Fact-checked by Grok 2 weeks ago

Covalent radius

The covalent radius of an atom is a measure of its size when it participates in a , defined as half the internuclear distance between two identical atoms joined by a single . This value, typically expressed in picometers (pm), provides a standardized way to quantify atomic dimensions in molecular contexts, accounting for the that holds the atoms together. Introduced by Linus Pauling in his seminal 1939 work The Nature of the Chemical Bond, the concept of covalent radius emerged from empirical measurements of bond lengths using techniques such as X-ray crystallography, electron diffraction, and spectroscopy. Pauling assigned specific single-bond radii to elements—for instance, 77 pm for carbon and 66 pm for oxygen—allowing the prediction of internuclear distances by summing the radii of bonded atoms, with adjustments for factors like electronegativity differences and bond multiplicity. Modern refinements, such as those based on quantum mechanical calculations and updated experimental data, have refined these values while preserving the core principle. Covalent radii exhibit clear that reflect configurations and nuclear charge effects: they generally increase down a group in the periodic due to the addition of principal quantum levels, which expand the atomic size, and decrease across a from left to right owing to the increasing that pulls electrons closer to the . For example, the covalent radius of carbon is about 76 , while fluorine's is 57 in the same . These trends differ from metallic radii, where atoms are typically larger due to delocalized electrons, and ionic radii, which shrink for cations and expand for anions compared to covalent values. In practice, covalent radii are essential for estimating lengths in and inorganic molecules, predicting molecular geometries, and understanding reactivity patterns, such as steric hindrance in crowded structures. They also inform computational modeling in and , where accurate radius data helps simulate crystal structures and molecular interactions. Variations exist for different hybridization states (e.g., sp³ carbon at 77 pm versus sp² at 73 pm) and bond orders, with and bonds shortening distances by about 15–20% relative to bonds.

Definition and Principles

Definition

The covalent radius of an atom is defined as half the internuclear distance between two identical atoms when they are joined by a single . This measure provides a standardized way to quantify atomic size specifically within the context of covalent bonding, where electrons are shared between atoms. A key principle underlying the use of covalent radii is the approximate additivity of lengths, whereby the between two dissimilar atoms A and B in a single is estimated as the sum of their individual covalent radii, R(\ce{A-B}) \approx r(\ce{A}) + r(\ce{B}). This additivity holds reasonably well for many homopolar and heteropolar bonds, facilitating predictions of molecular geometries. Covalent radii are typically expressed in picometers (), the standard SI-derived unit for such lengths, though angstroms () were historically common, with the conversion $1 \, \AA = 100 \, \mathrm{pm}. Values may be empirical, derived from experimental measurements in crystals or molecules, or calculated using quantum mechanical methods for consistency across the periodic table. For example, the H–H bond length in the dihydrogen molecule is 74 pm, so the covalent radius of hydrogen from this homonuclear bond is 37 pm.

Relation to Bond Lengths

The covalent radius of an atom is fundamentally defined as half the internuclear distance in a single covalent bond between two identical atoms, leading to the additivity rule for estimating bond lengths in heteronuclear diatomic molecules. For a single bond between atoms A and B, the bond length R(\ce{A-B}) is approximated by the sum of their individual covalent radii: R(\ce{A-B}) \approx r_\ce{A} + r_\ce{B}, where r_\ce{A} = \frac{1}{2} R(\ce{A-A}) and r_\ce{B} = \frac{1}{2} R(\ce{B-B}). This assumption stems from the idea that each atom contributes a fixed "share" to the bond length, analogous to the homonuclear case, and was first systematically applied by Linus Pauling in his analysis of molecular structures. While the additivity rule provides a reliable first-order approximation, it exhibits small deviations typically on the order of 1-5% (or about 2-3 pm for bonds around 100-150 pm) due to factors such as differences in orbital overlap and bond polarity arising from electronegativity variations. Modern compilations of covalent radii achieve a standard deviation of 2.8 pm across hundreds of bond lengths when fitting the additivity model, underscoring its practical accuracy despite these limitations. For hydrogen, the homonuclear radius (37 pm) differs from the effective value (~31 pm) used in heteronuclear bonds to better fit experimental data. In practice, covalent radii enable the prediction of unknown bond lengths by combining established values for the constituent atoms, facilitating rapid estimates in molecular design and structural chemistry without direct measurement. For instance, the C-N length can be estimated as the sum of the carbon (76 pm) and (71 pm) covalent radii, yielding approximately 147 pm, which aligns closely with experimental values. A example is the C-H bond: using the effective covalent radius of (31 pm) and carbon (76 pm), the predicted is 107 pm, compared to the experimental value of 109 pm in , demonstrating the rule's utility with minimal error.

Historical Development

Early Concepts

Linus Pauling advanced these ideas significantly in the 1930s, culminating in his 1939 book The Nature of the Chemical Bond and the Structure of Molecules and Crystals, where he formalized covalent radii by analyzing bond lengths obtained primarily from X-ray crystallography of crystals and molecules. Pauling derived radii as half the internuclear distance in homonuclear single bonds, assuming additivity and constancy for similar bonding environments across compounds, which enabled the estimation of heteronuclear bond lengths by summing constituent atomic radii. In his initial tables, Pauling provided values such as 77 for carbon in tetrahedral single bonds (based on the 154 pm C–C distance in ), 70 for , and 66 for oxygen, reflecting empirical data from diatomic gases, molecules, and while accounting for minor adjustments due to differences. These radii demonstrated transferability, as seen in predictions matching observed bond lengths in hydrocarbons and other . Pauling's framework established covalent radius as an indispensable practical tool for structural chemistry, facilitating rapid bond length calculations and molecular modeling long before comprehensive quantum mechanical derivations of atomic sizes became routine.

Modern Refinements

In the late 20th and early 21st centuries, refinements to covalent radii increasingly relied on statistical analyses of vast crystallographic datasets, enabling more precise averages and uncertainties. A landmark study by Cordero et al. in 2008 analyzed over 200,000 covalent bonds from the Cambridge Structural Database, deriving updated radii for elements up to atomic number 96 by averaging bond lengths and incorporating standard deviations to reflect variability. This approach yielded, for example, a refined covalent radius for sp³-hybridized carbon of 76(1) pm, where the uncertainty in parentheses denotes one standard deviation, providing a more robust empirical basis than earlier smaller-scale compilations. Building on electronegativity principles, R.T. Sanderson's 1983 model introduced dynamic adjustments to covalent radii based on electron density redistribution during bond formation, treating radii as variable quantities that equilibrate according to atomic electronegativities. In this framework, the effective radius of an atom in a scales inversely with the electronegativity difference between bonded atoms, allowing predictions of bond lengths in polar compounds without fixed values. This electronegativity equalization concept, detailed in Sanderson's work on polar covalence, has influenced subsequent models by emphasizing the context-dependent nature of atomic sizes in molecules. Further modern refinements account for hybridization effects through , distinguishing radii based on orbital overlap and geometry. For carbon, these yield specific values of 76 pm for sp³ hybridization (as in tetrahedral structures), 73 pm for sp² (as in alkenes), and 69 pm for (as in alkynes), reflecting shorter bonds due to increased s-character in the hybrid orbitals. These hybridization-dependent radii, integrated into updated tables like those from Cordero et al., enhance accuracy in predicting geometries for organic and organometallic compounds. For heavy and superheavy elements, relativistic effects have necessitated adjustments to covalent radii, as inner electrons approach speeds nearing that of light, contracting s-orbitals and expanding p/d/f orbitals. Pyykkö's 2012 review highlighted these impacts, extending covalent radius estimates to elements up to Z=118 () by incorporating Dirac-Fock calculations that account for spin-orbit coupling and mass-velocity corrections, resulting in larger radii for atoms compared to non-relativistic predictions. This work underscores how stabilizes unexpected oxidation states and bond lengths in transactinide . More recent theoretical developments, as of 2025, include first-principles derivations of covalent radii using quantum chemical calculations, providing values independent of experimental data for elements like H through . Additionally, radii based on the expectation value ⟨r⁴⟩ offer a new quantum mechanical perspective on size trends.

Methods of Determination

Experimental Techniques

and serve as primary experimental techniques for determining lengths in solid-state compounds, where internuclear distances are measured through the diffraction patterns produced by crystalline lattices. In diffraction, X-rays scatter off the electron clouds surrounding nuclei, allowing precise mapping of positions in single crystals, with resolutions often reaching 0.8 or better for covalent structures. diffraction complements this by scattering from nuclei, providing superior accuracy for light elements like and enabling direct measurement of all positions, including those in s, without the bias toward heavier atoms seen in methods. These techniques yield internuclear distances that, for homonuclear diatomic bonds (A-A), define the covalent radius r_{\text{cov}}(A) as half the averaged , accounting for multiple observations to mitigate structural variations. For gaseous molecules, in the or range determines lengths by analyzing transitions between quantized levels, which depend on the molecule's . The rotational constant B, derived from spacings, relates to the r via B = \frac{h}{8\pi^2 c \mu r^2}, where \mu is the , allowing inversion to obtain r. A classic example is the (H_2), where yields a of 74.14 pm, establishing the covalent radius of as approximately 37 pm. Gas-phase electron diffraction provides another key method for volatile compounds, scattering electrons off molecular electron densities to reconstruct internuclear distances without requiring crystallinity. This technique achieves high precision, typically on the order of 0.004 (0.4 ) for , by analyzing intensities as a function of and applying least-squares refinement to dynamic molecular models. To derive standardized covalent radii, experimental data are aggregated from vast repositories like the Structural Database (), which compiles over 1.36 million curated crystal structures from and studies of organic and metal-organic compounds. Analysis involves averaging internuclear distances across similar bonds, with statistical weighting to handle errors from thermal motion, , and environmental effects, ensuring robust values for and applications. Such databases facilitate error handling through outlier rejection and variance estimation, often cross-validating with gas-phase measurements for consistency.

Computational Methods

Ab initio methods provide a foundational approach for computing covalent radii by solving the to optimize molecular geometries and determine equilibrium bond lengths. In the Hartree-Fock (HF) method, the wave function is approximated as a single , enabling the calculation of densities and bond distances without empirical parameters beyond the basis set. (DFT), which incorporates exchange-correlation effects more efficiently, has become prevalent for such optimizations due to its balance of accuracy and computational cost. These methods typically derive covalent radii by halving the computed homonuclear single- length, such as in diatomic molecules, or by fitting to a series of homologous compounds. A representative application of DFT involves the B3LYP , which combines Hartree-Fock with Becke's gradient-corrected and Lee-Yang-Parr . For (C₂H₆), B3LYP/6-311+G(3df,2p) calculations yield a C-C single-bond of 1.531 , corresponding to a carbon covalent radius of 76.6 pm when halved. This value aligns closely with empirical estimates and demonstrates DFT's utility in predicting bond lengths for light elements. For heavier elements, relativistic effects are incorporated via scalar-relativistic pseudopotentials or Dirac-Hartree-Fock approaches to account for orbital contraction. Molecular dynamics (MD) simulations extend these static optimizations by incorporating dynamic effects, particularly thermal vibrations, to compute time-averaged bond lengths that better reflect experimental conditions at finite temperatures. In ab initio MD or density functional theory-based MD, nuclear trajectories are propagated using forces derived from on-the-fly electronic structure calculations, allowing the extraction of root-mean-square bond fluctuations and effective radii. For instance, vibrational averaging in simple hydrocarbons like methane shows bond length variations of ~0.01 Å, which adjust the nominal covalent radius by a few picometers depending on temperature. These simulations are essential for systems where zero-point motion or anharmonic effects significantly influence observed bond dimensions. Semi-empirical models, such as the extended Hückel theory (EHT), offer faster approximations for estimating covalent radii in large molecules or solids by parameterizing overlap and elements based on ionization potentials and electronegativities. EHT computes molecular orbitals and geometries iteratively, providing lengths that can be used to derive radii with errors typically under 5% for systems. Refinements to EHT, including adjustable Wolfsberg-Helmholtz parameters, enhance its ability to reproduce in covalent radii across the main-group elements, making it suitable for screening before more rigorous treatments. Validation of these computational approaches relies on benchmarking against experimental bond lengths from techniques like or gas-phase . For superheavy elements inaccessible to experiment, relativistic DFT provides ; Pyykkö's calculations using the PBE functional and small-core relativistic pseudopotentials yield a single-bond covalent radius for (element 118) of 157 pm, highlighting the relativistic expansion compared to lighter like (131 pm). Such comparisons confirm that DFT radii reproduce experimental values with standard deviations of ~3 pm for elements up to Z=86.

Standard Covalent Radii

Values for Single Bonds

The standard covalent radii for single bonds are derived from extensive crystallographic data in the , where bond lengths are averaged assuming additivity such that the distance between atoms A and B equals the sum of their individual covalent radii. These values, established by Cordero et al. in 2008, cover elements from to (atomic number ) and serve as a benchmark for predicting single-bond lengths in molecular structures. The dataset relies on over 100,000 bond distances for common elements like carbon and oxygen, with typical standard deviations of approximately 6 pm across the set, indicating high consistency in the experimental measurements. For main-group elements, the radii reflect typical sp³ hybridization in saturated compounds. Transition metal radii, however, depend on coordination number and spin state; the tabulated values use a default coordination number of 4 for consistency, though adjustments may be needed for other geometries. No major updates to this dataset have superseded it for single-bond applications, though complementary theoretical sets exist for superheavy elements. The following table presents representative covalent radii for single bonds in picometers (pm) for main-group elements, drawn from the Cordero compilation. These derive from half the A–A homonuclear or averaged A–X heteronuclear distances to electronegative partners like F, O, or N.
ElementSymbolRadius (pm)
H31
He28
Li128
Be96
B84
CarbonC (sp³)76
N71
OxygenO66
F57
Ne58
SodiumNa166
MagnesiumMg141
AluminumAl121
Si111
P107
S105
Cl102
Ar106
K203
CalciumCa176
Ga122
Ge120
As119
Se120
BromineBr120
Kr116
Rb220
Sr195
In142
TinSn139
Sb139
Te138
IodineI139
Xe140
CesiumCs244
Ba215
Tl145
LeadPb146
Bi148
Po140
At150
Rn150
For transition metals, examples include iron (low spin: 132 pm; high spin: 152 pm for CN=4) and (132 pm for CN=4), highlighting sensitivity to electronic configuration. As an illustration, silicon's covalent radius of 111 pm exceeds carbon's 76 pm, consistent with increasing atomic size down group 14. Noble gas values, such as argon's 106 pm, are interpolated from limited data due to their rarity in covalent bonding. The covalent radius of exhibits a systematic decrease across each of the periodic table from left to right, primarily due to the increasing experienced by valence electrons as protons are added to the without a corresponding increase in shielding from inner electrons. This trend results in a contraction of approximately 20-30 pm per for main-group . For instance, in 2, the covalent radius diminishes from 76 pm for carbon to 57 pm for . Similar patterns are observed in other , where the enhanced nuclear attraction pulls the electron cloud closer, reducing the atomic size. In contrast, covalent radii increase down a group as additional shells are occupied, extending the valence electrons farther from the despite the increasing charge. This expansion arises from the radial distribution of higher orbitals. An illustrative example is group 14, where the covalent radius grows from 76 pm for carbon, to 111 pm for , and 120 pm for . The increment per is typically larger in the p-block than in the s-block, reflecting differences in orbital and shielding efficiency. Notable anomalies disrupt these general trends. The causes a gradual decrease in covalent radii across the series (elements 57-71), stabilizing around 160-190 pm for late lanthanides like (198 pm) and (187 pm), due to poor shielding by electrons, which leads to a stronger without proportional size increase. In superheavy elements (Z > 100), relativistic effects further contribute to a slight contraction; the high nuclear charge accelerates inner electrons to near-relativistic speeds, stabilizing s-orbitals and indirectly compressing orbitals, as incorporated in theoretical compilations such as Pyykkö (2008) for elements up to 118. These periodic variations are often visualized in plots of covalent radius versus , revealing smooth declines across periods interrupted by group ascents and subtle inflections at transition series or f-block regions; for example, period 2 shows a steep ~19 pm drop from C to F, while group 14 illustrates a ~35 pm rise from C to Si. Such graphical representations underscore the interplay of charge, shielding, and quantum effects in dictating dimensions.

Variations in Covalent Radii

Multiple Bonds

In covalent bonds with higher bond orders, such as and bonds, the effective covalent radii of the atoms involved are smaller than those for bonds, reflecting the shorter interatomic distances observed experimentally. This shortening arises primarily from the additional pi-bonding in multiple bonds, which increases the between the nuclei and enhances the attractive forces, pulling the atoms closer together; the associated increase in s-character of the orbitals further contributes to this contraction by concentrating nearer to the nuclei./21%3A_Resonance_and_Molecular_Orbital_Methods/21.09%3A_Bond_Lengths_and_Double-Bond_Character) Typically, bonds result in covalent radii that are about 10-15% smaller than -bond values, while bonds are approximately 20-25% smaller, though these factors vary slightly by . A representative example is the carbon-carbon bond in ethane (C₂H₆), where the single bond length is 154 pm, corresponding to a covalent radius of about 77 pm per carbon atom, compared to ethene (C₂H₄), where the double bond length is 134 pm, yielding a radius of 67 pm per carbon./01%3A_Structure_and_Bonding/1.13%3A_Ethane_Ethylene_and_Acetylene) Similarly, in ethyne (C₂H₂), the triple bond length of 120 pm gives a carbon radius of 60 pm./01%3A_Structure_and_Bonding/1.13%3A_Ethane_Ethylene_and_Acetylene) These differences can be approximated by adjustment factors, such as R(\text{double}) \approx 0.85 \times R(\text{single}) for many elements, derived from empirical fits to bond length data. Bond-order-specific covalent radii have been systematically determined through self-consistent fits to extensive experimental bond length data from spectroscopy (e.g., X-ray crystallography and electron diffraction) and high-level computational methods (e.g., Dirac-Coulomb relativistic calculations). The following table presents such radii (in pm) for selected common elements, based on these analyses:
ElementSingle BondDouble BondTriple Bond
C756760
N716054
O635753
F645953
Si116111106
P11110294
S1039493
Cl999595

Effects of Hybridization and Electronegativity

The covalent radius of an atom varies with its hybridization state due to differences in the s-character of the hybrid orbitals. In sp³-hybridized atoms, such as carbon in alkanes or , the hybrid orbitals contain 25% s-character, resulting in a larger effective radius of approximately 76 pm for carbon. This larger size arises because the lower s-character leads to less contraction of the orbitals toward the . In contrast, sp²-hybridized carbon, as found in alkenes or , has 33% s-character, yielding a smaller radius of about 71 pm, while sp-hybridized carbon in alkynes exhibits 50% s-character and the smallest radius of roughly 69 pm. The trend of decreasing radius with increasing s-character (sp³ > sp² > sp) occurs because higher s-character concentrates closer to the , effectively contracting the orbital size and shortening bond lengths. Electronegativity differences in polar introduce asymmetry in effective atomic radii, as shifts toward the more atom, making its effective radius smaller and the less electronegative atom's larger than in homonuclear bonds. For example, in the , 's high (4.0 on the Pauling scale) compared to (2.2) pulls electrons closer, resulting in an effective fluorine radius smaller than the 72 pm observed in F-F. This asymmetry is quantified by the Schomaker-Stevenson rule, which predicts shortening Δd ≈ 6 pm per unit difference in electronegativity (ΔEN), so for high ΔEN values around 1.8 (as in H-F), the total shortening is about 10-12 pm, with the more electronegative atom's radius decreasing accordingly. Beyond hybridization and , other environmental factors like in molecular complexes can alter effective covalent radii. Higher s increase bond lengths by 5-10 pm due to greater ligand-ligand repulsion, effectively enlarging the central atom's ; for instance, mean M-O bond lengths in oxyanions vary by 6-9 pm with changes from 3 to 6. in further modulate radii through , where polar solvents stabilize charged or polar bonds, potentially lengthening them by 1-5 pm compared to gas phase, though these variations are typically small and context-dependent. A representative example is carbon, where the sp³-hybridized form in has a covalent radius of 76 pm (C-C 154 pm), while the sp²-hybridized form in shows 71 pm (C-C 142 pm), illustrating how hybridization influences radius without bond multiplicity effects.

Comparisons with Other Atomic Radii

Versus

The covalent radius of an atom is defined as approximately half the internuclear distance between two covalently bonded atoms of the same element in a or within a larger covalent compound, reflecting the size of the atom when involved in strong, shared-electron bonding. In contrast, the represents half the distance of closest approach between the nuclei of two non-bonded atoms of the same element, typically observed in intermolecular contacts within molecular crystals or van der Waals complexes, where weaker dispersion forces dominate. This distinction arises because covalent bonding pulls atoms closer together through orbital overlap, whereas van der Waals interactions maintain larger separations to avoid repulsion in non-bonded scenarios. Quantitatively, the van der Waals radius is generally 1.5 to 2.2 times larger than the , depending on the element; for carbon, the covalent radius is 76 pm, while the is 170 pm, yielding a of about 2.2. Covalent radii are measured primarily from bond lengths determined by techniques such as of covalent compounds or gas-phase spectroscopy, whereas van der Waals radii are derived from intermolecular distances in crystal structures of molecular solids or from second virial coefficients in gas-phase equations of state. These differences have key implications for molecular behavior: the covalent radius informs the and reactivity at bonding sites within molecules, such as the C–H of 109 pm in (CH₄), while the van der Waals radius governs packing efficiency and stability in condensed phases, exemplified by the intermolecular center-to-center distance of approximately 380 pm between molecules in its solid phase. Thus, covalent radii are crucial for predicting intramolecular structures and reaction pathways, whereas van der Waals radii are essential for understanding intermolecular forces, , and energies.

Versus Ionic Radius

The covalent radius describes the size of a neutral atom involved in a covalent bond, where electrons are shared between atoms, whereas the ionic radius pertains to the size of ions in ionic compounds, characterized by complete electron transfer from a cation to an anion, leading to electrostatic attraction in a crystal lattice. This fundamental distinction in bonding type results in significant differences in measured sizes; for example, the covalent radius of sodium is 166 pm, reflecting its neutral atomic size in metallic or covalent contexts, while the ionic radius of Na⁺ is 102 pm for sixfold coordination, as the loss of an electron increases the effective nuclear charge and contracts the ion. A key consequence of ionization is the size inversion observed across the periodic table: cations are invariably smaller than their parent neutral atoms due to higher pulling electrons closer, while anions are larger owing to added electrons increasing electron-electron repulsion without a corresponding increase in nuclear charge. For , the covalent radius is 99 pm in neutral Cl₂ molecules, but the Cl⁻ anion has an of 181 pm in sixfold coordination, expanding the electron cloud in ionic lattices like NaCl. These ionic radii were initially modeled by , who assumed additivity of cation and anion radii to match observed interionic distances in crystals such as NaCl (with a Na⁺-Cl⁻ distance of 281 pm), deriving values based on and differences. Subsequent refinements by and Prewitt accounted for and structural effects in oxides and halides, providing effective ionic radii that vary systematically—for instance, Na⁺ radii increase from 102 pm at 6 to 116 pm at 8—enhancing accuracy for predicting parameters in ionic solids. In contrast to purely covalent bonds, polar covalent bonds represent transitional cases where partial ionic character, driven by differences, blends the radii; in HCl, for example, the of 127 pm is shorter than the sum of pure covalent radii ( at ~31 pm and at 99 pm) due to ~17% ionic contribution, adjusting effective sizes accordingly.

References

  1. [1]
    Covalent Radius - an overview | ScienceDirect Topics
    Covalent radii refer to the characteristic values associated with each type of atom that indicate their size when forming covalent bonds, which are used to ...
  2. [2]
    2.3.3: Covalent and Ionic Radii - Chemistry LibreTexts
    Jan 31, 2024 · The page discusses various methods for measuring the radii of atoms and ions, including nonpolar atomic radii, van der Waals radii, ...
  3. [3]
    [PDF] THE NATURE OF THE CHEMICAL BOND MMI ij m
    Mar 1, 2025 · A detailed discussion of the valence-bond theory of the electronic structure of metals and intermetallic compounds is also presented.
  4. [4]
    Periodic Trends - EdTech Books
    Covalent radius increases as we move down a group because the n level (orbital size) increases. Covalent radius mostly decreases as we move left to right across ...
  5. [5]
    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 ...Missing: IUPAC | Show results with:IUPAC
  6. [6]
    Size of Atoms
    The covalent radius for an element is usually a little smaller than the metallic radius. This can be explained by noting that covalent bonds tend to squeeze the ...
  7. [7]
  8. [8]
    WebElements Periodic Table » Hydrogen » radii of atoms and ions
    The bond length in HH is: 74.130pm. There are several other ways ways to define radius for atoms and ions. Follow the appropriate hyperlinks for literature ...
  9. [9]
  10. [10]
    None
    ### Extracted Information on van Arkel and Ketelaar's Work in the 1920s on Bonding Types, Relation to Covalent Bonding and Radii
  11. [11]
    Covalent radii revisited - Dalton Transactions (RSC Publishing)
    A new set of covalent atomic radii has been deduced from crystallographic data for most of the elements with atomic numbers up to 96.
  12. [12]
    (PDF) Relativistic Effects in Chemistry: More Common Than You ...
    Jan 30, 2012 · ArticlePDF Available. Relativistic Effects in Chemistry: More Common Than You Thought. January 2012; Annual Review of Physical Chemistry 63(1): ...
  13. [13]
    Single-crystal X-ray Diffraction - SERC (Carleton)
    Jun 15, 2018 · Single-crystal X-ray Diffraction is a non-destructive analytical technique which provides detailed information about the internal lattice of crystalline ...
  14. [14]
    Evaluation of models determined by neutron diffraction and ... - NIH
    The advantage of neutron diffraction experiments is that the positions of all atoms can be determined, including H atoms, which are hardly detectable by X-ray ...Missing: solids | Show results with:solids
  15. [15]
    Single-crystal x-ray diffraction structures of covalent ... - Science
    Jul 6, 2018 · The high quality of the crystals allowed collection of single-crystal x-ray diffraction data of up to 0.83-angstrom resolution, leading to ...
  16. [16]
    Single-Crystal Neutron Diffraction Study on Guanidine, CN3H5
    Feb 14, 2013 · Our data show that such-optimized C-H, N-H, O-H, and B-H bond lengths coincide well with results from neutron diffraction-the exptl.
  17. [17]
    Experimental data for H 2 (Hydrogen diatomic) - CCCBDB
    Computational Chemistry Comparison and Benchmark DataBase Release 22 (May 2022) Standard Reference Database 101 National Institute of Standards and ...
  18. [18]
    The Cambridge Structural Database - IUCr Journals
    Mar 8, 2016 · The Cambridge Structural Database (CSD) contains a complete record of all published organic and metal–organic small-molecule crystal structures ...
  19. [19]
    Applications of the Cambridge Structural Database in chemical ...
    The CSD records data from single-crystal X-ray and neutron diffraction studies of C-containing compounds: organics and metal–organics. Powder structures are ...
  20. [20]
    CCCBDB bond length model
    Calculated at B3LYP/6-311+G(3df,2p). Species, Name, Bond type, Bond Length ... Ethane, 1,2-dichloro-, rCC, 1.531, 1.512, -0.019, 1, 2. C6H8, (Z)-hexa-1,3,5-triene ...
  21. [21]
    mean bond lengths from radial distribution functions - AIP Publishing
    Jan 13, 2017 · This paper is concerned with the questions of how accurate conventional estimates of the average bond length are and how they can be improved.
  22. [22]
    Extended Hückel theory for band structure, chemistry, and transport ...
    Aug 29, 2006 · We describe a semiempirical atomic basis extended Hückel theoretical (EHT) technique that can be used to calculate bulk band structure, ...Missing: estimation covalent
  23. [23]
  24. [24]
    Molecular Single‐Bond Covalent Radii for Elements 1–118
    Dec 18, 2008 · A self-consistent system of additive covalent radii, R(AB)=r(A) + r(B), is set up for the entire periodic table, Groups 1–18, Z=1–118.
  25. [25]
    Molecular Double‐Bond Covalent Radii for Elements Li–E112
    Nov 18, 2009 · The aim of additive covalent radii is to express approximately a bond length as the sum of two atomic radii.<|control11|><|separator|>
  26. [26]
    Hybridization Trends for Main Group Elements and Expanding the ...
    Apr 28, 2014 · according to sp, sp2, and sp3 bonding permits prediction of bond lengths, bond angles, relative electronegativity, and basicity of valence.
  27. [27]
    Some Revisions of the Covalent Radii and the Additivity Rule for the ...
    ... Covalent Radii and the Additivity Rule for the Lengths of Partially Ionic Single Covalent Bonds ... An equation to calculate internuclear distances of covalent ...Missing: deviations | Show results with:deviations
  28. [28]
    Solute–Solvent Interactions in Modern Physical Organic Chemistry
    Nov 11, 2020 · First, the effects of solvents on covalent chemistry are discussed; more specifically the impact of these effects on reactivity and the ...
  29. [29]
    Atomic Radius | Periodic Table of Elements - PubChem - NIH
    The atomic radius of a chemical element is a measure of the size of its atom, usually, the distance from the center of the nucleus to the outermost isolated ...
  30. [30]
    WebElements Periodic Table » Carbon » radii of atoms and ions
    Pauling ionic radii. This table shows Pauling radii for carbon. Ion, Pauling radius / pm, Periodicity link. C(I), 29, Pauling radius for +1 ions of the chemical ...
  31. [31]
    Direct observation of van der Waals two-dimensional crystal and ...
    Sep 17, 2018 · The intermolecular distance of the methane clusters is about 3.6 Å which is smaller than that of the two-dimensional methane layer on the ...
  32. [32]
    WebElements Periodic Table » Sodium » radii of atoms and ions
    Covalent radius (2008 values), 166, 2008 values of covalent radius ... Pauling ionic radii. This table shows Pauling radii for sodium. Ion, Pauling radius / pm ...
  33. [33]
    WebElements Periodic Table » Chlorine » radii of atoms and ions
    Pauling ionic radii. This table shows Pauling radii for chlorine. Ion, Pauling radius / pm, Periodicity link. Cl(I) ...