounds. The foundational ideas of bond length trace back to the early 20th century, emerging from Gilbert N. Lewis's 1916 introduction of electron-pair sharing in covalent bonds via dot structures, which visualized bonds as localized pairs achieving stable octets. This was complemented by , pioneered by and in 1927 through quantum mechanical descriptions of orbital overlap in the , and further refined by in the 1930s, who integrated hybridization concepts to explain directional bonding and associated lengths. These developments shifted chemical understanding from empirical valence rules to a quantum framework, establishing bond length as a quantifiable manifestation of electronic interactions.

Units and Notation

Bond lengths are most commonly expressed in picometers (pm), where 1 pm equals 10^{-12} meters, aligning with the (SI) for precise atomic-scale measurements. Alternatively, the (Å), defined as 10^{-10} meters or exactly 100 pm, remains a widely used non-SI unit in chemical literature due to its convenience for bond scales typically ranging from 50 to 300 pm. In structural formulas and diagrams, bond multiplicity is denoted by standard symbols: a as a solid line (-), a as two (=), and a as three (≡), facilitating quick visual representation of without specifying lengths. Scientific databases standardize reporting for consistency; for instance, the NIST Chemistry WebBook presents bond lengths primarily in angstroms (), while the CRC of Chemistry and Physics favors picometers (pm) in its tables of characteristic bond lengths. Historically, bond lengths were often reported in following its introduction in the early for studies, but adoption of the system in the late shifted preference toward picometers, with nanometers (, where 1 = 1000 pm) occasionally used for larger molecular dimensions though less common for individual bonds due to decimal awkwardness (e.g., C-C bonds at ~0.15 ). Conversion between units is straightforward: 1 Å = 100 pm = 0.1 , allowing seamless translation across references. Precision in bond length reporting depends on the measurement technique, with achieving typical accuracies of ±0.01 Å (±1 pm) for heavy-atom bonds under optimal conditions, ensuring reliable comparisons in analyses.

Determination Methods

Experimental Techniques

is one of the most widely used experimental techniques for determining bond lengths in crystalline solids. It relies on the of X-rays by the clouds surrounding nuclei in a , producing patterns that allow for the precise determination of positions. From these positions, interatomic distances corresponding to lengths are calculated, typically achieving accuracies of about 0.001 Å for bonds involving heavy atoms and around 0.01 Å or better for lighter atoms. However, the method measures distances between the centers of rather than nuclear positions, leading to systematic underestimation of lengths involving (e.g., C–H bonds appear ~0.15 Å shorter than actual values). Thermal motion and disorder in the further introduce averaging effects, limiting precision in dynamic structures, and the technique is generally inapplicable to amorphous materials or non-crystalline phases. Neutron diffraction complements X-ray methods, particularly for compounds containing hydrogen or other light elements where X-ray scattering is weak due to low electron density. Neutrons interact directly with atomic nuclei, enabling accurate localization of hydrogen positions and providing true nuclear interatomic distances without the electron cloud bias. For instance, neutron diffraction yields a C–H bond length of approximately 1.09 Å in organic molecules, contrasting with the shorter values from X-ray data. Accuracies are comparable to X-ray for non-hydrogen bonds (often within 0.001–0.005 Å), but the method requires large single crystals and access to specialized facilities like nuclear reactors or spallation sources, making it less routine. Limitations include challenges with sample size, radiation sensitivity, and absorption by certain isotopes, restricting its use primarily to solid-state studies. Microwave spectroscopy determines bond lengths in gas-phase molecules by measuring rotational transitions, from which rotational constants are derived to calculate moments of and thus equilibrium interatomic distances. The technique excels for small, polar molecules, offering high precision (e.g., bond lengths accurate to 0.001 or better) through analysis of spectral line spacings. An example is the carbon monoxide (CO) molecule, where spectroscopic data yield a bond length of about 1.13 , reflecting vibrationally averaged values that can be corrected to equilibrium lengths. However, it is limited to volatile, gaseous and requires the molecule to have a permanent for strong signals, excluding non-polar compounds. Electron diffraction, performed in the gas phase, uses high-energy electron beams scattered by molecular electrons to generate diffraction patterns, from which structural parameters including bond lengths are refined via least-squares fitting. This method is particularly suited for volatile or sublimable compounds, providing bond lengths with uncertainties as low as 0.002 Å for small molecules, such as the 1.484 Å C–C bond in thiirane. It captures dynamic, vibrationally averaged structures in isolation, avoiding solid-state packing effects. Limitations include the need for sufficient vapor pressure, challenges with larger or non-volatile molecules, and sensitivity to molecular vibrations that complicate data interpretation, making it less ideal for complex systems compared to diffraction methods for solids. Overall, these techniques are constrained by phase: and apply to crystalline solids, while and target gas-phase molecules, leaving solution-phase or amorphous materials reliant on indirect methods or approximations. Dynamic bonds in liquids remain particularly challenging due to motional averaging and lack of long-range order.

Theoretical and Computational Approaches

Theoretical approaches to bond length prediction rely on quantum mechanical models that describe the electronic structure of molecules, allowing computation of equilibrium geometries without experimental measurement. These methods, grounded in the , separate nuclear and electronic motions via the Born-Oppenheimer approximation, enabling the calculation of surfaces where bond lengths correspond to energy minima. Valence bond (VB) theory provides a qualitative framework for understanding bond lengths through the overlap of atomic orbitals, where stronger overlap between valence orbitals leads to shorter, more stable bonds due to enhanced electron sharing. In VB descriptions, bond length is influenced by the between contributing structures and the hybridization of orbitals, with better spatial overlap reducing the internuclear distance; for instance, sp hybridization in predicts shorter C-H bonds compared to sp³ in . This approach, originating from early quantum mechanical treatments, emphasizes localized bonding pairs and remains useful for interpreting bond shortening in conjugated systems. Molecular orbital (MO) theory extends this to a quantitative level by constructing delocalized orbitals from linear combinations of atomic orbitals, yielding potential energy curves that reveal the equilibrium bond length as the internuclear distance minimizing the total energy. The bond length r_e is the value of r that minimizes the V(r) = E_\text{electronic}(r) + V_\text{nuclear-nuclear}(r), where E_\text{electronic}(r) is the electronic energy at fixed nuclear separation r obtained by solving the electronic , and the minimum satisfies \frac{dV}{dr} = 0. This method accurately captures effects, such as shorter bonds in species with higher MO occupancy in bonding orbitals, as seen in the progression from N₂ to O₂. Ab initio methods solve the electronic variationally without empirical parameters, with Hartree-Fock () theory serving as the foundational approach by optimizing a single to approximate the wavefunction and compute bond lengths. typically shortens bonds by underestimating electron correlation, yielding mean absolute errors of about 0.01 for small molecules. Post- methods like second-order Møller-Plesset () incorporate dynamic electron correlation, lengthening bonds toward experimental values and reducing errors to around 0.005 , particularly improving multiple bonds where correlation effects are pronounced. Density functional theory (DFT) offers an efficient alternative for larger systems by approximating the exchange-correlation energy functional, enabling bond length predictions with computational cost scaling favorably compared to post-HF methods. The B3LYP, combining Hartree-Fock exchange with Becke's gradient-corrected exchange and Lee-Yang-Parr correlation, is widely adopted for its balance of accuracy and speed, achieving mean unsigned errors of approximately 0.01 for bond lengths in and inorganic small molecules. DFT's efficiency stems from its O(N³) scaling, making it suitable for systems beyond the reach of traditional techniques while maintaining predictive power for equilibrium geometries. Validation of these computational methods involves benchmarking predicted bond lengths against experimental data from techniques like , with high-level approaches such as coupled-cluster with single, double, and perturbative triple excitations [CCSD(T)] serving as gold standards; for small molecules, typical errors are less than 0.01 when using correlation-consistent basis sets like cc-pVQZ.

Influencing Factors

Bond Order and Multiplicity

Bond order is defined as the number of shared electron pairs between two atoms forming a , with a value of 1 for a , 2 for a , and 3 for a ./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Bond_Order_and_Lengths) This parameter directly influences bond length through an inverse relationship: as bond order increases, the bond length decreases because greater electron sharing enhances the attractive forces between the atomic nuclei./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Bond_Order_and_Lengths) In hydrocarbons, representative examples illustrate this trend clearly. A carbon-carbon single bond typically measures about 154 pm, a double bond around 134 pm, and a triple bond approximately 120 pm, reflecting the progressive shortening with each additional shared pair. Bond orders can also be fractional in systems involving resonance, where electron delocalization averages the bonding character across multiple structures. For instance, in benzene, the six carbon-carbon bonds exhibit an average bond order of 1.5 due to resonance between Kekulé structures, resulting in a uniform bond length of about 139 pm—intermediate between single and double bonds. An empirical relation often approximates this inverse dependence as bond length d \approx a - b \times n, where n is the and a and b are constants specific to the atomic pair, such as a \approx 154 pm and b \approx 17 pm for carbon-carbon bonds. This linear model captures the general trend observed in experimental data for multiple bonds, though more precise formulations like Pauling's exponential relation may apply over wider ranges. The impact of bond order extends to allotropes of elements like carbon. In , each carbon atom forms four single bonds (bond order 1) via sp³ hybridization, yielding a C-C bond length of 154 pm. In contrast, features layers of carbon atoms in sp² hybridization with delocalized π electrons imparting partial double-bond character (effective >1), shortening the in-plane C-C bonds to 142 pm./08:_Chemistry_of_the_Main_Group_Elements/8.07:_Group_14/8.7.01:_The_Group_14_Elements_and_the_many_Allotropes_of_Carbon)

Atomic Size and Electronegativity

The length of a is fundamentally influenced by the atomic sizes of the bonded atoms, with larger atoms generally forming longer bonds due to the greater distance between their nuclei. As increases down a group in the periodic table, the bond lengths to a fixed partner atom also increase. For example, the C–F length is approximately 135 pm, whereas the C–Cl is 177 pm, illustrating the effect of the progressively larger atomic radii of (57 pm ) and (102 pm ). In nonpolar covalent bonds, where the electronegativities of the atoms are similar, bond lengths can be reliably estimated using the additivity rule: the bond length approximates the sum of the of the two atoms, d_{AB} \approx r_A + r_B. This approach works well for homonuclear bonds (e.g., at 199 pm, twice the chlorine ) and many heteronuclear ones, providing a baseline for predicting molecular dimensions. Standard , derived from experimental bond length data and refined through quantum chemical calculations, are tabulated below for selected main-group elements; these values assume single bonds in neutral molecules and coordination numbers typical for the group.
ElementSymbolCovalent Radius (pm)
HydrogenH37
CarbonC76
NitrogenN71
OxygenO66
FluorineF57
SiliconSi111
PhosphorusP107
SulfurS105
ChlorineCl102
Electronegativity differences between bonded atoms introduce polarity that modifies bond lengths across the spectrum of bond types. On the Pauling scale, which quantifies an atom's ability to attract electrons in a bond (ranging from cesium at 0.79 to fluorine at 3.98), small differences (ΔEN < 0.5) yield nonpolar covalent bonds with lengths close to the sum of covalent radii; moderate differences (0.5 < ΔEN < 1.7) produce polar covalent bonds that are slightly shorter due to partial ionic character pulling the atoms closer; and large differences (ΔEN > 1.7) result in ionic bonds, where lengths approximate the sum of ionic radii rather than covalent ones. For instance, the C–F bond (ΔEN = 1.43, with carbon at 2.55 and fluorine at 3.98) measures 135 pm, shorter than the 133 pm sum of their covalent radii, owing to the enhanced ionic contribution from fluorine's high electronegativity. In contrast, the ionic Na–Cl distance in solid NaCl is 282 pm, matching the sum of Na⁺ (102 pm) and Cl⁻ (181 pm) ionic radii for sixfold coordination. Exceptions to simple additivity and electronegativity effects occur in hypervalent molecules, where central atoms exceed the and bond polarity extends beyond standard covalent models. In , for example, the S–F bonds are 156 pm long—shorter than the 162 pm sum of (105 pm) and (57 pm) covalent radii—due to the high electronegativity difference (ΔEN = 1.40) promoting greater ionic character in the polar 3-center-4-electron bonds that describe the hypervalent structure.

Molecular Geometry and Hybridization

Molecular geometry plays a crucial role in determining bond lengths through the influence of orbital hybridization and spatial arrangements of atoms. In hybridized orbitals, the percentage of s-character affects the bond length: higher s-character brings the bonding electrons closer to the , resulting in shorter bonds. For carbon atoms, sp³ hybridization in tetrahedral geometries, as in (CH₄), features 25% s-character, leading to longer C-H bonds of approximately 109 . In contrast, sp² hybridization in trigonal planar structures, such as ethene (C₂H₄), has 33% s-character and C-H bonds of about 108 , while sp hybridization in linear arrangements, like ethyne (C₂H₂), with 50% s-character yields the shortest C-H bonds at 106 . Bond angle distortions arising from or constrained geometries can further modify bond lengths by altering orbital overlap. In crowded molecules, repulsive interactions between non-bonded atoms may lengthen bonds to relieve , while acute angles can compress them. For instance, in , the enforced 60° C-C-C bond angles deviate significantly from the ideal 109.5° tetrahedral angle, causing the C-C bonds to shorten to about 151 pm compared to 154 pm in due to increased p-character and bent-bond character. The Valence Shell Electron Pair Repulsion () theory integrates these effects by predicting molecular shapes based on minimizing repulsions between electron pairs, which indirectly governs bond lengths through the resulting and strain. Lone pairs or multiple bonds can distort angles, leading to variations in bond distances; for example, in trigonal bipyramidal molecules, equatorial bonds are often shorter than axial ones due to differing repulsion patterns. In conjugated pi systems, facilitates electron delocalization across overlapping p-orbitals, slightly shortening single bonds by imparting partial double-bond character through , as seen in where C-C single bonds are marginally shorter than in isolated alkanes. Bond lengths in covalent compounds exhibit systematic variations across periods of the periodic table, primarily due to the increasing experienced by valence electrons as increases from left to right. This enhanced nuclear attraction pulls the electrons closer to the nucleus, reducing atomic radii and thus shortening bond lengths between similar atoms or with a common partner. For instance, the C-H bond length is approximately 109 pm, the N-H bond is 101 pm, and the O-H bond is 96 pm, illustrating the contraction across the second period. In contrast, bond lengths generally increase down a group in the periodic table as size expands with additional electron shells, leading to greater internuclear distances despite similar characteristics. This trend is evident in the halogen-carbon bonds, where the C-Cl bond measures 177 pm, the Si-Cl bond 202 pm, and the Ge-Cl bond 211 pm, reflecting the progressive increase in the size of the group 14 elements. differences between bonded atoms can modulate these lengths but primarily follow the size pattern in . Homonuclear diatomic molecules of second-period elements further highlight these periodic patterns, with bond lengths influenced by and atomic size. The N≡N in N₂ is notably short at 110 pm, while the O=O in O₂ is 121 pm, and the F-F in F₂ extends to 142 pm, longer than expected for its position due to inter-lone-pair repulsions between the highly electronegative atoms. These variations underscore how increasing nuclear charge shortens bonds across the period, modulated by electron repulsion in later elements. In , where electrons are delocalized across a of metal atoms, bond lengths are typically longer and more variable than in covalent bonds, reflecting the non-directional nature of the interaction and dependence on radii. For example, nearest-neighbor distances in metals increase down groups, such as from Cu-Cu at 256 pm in to Ag-Ag at 289 pm in silver, due to expanding atomic sizes in group 11. This delocalized character results in effective "" lengths that are averages over the rather than fixed distances. Anomalies in these trends occur, notably in the F-F bond, which is weaker than the Cl-Cl bond despite fluorine's smaller , primarily because of strong repulsions between the lone pairs on the compact fluorine atoms that destabilize the bond. Such exceptions arise from electron density overcrowding in small, highly electronegative atoms, deviating from the general left-to-right shortening.

Bond Lengths in Organic Compounds

In compounds, carbon-carbon (C-C) bond lengths vary significantly depending on the hybridization of the carbon atoms involved. For sp³-hybridized carbons, as in alkanes like , the typical C-C single bond length is 154 pm. In contrast, sp²-hybridized carbons in alkenes feature C=C double bonds measuring approximately 134 pm, while sp-hybridized carbons in alkynes exhibit C≡C triple bonds around 120 pm. These differences arise from the increasing s-character in the hybrid orbitals, which brings the nuclei closer together. Strain in small rings further modifies these lengths; for instance, the C-C bonds in cyclobutane are elongated to about 156 pm due to angle compression. Carbon-hydrogen (C-H) bonds in molecules typically range from 106 to 110 pm, with variations influenced by the hybridization of the adjacent carbon and nearby substituents. In sp³-hybridized systems like or alkanes, C-H bonds average 109 pm, shortening slightly to 108 pm in sp²-hybridized alkenes and to 106 pm in sp-hybridized alkynes such as , reflecting the higher s-character. Adjacent multiple bonds can also reduce C-H lengths marginally, as seen in terminal alkynes where the bond is notably acidic and compact. Bonds between carbon and heteroatoms exhibit characteristic lengths that depend on bond order and electronegativity differences. Single C-O bonds, as in alcohols or ethers, measure 143 pm, while C=N double bonds in imines are shorter at 128 pm. Carbonyl groups (C=O) in aldehydes, ketones, and carboxylic acids feature double bonds around 120 pm, contributing to their reactivity. The following table summarizes representative bond lengths for common functional groups in organic compounds:
Functional GroupBond TypeLength (pm)
C-C (sp³)154
C=C (sp²)134
C≡C (sp)120
Alcohol/EtherC-O143
C=N128
CarbonylC=O120
Terminal Alkyne≡C-H106
Substitution effects, such as , can subtly alter bond lengths in saturated and unsaturated systems. In alkenes like propene, involving σ C-H bonds and the π* orbital of the slightly lengthens the adjacent C-C compared to unbranched alkanes. In aromatic systems, electron delocalization leads to intermediate bond lengths. exemplifies this with all C-C bonds at 139 pm, longer than a typical C=C but shorter than a C-C , distinguishing it from alternant structures like localized polyenes.

Bond Lengths in Inorganic and Coordination Compounds

In main group binary compounds, bond lengths exhibit systematic trends influenced by atomic size and differences. For instance, the P-O length is typically 151 pm in phosphates and related oxides, reflecting the covalent character in these structures. Similarly, the S-S measures 206 pm in elemental sulfur and disulfides, with lengths increasing down group 16 due to larger atomic radii, as seen in where Se-Se bonds extend to around 232 pm. In oxides, main group metal-oxygen bonds generally shorten with higher electronegativity of the central atom, such as Si-O at 161 pm versus Pb-O at over 220 pm, while halides show analogous lengthening down a group, with C-F at 135 pm contrasting Si-F at 156 pm, attributable to expanding orbital sizes. Ionic compounds lack discrete covalent bonds but feature interatomic distances modeled as the sum of ionic radii, providing effective "bond lengths" in crystal . In NaCl, for example, the Na-Cl distance is 281 pm, corresponding to the sum of Na⁺ (102 pm) and Cl⁻ (181 pm) radii in the rock salt structure. This approach extends to other halides, where distances increase with cation or anion size, such as 314 pm for K-Cl, influencing stability and properties. Coordination compounds display metal-ligand bond lengths that vary significantly with the metal's , as higher states contract the and enhance electrostatic attraction. For -O bonds, lengths are shorter in Fe(III) complexes (around 1.95 ) compared to Fe() (around 2.10 ), as observed in octahedral environments like ferrites and analogs. The trans influence further modulates these lengths, where a strong σ-donor lengthens the opposing metal-ligand bond by 0.1-0.2 , exemplified by Pt() complexes where trans to NH₃ extends to 2.30 versus 2.27 for . Multiple bonds in inorganic compounds often deviate from free molecular values due to coordination effects. The N≡N in free N₂ is 110 pm, but in metal nitrides and dinitrogen complexes, it elongates to 120-128 pm as shifts toward metal-nitrogen interactions, reducing . This is evident in V(III) dinitrogen where N-N stretches to 1.25 . Cluster compounds like and phosphazenes feature variable bond lengths arising from bridging interactions and delocalized bonding. In such as B₂H₆, terminal B-H bonds are 119 pm, while bridging B-H-B bonds extend to 133 pm, with B-B distances around 177 pm reflecting three-center two-electron bonds. Phosphazenes exhibit P-N bonds alternating between 156 pm (double-bond-like) and 174 pm (single), as in cyclic (NPCl₂)₃, due to partial π-delocalization in the . These variations underscore the multicenter bonding in such clusters.

References

  1. [1]
    Bond Lengths and Energies - Chemistry LibreTexts
    Jan 29, 2023 · Bond Lengths and Energies. Last updated: Jan 29, 2023. Save as PDF. Page ... Bond Lengths and Energies is shared under a CC BY-NC-SA 4.0 ...
  2. [2]
    1.3 Chemical Bonding — Introduction to Geochemistry - Vijay Vulava
    Bond length measured in pm ( = 10 − 12 m ) or Å Å ( = 10 − 10 m ), is defined as the distance between nuclei of two covalently bonded atoms in a molecule.
  3. [3]
    7.1 Molecular Structure and VSEPR Theory - UCF Pressbooks
    A bond distance (or bond length) is the distance between the nuclei of two bonded atoms along the straight line joining the nuclei. Bond distances are measured ...
  4. [4]
    [PDF] Bond Lengths, Stretching Force Constants, and Bond Dissociation ...
    Jan 16, 2009 · It is the energy needed to separate two atoms A and B directly connected by a bond without changing their electronic features (degree of ...
  5. [5]
    bond length (BT07003) - IUPAC
    The distance between atomic centers involved in a chemical bond. The notion of bond length is defined differently in various experimental methods.Missing: authoritative | Show results with:authoritative
  6. [6]
    Chemical Bonding: The Journey from Miniature Hooks to Density ...
    This review article traces the development of our understanding of bonding from prehistory, through the debates in the 19th century CE bearing on valence, to ...
  7. [7]
    8.2: Ionic Bonding - Chemistry LibreTexts
    Apr 12, 2023 · Ionic bonds are formed when positively and negatively charged ions are held together by electrostatic forces.Learning Objectives · Note the Pattern · The electrical behavior of...
  8. [8]
    Metallic Bonding - Chemistry LibreTexts
    Jan 29, 2023 · Metallic bonds occur among metal atoms. Whereas ionic bonds join metals to non-metals, metallic bonding joins a bulk of metal atoms.
  9. [9]
    Covalent Bond Distance, Radius and van der Waals Radius
    Jan 29, 2023 · While van der Waals radius is used to define half of the distance between the closest approach of two non-bonded atoms of a given element.
  10. [10]
    Gilbert N. Lewis and the chemical bond: The electron pair and the ...
    Nov 15, 2006 · We describe the development of Lewis's ideas concerning the chemical bond and in particular the concept of the electron pair bond and the octet rule.
  11. [11]
    8.9: Covalent Bond Properties: Order, Length, and Energy
    May 7, 2022 · Please note that this particular table is in picometers (10-12), and that a common unit for bond length is the Angstrom Å (10-10), so for ...
  12. [12]
    Definition of Angstrom in Physics and Chemistry - ThoughtCo
    Jun 9, 2025 · An angstrom or ångström is a unit of length used to measure very small distances. One angstrom is equal to 10 −10 m (one ten-billionth of a meter or 0.1 ...
  13. [13]
    9.7: Multiple Covalent Bonds - Chemistry LibreTexts
    Mar 20, 2025 · A double covalent bond is a covalent bond formed by atoms that share two pairs of electrons. A triple covalent bond is a covalent bond formed by ...
  14. [14]
    Chempendix - Bond Lengths
    Average Bond Lengths in Picometers (pm). Adapted from: Olmsted & Williams, Chemistry, 5th Edition, 2011. Reference: CRC Handbook of Chemistry and Physics, 2007.
  15. [15]
    Angstrom - Wikipedia
    Yet the angstrom was never part of the SI system of units, and has been increasingly replaced by the nanometre (10−9 m) or picometre (10−12 m).
  16. [16]
    The MSDS HyperGlossary: Distance Unit Conversions
    The bond distances between atoms are generally in the range of 100 to 300 pm. Angstrom (Å, metric), An Angstrom = 1 x 10-10 meters. 1 Angstrom = 100 pm = 0.1 nm ...
  17. [17]
    Hydrogen atoms can be located accurately and precisely by x-ray ...
    May 27, 2016 · We show that, contrary to widespread expectation, hydrogen atoms can be located very accurately using x-ray diffraction, yielding bond lengths ...
  18. [18]
    Orbital Overlap and Chemical Bonding - Krapp - 2006 - Chemistry
    Dec 1, 2006 · Bond strength and length are determined by orbital interaction, Pauli repulsion, and electrostatic attraction, not just maximum orbital overlap.
  19. [19]
    Recent Developments in the Methods and Applications of the Bond ...
    Distortion theorem; Keeping the bond valence sum constant, the average bond length in a coordination sphere increases the more the individual bond lengths ...
  20. [20]
    The Electronic Structure and Bonding in Some Small Molecules - PMC
    The calculation of bond strengths and potential energy curves requires one to consider electron correlation, both dynamic and static, i.e., treatments that go ...
  21. [21]
    The accuracy of ab initio molecular geometries for systems ...
    Nov 9, 2005 · In a large basis-set study using cc-pVQZ sets, it was found that the mean (maximum) absolute bond-length errors are 2.6 (8.5), 0.5 (1.7), 0.7 ( ...
  22. [22]
    [PDF] The prediction of molecular equilibrium structures by the standard ...
    Moreover, the Hartree–Fock bond lengths are too short and the MP4 bond lengths are too long relative to experiment. Thus, improvements in the description of ...
  23. [23]
    General Performance of Density Functionals - ACS Publications
    Bond Lengths. The performance of the various density functionals is usually quite good in the prediction of minimum energy structures. The density functionals ...
  24. [24]
  25. [25]
    On the correlation of bond order and bond length - ScienceDirect.com
    We found that the formula Pauling proposed to define “chemist's bond order”, n=exp[−(R−R0)/b] (where R0 is the equilibrium bond length of a single bond) ...
  26. [26]
    Common Bond Energies (D - Wired Chemist
    Common Bond Energies (D. ) and Bond Lengths (r). Hydrogen. Bond, D (kJ/mol), r (pm). H-H, 432, 74. H-B, 389, 119. H-C, 411, 109. H-Si, 318, 148. H-Ge, 288, 153.
  27. [27]
    Covalent Radius of the elements - Photographic Periodic Table
    Covalent Radius of the elements ; Oxygen, 66 pm, Osmium ; Nitrogen, 71 pm, Silver ; Carbon, 76 pm, Thallium ; Boron, 85 pm, Ruthenium ; Beryllium, 96 pm, Lead ...
  28. [28]
    Electronegativity | Periodic Table of Elements - PubChem - NIH
    Tabular Electronegativity (Pauling Scale) Data ; 3, Li, Lithium · 0.98 ; 4, Be, Beryllium · 1.57.
  29. [29]
    Electronegativity effects and single covalent bond lengths of ...
    Mar 31, 2014 · The results also show that electronegativity is a major influence on covalent bond lengths and the set of electronegativity scale and covalent ...
  30. [30]
    Ionic Radii - WebAssign
    Li, 1.52, Be, 1.11, N, 0.75, O, 0.73, F, 0.71. Li1+, 0.60, Be2+, 0.31, N3–, 1.71, O2–, 1.40, F1–, 1.36. Na, 1.86, Mg, 1.60, Al, 1.43, S, 1.04, Cl, 0.99.
  31. [31]
    A quantitative definition of hypervalency - RSC Publishing
    Aug 14, 2015 · Although O is more electronegative than Cl, many of the oxides require structures with double bonds, which might lead to higher γ values. The ...
  32. [32]
    [PDF] Bond Length, Bond Strength & Acidity
    Effect of hybridization on length of single bonds: C–H and C–C bonds shorten slightly with increased s character on carbon sp3–sp3.
  33. [33]
    [PDF] The re Structure of Cyclopropane - SMU
    Most important, we note that the CC bond in cyclopropane is shorter than that in ethane by about 0.021 Å. 40 This is considerably less than the 0.04 Å often ...
  34. [34]
    Experimental data for C 2 H 6 (Ethane) - CCCBDB
    Calculated vibrational frequencies for C2H6 (Ethane). Gas-phase IR spectra can be found in the NIST Chemistry Webbook here. Rotational Constants (cm-1) ...Missing: ethene | Show results with:ethene
  35. [35]
    9.2: The VSEPR Model - Chemistry LibreTexts
    Jul 7, 2023 · The premise of the VSEPR theory is that electron pairs located in bonds and lone pairs repel each other and will therefore adopt the geometry ...
  36. [36]
  37. [37]
    [PDF] Average Bond Lengths - WebAssign
    Average Bond Lengths bond r bond (pm). C–H. 109. C–C. 154. C=C. 134. C≡C. 120. C–N. 147. C=N. 129. C≡N. 116. C–O. 143. C=O. 120. C≡O. 113. C–F. 135. C–Cl. 177.
  38. [38]
  39. [39]
    Table of bond lengths in chemical molecules - Calculla
    Substance, Molecular formula, Bond lengths [pm] ; ethanol, C2H5OH, C-C : 151.2 pm C-O : 143.1 pm O-H : 97.1 pm C-H : 110 pm ; methanol, CH3OH, C-H : 109.4 pm C-O ...
  40. [40]
    Oxygen - Wikipedia
    2, the major part of the Earth's atmospheric oxygen (see Occurrence). O2 has a bond length of 121 pm and a bond energy of 498 kJ/mol. Trioxygen (O 3) ...
  41. [41]
    molecular orbital theory - Diatomic halogens - bond strength
    Oct 9, 2014 · If we go for the Lewis or valence bond view, we'd argue that the large number of lone pairs on a small atom will be highly repulsive.Missing: length | Show results with:length
  42. [42]
    Bond-length distributions for ions bonded to oxygen - PubMed Central
    Bond-length distributions are examined for three configurations of the H+ ion, 16 configurations of the group 14–16 non-metal ions and seven configurations ...
  43. [43]
    Sulfur-sulfur bond lengths, or can a bond length be estimated from a ...
    Antiphase Boundaries in the Turbostratically Disordered Misfit Compound (BiSe)1+δNbSe2. Inorganic Chemistry 2015, 54 (21) , 10309-10315. https://doi.org ...Missing: reliable sources
  44. [44]
    mp-22862: NaCl (Cubic, Fm-3m, 225) - Materials Project
    All Na-Cl bond lengths are 2.79 Å. Cl¹⁻ is bonded to six equivalent Na¹⁺ atoms to form a mixture of corner and edge-sharing ClNa₆ octahedra. The corner ...
  45. [45]
    Experimental data for NaCl (Sodium Chloride) - CCCBDB
    Bond length · Rotational Constant · Moment of Inertia · Dipole and Quadrupole ... Calculated geometries for NaCl (Sodium Chloride). Bond descriptions.Missing: pm | Show results with:pm
  46. [46]
    Data mining of iron(II) and iron(III) bond-valence parameters, and ...
    Although both Fe atoms were assigned an oxidation state of 2 and have a coordination number of 6, Fe1II—N distances span 1.911–2.067 Å, while Fe2II—N distances ...
  47. [47]
    First proof of the 'surface <i>trans</i>-effect' - - Diamond Light Source
    Strongly trans-directing ligands weaken the ligand-metal bond that lies directly opposite, in the trans position. This causes the metal-ligand bond length in ...
  48. [48]
    Experimental data for N 2 (Nitrogen diatomic) - CCCBDB
    Bonds, angles · Bond, angle, or dihedral · DFT grid size on point group · DFT grid on bond length · Core correlation - bond length · Same bond/angle many ...Missing: pm | Show results with:pm
  49. [49]
    Dinitrogen Binding and Activation: Bonding Analyses of Stable V(III/I ...
    Aug 25, 2022 · For these complexes, the V–N bond length is between 1.704 and 1.767 Å, and the N–N bond length is in the range of 1.204–1.280 Å. The N2 unit in ...
  50. [50]
    Bond length based bond orders for borane cluster compounds
    This relationship is useful in deducing the distribution of electronic charge, and the approach can be extended to other homonuclear clusters. Previous article ...
  51. [51]
    Molecular structures of non-geminally substituted phosphazenes ...
    There are two significantly different P–N bond lengths in the ring, 1·586 and 1·555 Å. The P–Cl bonds (2·100 Å) are considerably longer than those in N4P4Cl8.