Fact-checked by Grok 2 weeks ago

Atomic radius

The atomic radius of a is a measure of the size of its neutral atoms, typically expressed in picometers and varying based on the measurement method and bonding context. It is most commonly defined as the , which is one-half the internuclear distance between two identical atoms joined by a single . This definition provides a standardized way to compare atomic sizes across elements, though alternative measures exist to account for different chemical environments. Several types of atomic radii are used depending on the atomic or molecular context. The metallic radius applies to elements in their solid metallic form and is defined as one-half the distance between the nuclei of nearest-neighbor atoms in the crystal lattice. The van der Waals radius, relevant for non-bonded interactions in molecular crystals or gases, is one-half the between the surfaces of two non-bonded atoms of the element. These distinctions arise because atoms do not have fixed boundaries; their effective size depends on cloud overlap and environmental factors. In the periodic table, atomic radii follow predictable trends that reflect and nuclear charge. Radii generally decrease from left to right across a due to increasing , which draws s closer to the without adding new shells. Conversely, radii increase down a group as additional principal electron shells are occupied, shielding inner s and expanding the atomic volume. These patterns underpin much of and predict behaviors like reactivity and bond lengths.

Definitions

Covalent radius

The of an atom is defined as half the internuclear distance between two identical atoms joined by a single . For example, the Cl–Cl bond length in gas is 199 pm, yielding a covalent radius for of 99.5 pm. This measure is particularly applicable to nonmetals, where atoms share electrons in covalent bonds, providing a way to quantify atomic size in molecular contexts. Covalent radii vary depending on bond order, with single-bond radii being longer than those for double or triple bonds due to increased orbital overlap in multiple bonds, which draws the nuclei closer together. established a foundational scale of single-bond covalent radii based on empirical bond lengths, assigning values such as 77 pm for tetrahedral carbon and 70 pm for tetrahedral . On this scale, double-bond radii are shorter (e.g., 67 pm for carbon), and triple-bond radii are even shorter (e.g., 55 pm for nitrogen). These radii enable prediction of bond lengths in molecules by adding the contributions from each atom. For a heteronuclear between atoms A and B, the approximate length is given by d_{AB} \approx r_A + r_B, where r_A and r_B are the respective covalent radii. To account for differences, Pauling introduced a correction: \Delta r = 0.028(\chi_A - \chi_B) , which adjusts the individual radii (with the more electronegative atom having a slightly smaller effective radius) before .

Van der Waals radius

The van der Waals radius represents half the distance between the nuclei of two adjacent non-bonded atoms in a , serving as a measure of the effective size of an atom when it is not involved in covalent bonding. This definition is particularly applicable to systems held together by weak intermolecular forces, such as crystals (e.g., solid ) or molecular solids where non-bonded contacts predominate. For instance, the van der Waals radius of carbon is 170 pm, determined from non-bonded interatomic distances in structures like . In 1964, Arrigo Bondi established a widely adopted scale of van der Waals radii by analyzing crystallographic data from molecular crystals and solids, selecting values that best fit observed interatomic distances. Bondi's compilation provides standardized radii for main-group elements, such as 147 pm for , 152 pm for oxygen, and 155 pm for , which have become benchmarks for estimating atomic sizes in non-bonded contexts. These values reflect averages derived from diverse environments, ensuring consistency across applications. The originates from the equilibrium distance governed by attractive van der Waals forces (dispersion and induction) balanced against short-range repulsion, leading to larger dimensions than covalent radii, which involve stronger shared-electron-pair bonds. For two identical non-bonded atoms, the contact distance is given by d_{\text{vdW}} = 2 \times r_{\text{vdW}}, where r_{\text{vdW}} is the . This formulation is essential for computing molecular volumes, surface areas, and packing coefficients in lattices, aiding in the of molecular and .

Ionic radius

The ionic radius refers to the effective size of a cation or anion within the crystal lattice of an ionic compound, where ions are stabilized by electrostatic attractions. This radius differs markedly from that of the atom due to electron gain or loss; for example, the sodium cation (Na⁺) has an ionic radius of 102 pm for coordination number 6, compared to 186 pm for the sodium atom. The Shannon-Prewitt scale provides a comprehensive set of effective ionic radii based on analyses of interatomic distances from , accounting for variations in oxidation states and coordination environments. These radii are calibrated such that the oxide anion (O²⁻) has a value of 140 pm for coordination number 6, serving as a reference for deriving other ionic sizes. Cations are generally smaller than their parent neutral atoms because removing electrons reduces electron-electron repulsion while the nuclear charge remains the same, resulting in a higher that contracts the . Anions, by contrast, are larger than their neutral atoms as added electrons increase repulsion in the outer shell, expanding the ion's size. These disparities intensify with higher charges: more positive charges shrink cations further, while more negative charges enlarge anions. In ionic bonds, the observed interionic distance is approximately the sum of the cation and anion radii, d \approx r_{+} + r_{-}. The radius ratio r_{+}/r_{-} helps predict and stability; for example, ratios exceeding 0.414 support octahedral coordination in structures like rock salt (NaCl).

Metallic radius

The metallic radius refers to half the internuclear distance between the nearest neighbor atoms in the crystal of a pure metal, providing a measure of atomic size specific to environments. This radius is typically determined from data on metallic crystals and standardized for a of 12, corresponding to close-packed structures like face-centered cubic (FCC) or hexagonal close-packed (HCP). formalized this concept in his analysis of interatomic distances, deriving metallic radii by adjusting observed bond lengths for effective and resonance energy in the metallic . Representative values illustrate how metallic radii vary with and . For in its FCC , the metallic is 128 , reflecting a nearest-neighbor distance of 256 . Iron, in its body-centered cubic (BCC) structure at ( 8), has an actual half nearest-neighbor distance of 124 , based on a nearest-neighbor distance of 248 ; the standard metallic adjusted for 12 is 126 . , adopting an FCC structure, exhibits a metallic of 144 , with a corresponding of 288 . These values are derived from empirical measurements and theoretical adjustments, often aligning closely with Pauling's tabulated radii for 12 (: 128 ; : 126 ; : 144 ).
ElementCrystal StructureMetallic Radius (pm)Nearest-Neighbor Distance (pm)
Copper (Cu)FCC128256
Iron (Fe)BCC124248
Gold (Au)FCC144288
Metallic bonding, involving delocalized valence electrons that form a "sea" surrounding positively charged metal ions, enables efficient close packing in the lattice, resulting in internuclear distances shorter than those in van der Waals interactions (typically 300–400 pm for non-bonded contacts) but generally longer than covalent bond lengths in nonmetallic compounds (often 100–200 pm). This arises because the non-directional, electrostatic nature of metallic bonds allows for variable coordination without the strong, localized sharing of electron pairs seen in covalent bonds. Pauling's model describes metallic bonds as resonating covalent structures, where the effective bond order (around 1 for many metals) leads to radii slightly larger than single covalent bonds but optimized for lattice stability. The metallic , given by the [formula d](/page/Formula_D) = 2 r_m where r_m is the metallic , directly informs calculations of parameters and . For example, in FCC structures, the parameter a = 2 \sqrt{2} r_m, which is used to estimate metallic densities and predict behaviors by modeling atomic packing in solid solutions.

History

Early concepts

The concept of atoms as indivisible and eternal particles originated in , particularly with in the BCE, who envisioned them as the fundamental building blocks of matter without specifying or quantifying their sizes. This qualitative notion of discrete particles persisted through but lacked empirical support or dimensional attributes until the . In the early 19th century, revived and formalized in 1808, proposing that all matter consists of indivisible atoms of finite size that combine in fixed ratios to form compounds, thereby implying inherent atomic dimensions essential for chemical behavior. Building on this, Amedeo Avogadro's 1811 hypothesis stated that equal volumes of different gases at the same temperature and pressure contain equal numbers of molecules, which indirectly linked gaseous volumes to molecular sizes and supported the idea of atoms occupying space. Advancements in the mid-19th century through the provided the first indirect quantitative estimates of atomic sizes. James Clerk Maxwell in 1860 derived effective atomic diameters from gas viscosity measurements, modeling collisions between hard-sphere atoms to explain transport properties like independent of density. extended this framework in the 1860s and 1870s, refining the of molecular interactions to yield atomic diameters on the order of angstroms, though these remained theoretical constructs without direct observation. Jean Perrin's 1908 experiments on offered empirical validation, analyzing the random displacements of suspended particles to estimate atomic scales around $10^{-10} m, aligning with kinetic theory predictions and confirming atoms as tangible entities of measurable size. A pivotal breakthrough came in 1912 when demonstrated diffraction by crystals, revealing interference patterns that directly measured interatomic spacings and marked the transition to precise structural determinations of atomic arrangements.

Key developments

The development of atomic radius concepts accelerated in the early with advancements in , which provided empirical data on atomic separations in crystals. and his son William Lawrence Bragg pioneered the use of X-ray diffraction to determine crystal structures starting in 1913, enabling the measurement of interatomic distances in ionic compounds such as NaCl and allowing for the estimation of ionic radii based on observed lattice parameters. Their work laid the groundwork for distinguishing between ionic and other bond types by quantifying atomic sizes in solid-state environments. The advent of in the 1920s revolutionized the theoretical understanding of atomic sizes by shifting focus from classical models to probabilistic electron distributions. Erwin Schrödinger's formulation of the wave equation in 1926 provided a mathematical framework for describing atomic orbitals, where the size of an atom is interpreted as the spatial extent of the electron probability density, rather than a fixed boundary. Building on this, John C. Slater introduced shielding rules in 1930 to calculate the experienced by electrons, which directly influences orbital contraction and thus atomic radii; these rules approximate the screening by inner electrons, revealing how increased nuclear charge pulls electrons closer, reducing atomic size across the periodic table. Linus Pauling synthesized these quantum insights with experimental data in the 1930s, establishing practical scales for atomic radii. In 1932, he developed the scale to quantify the electron-attracting power of atoms, which correlates with bond polarity and variations in covalent radii between elements. Pauling's seminal 1939 book, , included the first comprehensive table of covalent radii derived from measurements in molecules, standardizing values on a scale where typical single-bond radii hover around 100 for main-group elements like carbon (77 ) and (99 ). This work emphasized how electronegativity differences adjust shared radii in heteronuclear bonds, providing a unified approach to predicting molecular geometries. In 1964, Arnold Bondi published a comprehensive set of van der Waals radii based on molecular crystal data, which became a standard reference for non-bonded atomic interactions. In the 1940s, Anton Eduard van Arkel and Jan Arnold Albert Ketelaar advanced the classification of atomic interactions by introducing a triangular that categorizes bonds as ionic, covalent, or metallic based on electronegativity differences and atomic size ratios, facilitating the distinction of corresponding radius types in compounds. Their framework, first outlined in van Arkel's 1941 textbook and refined in subsequent publications, highlighted how ionic radii dominate in electrovalent crystals, covalent radii in molecular structures, and metallic radii in delocalized systems. These efforts ensured that atomic radii could be reliably applied across disciplines, from to .

Empirical measurements

Determination methods

Experimental determination of atomic radii involves techniques that directly probe interatomic distances in crystals, molecules, or surfaces, yielding empirical values for various radius types such as covalent, ionic, and metallic. serves as the primary method for measuring lattice parameters and bond lengths in crystalline solids, from which radii are inferred. In this technique, X-rays are scattered by the surrounding atomic nuclei, producing diffraction patterns that are analyzed to reconstruct atomic positions via methods or direct refinement. Bond lengths derived from typically achieve an accuracy of approximately 0.01 for heavy atoms, enabling precise estimation of ionic radii in salts (e.g., NaCl) by assuming the additivity of cation and anion contributions, and covalent radii from maps in molecular crystals. Neutron diffraction addresses limitations of XRD for light elements, particularly , which scatter X-rays weakly due to their low . Neutrons interact with atomic nuclei through strong, short-range forces, providing clear signals for positioning light atoms in crystal structures, such as in molecules or networks in hydrates. This method yields interatomic distances comparable to XRD for heavier elements but excels in resolving H positions, with typical precisions of 0.01–0.02 , facilitating accurate covalent and van der Waals radii in hydrogen-containing compounds. Spectroscopic approaches, including gas-phase electron diffraction (GED) and (EXAFS), target specific bonding environments. GED scatters high-energy electrons through gaseous molecules to measure internuclear distances directly, avoiding crystal-packing distortions; it has been instrumental in determining covalent radii for hydrocarbons like and , with bond length uncertainties around 0.002–0.005 . , conversely, probes local coordination in amorphous or disordered materials by examining modulations in absorption spectra above the of a target atom, revealing radial distribution functions and effective s up to several angstroms from the absorber; this is particularly useful for metallic radii in alloys or glasses. Scanning tunneling microscopy (STM) and atomic force microscopy (AFM) provide nanoscale imaging of surface atoms, offering direct observations of atomic protrusions on substrates. STM maps surface electronic density via quantum tunneling currents from a conductive tip, resolving individual atoms like silicon on Si(111) with sub-angstrom lateral resolution, while AFM senses short-range repulsive forces to image insulating surfaces, such as alkali halides, achieving vertical sensitivities below 0.1 Å for surface atomic radii. These techniques are limited at elevated temperatures, where anharmonicity in atomic vibrations introduces thermal broadening, reducing positional accuracy in diffraction-based methods like XRD by up to 10–20% due to non-Gaussian displacement distributions.

Tabulated values

Tabulated values of atomic radii are compiled from empirical measurements of lengths and interatomic distances in various compounds and . These compilations, such as those in the CRC Handbook of Chemistry and Physics, provide standardized data for different radius types, drawing from seminal works like Pauling's covalent radii (1931) and Bondi's van der Waals radii (1964). A modern empirical set for covalent radii was compiled by Cordero et al. (2008) from crystallographic data. Covalent radii, defined as half the single-bond distance between identical atoms, are widely used for nonmetals and represent bonding sizes. Below is a selection of empirical covalent radii in picometers (pm) for representative elements, illustrating the range from (31 pm) to (140 pm).
ElementCovalent Radius (pm)
H31
C76
N71
O66
F57
Cl102
Br120
I139
Xe140
Data from Cordero et al. (2008). For example, across period 2 nonmetals (C to F), values decrease from 76 pm to 57 pm, reflecting empirical trends. Van der Waals radii measure non-bonding contact distances and are larger than covalent radii, typically 1.5–2 times greater. Bondi's 1964 , derived from molecular structures, is a standard reference, with values often extrapolated from limited solid-state data. These radii range from (140 pm) to (202 pm) for .
ElementVan der Waals Radius (pm, Bondi)
He140
Ne154
Ar188
Kr202
Xe216
H120
O152
Cl175
Br185
I198
Compiled from Bondi's analysis of over 100 crystal structures. Uncertainties arise for due to , with differences up to 10 pm between scales. Ionic radii, applicable to ions in crystals, depend strongly on (CN) and charge, with values increasing by about 10% for higher CN in metallic ions. Shannon's 1976 effective ionic radii, based on refinements of over 1000 and structures assuming fixed O²⁻ (140 pm) and F⁻ (133 pm) radii, are the most comprehensive set. For instance, Li⁺ at CN=6 is 76 pm, while at CN=4 it is 59 pm.
IonIonic Radius (pm)CN
Li⁺594
Li⁺766
Na⁺1026
O²⁻1406
F⁻1336
These values from Shannon's tables in Acta Crystallographica show coordination dependence, with higher CN leading to larger radii due to expanded coordination spheres. Metallic radii apply to elements in their solid metallic form and are defined as half the distance between nearest-neighbor atoms in the crystal lattice. Values depend on and (CN), typically 8–12 for metals. Below is a selection of empirical metallic radii in pm for representative metals.
ElementMetallic Radius (pm)CN
1558
1868
2318
12812
14412
Data compiled from standard references like the (values for typical CN in elemental structures). Overall, empirical tables like those in the and IUPAC-referenced data ensure consistency across scales, though metallic radii may vary by up to 10% with CN.

Across periods

In the periodic table, the atomic radius decreases from left to right across a due to the increasing experienced by electrons. This trend arises because each successive element adds a proton to the , enhancing its positive charge, while the additional occupies the same principal quantum , offering minimal shielding from this increased attraction. As a result, electrons are drawn closer to the , contracting the overall atomic size without the addition of new electron shells. For example, in period 2, the calculated atomic radius diminishes from 167 pm for to 38 pm for . This left-to-right shrinkage is evident in the transition from s-block metals, which exhibit larger radii, to p-block nonmetals with smaller sizes. Covalent radii, often used for nonmetals, typically decrease by approximately 20–30 pm per element across a , underscoring the progressive . In 3, for instance, the atomic radius falls from 190 pm for sodium to 99 pm for , illustrating how alkali metals maintain extended sizes compared to in the same row. The poor shielding effectiveness of electrons within the same shell exacerbates this trend, as they incompletely screen the nucleus from the growing proton count, thereby amplifying the pull on valence electrons.

Down groups

As elements descend a group in the periodic table, their atomic radii generally increase due to the addition of successive electron shells, which occupy higher principal quantum numbers and extend farther from the nucleus. This trend results in an enlargement of approximately 50-100 pm per period, reflecting the growing spatial extent of the electron cloud. For instance, in group 1, the metallic radii progress from at 157 pm to sodium at 191 pm, at 235 pm, and cesium at 272 pm, illustrating a steady that facilitates the increasing metallic character down the group. Similarly, for the halogens in group 17, covalent radii increase from at 71 pm to at 99 pm, at 114 pm, and iodine at 133 pm, underscoring the trend across nonmetals as well. The primary driver of this vertical increase is the , where inner-core electrons repel valence electrons, thereby reducing the experienced by the outer electrons and allowing them to occupy larger orbitals. Although the nuclear charge rises with , the additional shielding from new inner shells outweighs this pull, leading to a net expansion of the atomic radius; this is particularly evident in metallic radii, which show consistent growth without significant contraction. In contrast to the contraction observed across periods due to unshielded nuclear attraction, the down-group trend emphasizes the dominance of increasing principal quantum levels. The rate of increase varies across the periodic table, with faster growth in the early periods (such as from the second to third) compared to later ones, where the increments per become more gradual. In the p-block elements, this growth slows notably between the fifth and sixth periods due to the in the preceding f-block, which imperfectly shields the nuclear charge and results in smaller-than-expected radii for elements like relative to . Overall, these patterns highlight how shielding and addition govern the vertical periodicity in atomic size.

Contractions and exceptions

The lanthanide contraction describes the gradual decrease in atomic and ionic radii across the lanthanide series (from to ), resulting from the ineffective shielding of the increasing nuclear charge by 4f electrons. This poor shielding causes a stronger pull on the outer electrons, leading to radii that are 10-20 pm smaller than anticipated for elements immediately following the lanthanides in the periodic table. For instance, (Hf) exhibits an atomic radius of 159 pm, nearly identical to that of (Zr) at 160 pm, despite the expectation of a larger size for the heavier 5d-series element based on typical down-group trends. Similarly, the arises from the suboptimal shielding provided by (or analogous nd) electrons in transition metals, which allows the nuclear charge to increase more effectively across a . This results in only a modest decrease in atomic radii from left to right in the d-block, but it notably affects elements in groups 11 and 12, making them smaller than would be predicted by extrapolation from s-block trends. Potassium (), an s-block element, has an atomic radius of 235 pm, whereas (Cu) in a comparable horizontal position shows a much smaller radius of 128 pm due to this enhanced . Other exceptions include relativistic effects prominent in heavy elements, which accelerate inner electrons and contract s and p orbitals while expanding d orbitals. In (Au), this leads to a notable contraction of the 6s orbital, yielding an atomic radius of 144 pm and contributing to its unique and color. Additionally, in the p-block, the —where the ns² electrons of heavier elements (e.g., in groups 13-15) are reluctant to participate in bonding due to relativistic stabilization—creates size anomalies in lower oxidation states, as these electrons remain more tightly bound closer to the . These contractions have significant chemical implications, particularly in coordination chemistry and . The reduced radii of 5d metals, stemming from , enable tighter binding of ligands, which increases metal-ligand bond strengths and enhances catalytic activity in processes like olefin and ; for example, complexes of (4d) versus (5d) show stronger interactions and higher selectivity in the latter due to the comparable yet more charged sizes.

Calculated radii

Theoretical models

The , introduced in 1913, offers a foundational theoretical approach to estimating atomic radii for hydrogen-like atoms, treating electrons as orbiting the in discrete shells. The radius of the nth orbit is given by r_n = \frac{n^2 a_0}{Z}, where n is the principal quantum number, a_0 = 52.9 pm is the (the ground-state radius of ), and Z is the . For the in its (n=1, Z=1), this yields r_1 = 52.9 pm, providing a benchmark for atomic size. For the in its (n=1, Z=1), this yields r_1 = 52.9 pm, providing a benchmark for atomic size. To extend the model to multi-electron atoms, [Z](/page/Z) is replaced by an effective nuclear charge [Z_\text{eff}](/page/Effective_nuclear_charge), which accounts for partial shielding of the by inner electrons, reducing the net attraction felt by electrons. This adjustment allows approximate radii calculations but relies on estimating [Z_\text{eff}](/page/Effective_nuclear_charge), as the simple Bohr framework assumes circular orbits and neglects electron-electron repulsions. Slater's rules, developed in 1930, provide a semi-empirical method to compute Z_\text{eff} = Z - \sigma, where \sigma is a shielding constant derived from the electron configuration. Electrons are grouped by principal quantum number into shells (e.g., 1s, 2s2p), and \sigma is calculated with contributions such as 1.00 from each electron in inner shells, 0.85 from electrons in the (n-1) shell, 0.35 from other electrons in the same group (ns,np), and 0 for electrons in higher groups. For valence electrons, this yields Z_\text{eff} values like 3.25 for carbon's 2p electrons (Z=6, \sigma=2.75) and 6.10 for chlorine's 3p electrons (Z=17, \sigma=10.90), which can then be inserted into the modified Bohr formula to estimate radii. These rules simplify wave function approximations and predict trends effectively for ionization energies and sizes. Empirical correlations further link atomic radii to other properties without full quantum calculations. In the 1870s, estimated atomic sizes indirectly through atomic volumes, calculated as divided by , which revealed periodic variations and predicted properties for undiscovered elements like . Similarly, the Allred-Rochow electronegativity scale (1958) relates \chi_\text{AR} to r via \chi_\text{AR} = 0.744 + \frac{0.359 Z_\text{eff}}{r^2}, where Z_\text{eff} is from Slater's rules, enabling radii to be inferred from measured electronegativities (e.g., solving for r in fluorine yields ~72 pm). These models, while insightful, have limitations due to their simplifications. The Bohr model fails for multi-electron atoms beyond hydrogen-like systems, as it ignores electron repulsion and orbital shapes, leading to inaccuracies in spectral predictions and sizes. Slater's rules provide fair accuracy for main-group elements by approximating shielding but oversimplify for transition metals, where d-electron penetration and poor shielding cause deviations in Z_\text{eff} and thus radii estimates. Empirical correlations like Mendeleev's and Allred-Rochow's depend on experimental inputs, limiting their predictive power for heavy or exotic elements.

Computational methods

Computational methods for determining atomic radii rely on quantum mechanical approaches that solve the or its relativistic counterparts to obtain wavefunctions or densities, from which radii are extracted as values or density contours. The Hartree-Fock () method, a foundational self-consistent field approach, approximates the many-electron wavefunction as a single and iteratively solves for atomic orbitals while accounting for -electron interactions via mean-field potentials. The atomic radius in HF calculations is commonly defined as the value \langle r \rangle = \int r |\psi(r)|^2 dV for the valence orbital, representing the average radial position of s, or as the radius of the 90% contour, where 90% of the electron probability lies within a of that size. For the , HF calculations using Gaussian basis sets yield \langle r \rangle \approx 53 pm, matching the closely since HF is exact for one-electron systems. These methods are implemented in software packages such as Gaussian and , which employ basis set expansions to numerically solve the equations efficiently. Density Functional Theory (DFT) extends these ideas by focusing on the \rho(\mathbf{r}) rather than the many-body wavefunction, solving the Kohn-Sham equations—a set of single-particle equations analogous to but incorporating exchange-correlation effects through an approximate functional. The atomic radius is derived from the resulting , often as the corresponding to half the in homonuclear dimers or the 90% contour for isolated atoms. like B3LYP, which blend exact exchange with -based approximations, provide covalent radii that agree with experimental values to within about 5 pm for many elements, offering a balance of accuracy and computational cost for multi-electron atoms. For instance, B3LYP calculations on main-group elements reproduce with mean absolute deviations of 2-3 pm compared to gas-phase data. These approaches build on earlier theoretical models by delivering precise numerical predictions without empirical parameters. For heavy elements, relativistic effects significantly influence atomic sizes, necessitating incorporation of the into or DFT frameworks to account for high velocities near the . In Dirac-Hartree-Fock (DHF) or relativistic DFT, the four-component wavefunctions capture spin-orbit coupling and mass-velocity corrections, leading to of s-orbitals and of p- and d-orbitals. A prominent example is mercury (), where relativistic effects contract the 6s orbital by approximately 20%, reducing the atomic radius and contributing to its low by weakening . Such calculations, performed using Gaussian basis sets with relativistic pseudopotentials or all-electron methods in , reveal deviations of up to 10-15% in non-relativistic approximations for Z > 50.

References

  1. [1]
    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 ...Missing: authoritative sources
  2. [2]
    4.3 Periodic Trends in the Size of Atoms – Chemistry Fundamentals
    In the periodic table, atomic radii decrease from left to right across a row and increase from top to bottom down a column. Because of these two trends, the ...
  3. [3]
    Atomic Radii - Chemistry LibreTexts
    Apr 28, 2024 · Definition. Atomic radius is generally stated as being the total distance from an atom's nucleus to the outermost orbital of electron.Missing: authoritative sources
  4. [4]
    CH150: Chapter 2 - Atoms and Periodic Table - Chemistry
    ... atom as you go across a period. So the overall periodic trend for atomic radius (size) is that atoms get smaller as you go across a period, and they get ...
  5. [5]
    Property Trends in the Periodic Table - Chemistry 301
    Since all atoms and ions are spherical, there size is typically given as a radius. The size of an atom (ion) is referred to as its atomic (ionic) radius. As ...
  6. [6]
    [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.
  7. [7]
    Experimental data for Cl 2 (Chlorine diatomic)
    Bond length · Rotational Constant · Moment of Inertia · Dipole ... 0.0000, 1.9879. Atom - Atom Distances bond lengths. Distances in Å. Cl1, Cl2. Cl1, 1.9879. Cl2 ...Missing: authoritative source
  8. [8]
    van der Waals Volumes and Radii | The Journal of Physical Chemistry
    Research Article March 1, 1964. van der Waals Volumes and Radii. Click to copy article linkArticle link copied! A. Bondi. ACS Legacy Archive. Open PDF. The ...Missing: original | Show results with:original
  9. [9]
    Consistent van der Waals Radii for the Whole Main Group
    Bondi's van der Waals radii represent average or typical values, and they cannot be equated precisely to a distance that can be uniquely defined in terms of ...Introduction · Methodology · Results · ReferencesMissing: original | Show results with:original
  10. [10]
    Sodium Element Facts - Chemicool
    Atomic radius, 186 pm. Ionic radius (1+ ion), 116 pm. Ionic radius (2+ ion) ... Source: Due to its high reactivity, sodium is found in nature only as a ...
  11. [11]
    Ionic Coordination and Pauling's Rules - Tulane University
    Sep 26, 2014 · Rule 1. Around every cation, a coordination polyhedron of anions forms, in which the cation-anion distance is determined by the radius sums and ...
  12. [12]
    Atomic Radii and Interatomic Distances in Metals - ACS Publications
    Click to copy article linkArticle link copied! Linus Pauling. ACS Legacy ... https://doi.org/10.1021/acs.jpca.5c01901. Joseph M. Parr, Cameron Jones ...
  13. [13]
    Metallic, Covalent and Ionic Radii(r)* - Wired Chemist
    Metallic radii for 12-coordination are given for all metals. Covalent radii are in parentheses. Ionic radii are for six-coordination.
  14. [14]
    A Brief History of Chemistry
    Dalton's atomic theory of matter contains four fundamental hypotheses: All matter is composed of tiny indivisible particles called atoms. All atoms of an ...
  15. [15]
    [PDF] An Introduction to the Atomic Theory - Thomas Aquinas College
    Are Avogadro's hypotheses consistent with Berzelius's theory of electro-chemistry and Dalton's supposition that particles of the same kind repel each other ...
  16. [16]
    New Foundations Laid by Maxwell and Boltzmann (1860 - 1872)
    Aug 24, 2001 · Maxwell showed that if the kinetic theory of gases is correct, both expectations will be wrong, because the mechanism that produces viscosity is ...
  17. [17]
    [PDF] Einstein's random walk
    Physics in motion – Einstein's theory of Brownian motion allowed Jean Baptiste Perrin to demonstrate the existence of atoms in 1908. Perrin, who won the 1926 ...
  18. [18]
    Single-crystal X-ray Diffraction - SERC (Carleton)
    May 17, 2007 · Max von Laue, in 1912, discovered that crystalline substances act as three-dimensional diffraction gratings for X-ray wavelengths similar to ...
  19. [19]
    The reflection of X-rays by crystals - Journals
    Bragg William Henry and; Bragg William Lawrence. 1913The reflection of X-rays by crystalsProc. R. Soc. Lond. A88428–438http://doi.org/10.1098/rspa.1913.0040 ...
  20. [20]
    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 involving ...Missing: angstrom | Show results with:angstrom
  21. [21]
    Estimating Atomic Sizes with Raman Spectroscopy | Scientific Reports
    Mar 19, 2013 · The radii reported in the literature are mainly determined from X-ray diffraction and neutron diffraction technologies. It can be seen that ...
  22. [22]
    Current status of neutron crystallography in structural biology - PMC
    Neutrons are better for observing light elements, such as hydrogen or lithium, than X-rays because neutrons interact directly with the nuclei of atoms and the ...Diffraction Data Measurement... · Figure 1 · Figure 3
  23. [23]
    Neutron Diffraction - an overview | ScienceDirect Topics
    Similar to X-ray diffraction, neutron diffraction (ND) measures the atomic structure of materials. The mechanism is essentially the same, except that here the ...
  24. [24]
    Carbon—Carbon Bond Distances. The Electron Diffraction ...
    The electron diffraction investigation of ethane, propane, isobutane, neopentane, cyclopropane, cyclopentane, cyclohexane, allene, ethylene, isobutene, ...
  25. [25]
    Structure determination of amorphous materials via exafs
    Abstract. An improved data reduction routine was developed and computer programs were written for treating EXAFS absorption spectra. By judicious application of ...
  26. [26]
    Unveiling the local structure of the amorphous metal Fe ( 1 − x ) Zr x ...
    Mar 27, 2023 · Extended X-ray absorption fine structure (EXAFS) data are sensitive to the short-range order and have been used to qualitatively validate ...
  27. [27]
    Detecting elusive surface atoms with atomic force microscopy - PNAS
    With the STM, quantum mechanical tunneling of electrons from a sharp metal tip to a conductive substrate is used to detect individual surface atoms with atomic ...
  28. [28]
    Quantifying the evolution of atomic interaction of a complex surface ...
    Aug 24, 2020 · AFM is capable of resolving the lattice of insulating substrates with atomic resolution. Previously, Ellner et al. considered AFM images of Cl ...<|control11|><|separator|>
  29. [29]
    Anharmonic Thermal Motion Modelling in the Experimental XRD ...
    The results of the Hamilton test strongly support the inclusion of anharmonicity into the thermal motion modelling of the 1-MUR crystal, even if the data ...
  30. [30]
    [PDF] CRC Handbook of Chemistry and Physics
    These include tables on atomic and molecular polarizabilities, diffusion in gases and liquids, vapor pressure and density of mercury, ionic radii in crystals, ...
  31. [31]
    Covalent radius » Periodic table gallery - Mark Winter
    When two atoms of the same kind are bonded through a single bond in a neutral molecule, then one half of the bond length is referred to as the covalent radius.Missing: scale | Show results with:scale
  32. [32]
    The first property to explore is atomic radius.
    Another way to explain the trend in atomic radius across a period is in terms of the effective nuclear charge experienced by the valence electrons. Since ...
  33. [33]
    [PDF] Effective Nuclear Charge (Zeff) and Atomic Radius
    Considering the trend of Zeff (effective nuclear charge) explain the trend of atomic radius across a period in terms of effective nuclear charge and ...
  34. [34]
    Atomic radii (Clementi) » period 2 - WebElements Periodic Table
    The values given here for atomic radius are calculated values using methods outlined in reference 1.
  35. [35]
    [PDF] The periodic trends in atomic radius are
    The periodic trends in atomic radius are: • As you go left to right across a period: Zeffective increases moving to the right, since the number of protons is.
  36. [36]
    Atomic radii (Clementi) » period 3 - WebElements Periodic Table
    Visual periodicity representations for atomic radii, covalent radii, and van der Waals radii. Ionic radii are also available.
  37. [37]
    Property Trends in the Periodic Table - Chemistry 301
    Effectively nuclear charges varies fairly systematically across a period and down a group, thus the idea of periodic trends. There are always excepts to these ...
  38. [38]
    Hybridization Trends for Main Group Elements and Expanding the ...
    Apr 28, 2014 · This work presents trends in hybridization of B, C, N, O, and F in HnX–YHm compounds, where the atom X belongs to the second period of the ...
  39. [39]
    1.18: The Effects of Shielding on Periodic Properties
    Aug 28, 2023 · Down the periodic table, larger atomic radii cause electrons in valence orbitals to be shielded by core electrons. Recall that shielding ...
  40. [40]
    Periodic Trends in Physical Properties of Elements - Atomic Radius
    Atomic radius includes covalent radius (half the distance between two bonded atoms), metallic radius (half the distance between adjacent metal atoms), and Van ...
  41. [41]
    2.2: Trends in Size - Chemistry LibreTexts
    Jul 16, 2020 · The atomic radius for the halogens increases down the group as n increases. ... atomic radius decreases as Z increases; for example, from K to Kr.
  42. [42]
    Lanthanide Contraction - Chemistry LibreTexts
    Jun 30, 2023 · The graph shows the atomic radius decreasing as the atomic number is increasing, Lanthanide Contraction. Shielding and its Effects on Atomic ...
  43. [43]
    Atomic Radius | The Elements Handbook at KnowledgeDoor
    Our table of atomic radii covers 84 elements. Each value has a full citation identifying its source. The integrated unit conversion calculator can quickly ...
  44. [44]
    The Lanthanide Contraction Revisited - PMC - PubMed Central
    The quadratic nature of the lanthanide contraction is derived theoretically from Slater's model for the calculation of ionic radii. In addition, the sum of all ...
  45. [45]
    Relativistic Effects in the Electronic Structure of Atoms | ACS Omega
    Sep 22, 2017 · Periodic trends in relativistic effects are investigated from 1 H through 103 Lr using Dirac–Hartree–Fock and nonrelativistic Hartree–Fock calculations.
  46. [46]
    Effects of relativistic motion of electrons on the chemistry of gold and ...
    The contraction is due to the relativistic mass increase as followed by the decrease of the Bohr radius. The contraction of the spherical p 1 / 2 subshell is ...
  47. [47]
    [PDF] The Lanthanide Contraction beyond Coordination Chemistry ...
    Abstract: Lanthanide chemistry is dominated by the 'lanthanide contraction', which is conceptualized traditionally through coordination chemistry.
  48. [48]
    Do f Electrons Play a Role in the Lanthanide−Ligand Bonds? A ...
    The present study addresses for the first time the role of f electrons in lanthanide-ligand bonds for the complete Ln series.
  49. [49]
    6.2 The Bohr Model - Chemistry 2e | OpenStax
    Feb 14, 2019 · ... Bohr radius, with a value of 5.292 × × 10−11 m: r = n 2 Z a 0 ... Bohr's model of the hydrogen atom provides insight into the behavior ...
  50. [50]
  51. [51]
    Atomic Shielding Constants | Phys. Rev.
    Atomic Shielding Constants. J. C. Slater. Jefferson Physical Laboratory ... J. C. Slater, Phys. Rev. 34, 1293 (1929); J. C. Slater, Phys. Rev. 35, 509 (1930).Missing: paper | Show results with:paper
  52. [52]
    Atomic size, ionization energy, polarizability, asymptotic behavior ...
    Sep 21, 2012 · Hartree–Fock radii. The two most common radii employed from Hartree ... expectation value of r, 〈r〉 [2]. For a hydrogenic atom these ...
  53. [53]
    radii of atoms and ions - Hydrogen - WebElements Periodic Table
    Atomic radius of the chemical elements displayed on a miniature periodic table ... Hartree-Fock wave functions and radial expectation values: hydrogen to ...
  54. [54]
    Basis Sets | Gaussian.com
    May 17, 2021 · The following table lists polarization and diffuse function availability and the range of applicability for each built-in basis set in Gaussian 16.
  55. [55]
    Additive Covalent Radii for Single-, Double-, and Triple-Bonded ...
    An accurate gas-phase microwave value of 192.87 pm (47) is available for ... covalent radius” per element, for the particular bonding situation. The ...
  56. [56]
    Detailed Density Functional Theory Study of the Cationic ...
    Sep 22, 2022 · The single bond covalent radius of F is 0.71 Å and 1.48 Å for Zr ... B3LYP without sacrificing too much accuracy in this particular case.Scheme 2 · Atomic Charges · Qtaim Bond Paths And Nbo...
  57. [57]
    [PDF] Why is mercury liquid? Or, why do relativistic effects not get into ...
    Sep 18, 2018 · The relativ- istic velocity-mass contraction of the radius of the 6s orbital of the freeAu(g) atom has been calculated to about 16% (8).