Fact-checked by Grok 2 weeks ago

Polarizability

Polarizability is a fundamental property in physics and that quantifies the ease with which the distribution in an or can be distorted by an external , resulting in the formation of an induced . This induced dipole moment \vec{p} is linearly related to the applied \vec{E} through the equation \vec{p} = \alpha \vec{E}, where \alpha is the polarizability, a tensor quantity with units typically expressed in cubic angstroms (ų) or 10^{-40} F m² in SI units. For isotropic systems, such as spherical s, \alpha is a scalar value, while for anisotropic molecules, it is described by a tensor reflecting directional dependencies. In atomic systems, polarizability arises primarily from the displacement of electrons relative to the under the of the field, with values increasing with atomic size and decreasing due to looser binding of electrons. For example, the polarizability of is approximately 0.67 × 10^{-24} cm³, while helium's is about 0.20 × 10^{-24} cm³ owing to its higher excitation energy. Molecular polarizability extends this concept to include contributions from distortion, atomic vibrations, and, in polar molecules, partial orientation of permanent dipoles, though the latter is distinct from pure polarizability effects. Atomic polarizabilities within molecules are often partitioned using models like the LoProp to analyze local responses. Polarizability plays a crucial role in determining dielectric properties of materials, where the macroscopic polarization \vec{P} is given by \vec{P} = N \alpha \vec{E}, with N as the of atoms or molecules, linking it to the \epsilon_r via relations like the Clausius-Mossotti equation: (\epsilon_r - 1)/(\epsilon_r + 2) = N \alpha / (3 \epsilon_0). In chemistry, it governs intermolecular interactions, particularly and dispersion forces, which scale with \alpha^2 and influence molecular recognition, reactivity, and properties like boiling points in nonpolar compounds. Additionally, polarizability affects spectroscopic observables, such as Raman intensities, and is essential in computational modeling of biomolecular systems using polarizable force fields to accurately capture electrostatic effects.

Fundamentals of Polarizability

Definition and Physical Interpretation

Polarizability refers to the ease with which the of an , , or material can be distorted by an external , resulting in the formation of an induced . It is quantitatively defined as the ratio of the magnitude of the induced dipole moment \mathbf{p} to the strength of the applied \mathbf{E}, expressed as \alpha = \frac{p}{E} in the scalar approximation for isotropic systems. This measure captures the responsive nature of the charge distribution within the system to external perturbations, where the induced dipole arises from the temporary displacement of electrons relative to the positively charged nuclei. A key distinction exists between permanent dipoles, which are inherent asymmetries in the charge distribution of polar molecules (such as ), and induced dipoles, which are transiently created in otherwise symmetric, nonpolar entities under an . Polarizability specifically quantifies the to forming these induced dipoles by assessing the deformability of the cloud, rather than the alignment of pre-existing permanent dipoles. In nonpolar systems, the external shifts the center of negative charge away from the positive core, creating a separation that scales linearly with for weak fields. Common examples of polarizable systems include neutral atoms like , where the field induces a by polarizing the spherical cloud; nonpolar molecules such as oxygen (O₂), which acquire a through displacement; and bulk materials like dielectrics, where collective leads to macroscopic effects such as reduced internal field strength. In these cases, polarizability governs phenomena ranging from intermolecular forces to the behavior of insulators in . For anisotropic molecules, the response is more generally described by the electric polarizability tensor, which accounts for directional dependencies. Several qualitative factors influence the magnitude of polarizability. Larger or molecular , corresponding to greater electron-nuclear separation, allows electrons to be more easily displaced, increasing polarizability, as seen in heavier atoms compared to lighter ones in the same group. Similarly, higher , particularly in valence shells with more loosely bound electrons, enhances the distortability of the cloud, making systems with more electrons generally more polarizable. These factors underscore polarizability's role as a measure of electronic "softness" in response to electrostatic influences.

Historical Background

The concept of polarizability emerged in the early through investigations into dielectrics, beginning with Michael Faraday's experimental observations of induction effects in insulating materials. In his eleventh series of experiments, Faraday demonstrated that dielectrics exhibit a specific inductive capacity when subjected to electric forces, effectively polarizing the material and altering its response to external fields, which laid the groundwork for understanding molecular deformation under electrical influence. This foundational work was formalized within electromagnetic theory by James Clerk Maxwell in the 1860s, who incorporated dielectric polarization into his dynamical model of the . accounted for the arising from time-varying polarization in dielectrics, unifying , , and propagation while treating polarizable media as responsive to electromagnetic . Building on these ideas, the Clausius-Mossotti relation, developed in the mid-19th century, provided an early link between macroscopic dielectric constants and microscopic polarizability, assuming local fields influence atomic responses. A significant advancement came in 1912 with Peter Debye's theoretical treatment of dipolar polarization, which connected molecular polarizability directly to dielectric constants through an equation accounting for both electronic distortion and orientational effects in polar substances. In the quantum era, Max Born's 1920 contributions to crystal lattice dynamics incorporated ionic polarizability into models of cohesive forces in ionic crystals, enabling predictions of lattice stability and vibrational properties. During the 1930s, applied to analyze bond polarities and deformability in molecules, using to quantify how sharing influences polarizability in covalent and polar bonds. Post-1950s computational developments marked a shift toward precise quantum mechanical calculations, with Hartree-Fock approximations enabling the first reliable estimates of ionic polarizabilities, as demonstrated by R. M. Sternheimer's 1954 work deriving electronic distortions from self-consistent field wave functions. By the 1960s, methods extended these to molecular systems, with early calculations using Gaussian basis sets to compute polarizabilities for small molecules like H2 and CH4, establishing benchmarks for electron correlation effects.

Electric Polarizability

Mathematical Formulation

The electric polarizability quantifies the linear response of a system's induced to an applied . For isotropic systems, such as atoms or spherical molecules, the polarizability is a scalar quantity \alpha, defined by the relation \mathbf{p} = \alpha \mathbf{E}, where \mathbf{p} is the induced and \mathbf{E} is the ; in component form along the field direction, this simplifies to \alpha = p_i / E_i. The SI units of \alpha are \mathrm{C}^2 \mathrm{m}^2 \mathrm{J}^{-1}, equivalent to \mathrm{F} \mathrm{m}^2 ( square meters), though it is commonly expressed in cubic angstroms (\AA^3) in and molecular contexts, with $1 \AA^3 \approx 1.11265 \times 10^{-40} \mathrm{C}^2 \mathrm{m}^2 \mathrm{J}^{-1}. In anisotropic systems, such as non-spherical molecules or , the polarizability is described by a second-rank tensor \boldsymbol{\alpha}, relating the induced to the field via p_i = \sum_{j=1}^3 \alpha_{ij} E_j, where \alpha_{ij} forms a $3 \times 3 (\alpha_{ij} = \alpha_{ji}) due to the reciprocity of energy in . This tensor can often be diagonalized in the principal axis frame, yielding components \alpha_{xx}, \alpha_{yy}, and \alpha_{zz} that reflect directional variations in response. The mean (or average) polarizability \bar{\alpha} is then obtained from the of the tensor as \bar{\alpha} = (\alpha_{xx} + \alpha_{yy} + \alpha_{zz})/3, providing an isotropic equivalent for comparisons or bulk approximations. For a dilute gas of N non-interacting isotropic molecules per unit volume, the macroscopic \mathbf{P} is \mathbf{P} = N \alpha \mathbf{E}, leading to the (dielectric constant) \varepsilon_r = 1 + N \alpha / \varepsilon_0, where \varepsilon_0 is the ; this relation neglects local field corrections that arise in denser media. In general, the polarizability \alpha (scalar or tensor components) is frequency-dependent, arising from the dynamic response to oscillating fields, though detailed derivations are addressed elsewhere.

Measurement Techniques

One common method for measuring electric polarizability in gases and liquids at low densities involves determining the n and using the Lorentz-Lorenz equation in its approximate form, n^2 - 1 = \frac{N \alpha}{\epsilon_0}, where N is the , \alpha is the polarizability, and \epsilon_0 is the ; this relates the macroscopic to the microscopic polarizability assuming dilute conditions where local field corrections are negligible. This technique has been applied to various fluids, including light crude oils, by evaluating molar refractive indices and volumes to extract electronic polarizabilities. For species in gaseous s, polarizability can be determined by observing the deflection of a collimated neutral beam in an inhomogeneous , where the force on each atom is \mathbf{F} = \frac{1}{2} \alpha \nabla E^2, leading to a deflection proportional to \alpha and the field gradient \nabla E; this method was pioneered for atoms like sodium and using velocity-selected beams to achieve precise measurements. More recent implementations, such as atom for , have refined this approach by comparing deflections to enhance accuracy in isolating scalar polarizability components. Stark effect spectroscopy provides another direct route to extract polarizability from atoms and molecules by applying a uniform E and measuring the quadratic shift in lines, given by \Delta E = -\frac{1}{2} \alpha E^2, which shifts or molecular levels and allows \alpha to be derived from the field-dependent changes. This technique has been used effectively for cesium states, such as the $7p ^2P_{1/2,3/2} levels, by analyzing absorption spectra under controlled fields to quantify scalar static polarizabilities with uncertainties below 1%. For tensor components, the method can be extended by resolving Stark sublevels, though it primarily yields the average scalar value. Modern techniques enhance precision for specific phases; for gases, (CRDS) enables high-sensitivity measurements of variations or field-induced effects, achieving accuracies around 1% for like and by monitoring light decay in an filled with the sample. For solids and surfaces, measures changes in light polarization upon reflection to determine the function, from which effective polarizability is extracted via relations like \epsilon_r = 1 + \frac{N \alpha}{\epsilon_0}, particularly useful for thin films and crystalline materials with sub-nanometer resolution. In dense media such as liquids and solids, measurements face challenges from effects, where the effective field at a differs from the applied field due to surrounding ; the Onsager model addresses this qualitatively by treating the molecule in a spherical cavity within a , providing a reaction field correction to relate macroscopic to microscopic polarizability without assuming Lorentz local fields.

Polarizability in Atoms and Molecules

Atomic Polarizabilities

Atomic polarizabilities describe the response of isolated atoms to an external electric field, characterized by the induced dipole moment proportional to the field strength. For atoms, this property is isotropic due to spherical symmetry in the ground state, and it generally increases with atomic number within a group in the periodic table. This trend arises from the larger size of the electron cloud in heavier atoms, which allows for greater distortion by the field; for example, the static dipole polarizability rises from 0.205 ų for helium to 4.04 ų for xenon in the noble gas group. Across a period, polarizability decreases from left to right as the effective nuclear charge increases, contracting the electron cloud and reducing deformability. Quantum mechanically, the static polarizability \alpha of an atom in its |0\rangle is derived from second-order applied to the interaction H' = -\vec{\mu} \cdot \vec{E}, where \vec{\mu} is the electric and \vec{E} is the . The expression is \alpha = 2 \sum_{k \neq 0} \frac{|\langle 0 | \mu_z | k \rangle|^2}{E_k - E_0}, summed over all excited states |k\rangle with energies E_k > E_0, assuming the field is along the z-axis. This formula captures the virtual excitations of the electron cloud, with contributions weighted by transition moments and energy denominators. For practical computations in multi-electron atoms, the sum is often approximated by including dominant excited states or using basis set expansions. In multi-electron atoms, additivity approximations simplify calculations by treating the total polarizability as a sum of contributions from individual electrons or orbitals, modeled using Slater-type orbitals to account for screening effects. These single-electron polarizabilities are estimated based on the orbital's radial extent and , providing a tractable way to approximate the full many-body response without explicit summation over all states. Such methods, while approximate, align well with exact values for lighter atoms and facilitate scaling to heavier systems. Experimental values for atomic polarizabilities are obtained through techniques like atomic beam deflection in inhomogeneous fields, which directly measure the Stark shift. The following table summarizes recommended static dipole polarizabilities for and metals, converted to ų (with uncertainties where available), highlighting the increase down each group.
ElementPolarizability (ų)
He0.205 ± 0.000003
0.395 ± 0.000004
1.64 ± 0.001
2.49 ± 0.003
Xe4.04 ± 0.03
24.3 ± 0.00007
24.1 ± 0.07
43.0 ± 0.04
47.4 ± 0.04
59.4 ± 0.1
These values, primarily from high-precision coupled-cluster calculations benchmarked against experiments like beam deflection for (yielding 24.1 ų), underscore the larger polarizabilities of compared to due to their in extended s-orbitals. Near thresholds, the distinction between static (\omega = 0) and dynamic polarizabilities becomes pronounced, as the latter \alpha(\omega) incorporates dependence via time-dependent . Dynamic polarizabilities exhibit enhanced values or resonances when the field \omega approaches energies, including the , leading to divergences in Rydberg-like states or near-threshold behaviors that static values do not capture. This effect is critical for understanding and high-field interactions in atomic systems.

Molecular Polarizabilities

Molecular polarizabilities describe the response of a molecule's cloud to an external , extending the concept from isolated atoms by incorporating intramolecular s and geometric arrangements. Unlike atomic polarizabilities, which treat atoms as isotropic units, molecular polarizabilities account for the distribution of charge and bonding effects that lead to tensorial behavior. Additive models approximate the total molecular polarizability as the sum of contributions plus corrections for bonds and interactions, providing a computationally efficient way to estimate values for larger systems. One widely used approach is the Applequist model, which employs a point dipole interaction framework where atomic polarizabilities interact via dipole-dipole couplings, particularly effective for conjugated systems like polyenes. In many molecules, polarizability is anisotropic due to , quantified by the difference Δα = α_∥ - α_⊥, where α_∥ and α_⊥ are the components parallel and perpendicular to the principal axis. This anisotropy arises from the elongated along bonds, making linear molecules like exhibit significant values, with Δα ≈ 0.5 ų, reflecting the stronger response along the molecular axis. Quantum mechanical calculations, such as (DFT) with time-dependent extensions, offer high accuracy for these tensors, typically within 5-10% of experimental results, by solving the perturbed . Representative examples illustrate these effects: the water molecule has an isotropic average polarizability of 1.44 ų, enhanced by the lone pairs on oxygen that facilitate electron distortion. In , the delocalized π electrons and aromatic ring currents substantially increase the in-plane polarizability components, yielding a mean value of approximately 10 ų, far exceeding simple additive atomic sums. For nonlinear optical responses, the first hyperpolarizability β extends this to second-order effects, where the induced depends quadratically on the field, relevant for molecules with broken inversion symmetry.

Applications of Polarizability

In Condensed Matter and

In , polarizability plays a crucial role in understanding the response of bulk materials, where local field corrections account for the influence of surrounding dipoles on the effective experienced by individual atoms or molecules. The Clausius-Mossotti provides a fundamental link between the microscopic polarizability \alpha and the macroscopic \epsilon_r, given by \alpha = 3 \epsilon_0 \frac{\epsilon_r - 1}{N (\epsilon_r + 2)}, where \epsilon_0 is the and N is the of polarizable units. This arises from considering the local field as the applied field plus the field from a spherical in the polarized medium, enabling predictions of bulk properties from atomic-scale responses. In crystalline solids, polarizability manifests as a second-rank tensor due to the lack of in non-cubic , leading to direction-dependent responses that reflect the crystal's . For instance, in \alpha- (SiO_2), the polarizability tensor exhibits principal components that differ by approximately 8%, contributing to its overall anisotropic behavior. This tensorial nature influences properties such as and wave propagation, with calculations often involving dipolar sums to relate atomic polarizabilities of Si and O to the observed refractivities. Polarizability also impacts by affecting electron density distributions and atomic factors, particularly through the , which treats as interaction with a potential derived from the electron cloud. In advanced refinement methods, polarizable atomic multipole models, such as those based on the , incorporate induced dipoles to better describe aspherical electron densities, improving agreement with experimental data; for example, in crystals, such models reduce the R-free factor by 0.45% compared to independent atom models. These approaches enhance the accuracy of structure determination by accounting for polarization effects in the amplitudes. The distinction between ionic and covalent solids highlights polarizability's dependence on electron delocalization: covalent solids exhibit higher values due to shared, mobile s that enhance response to fields, whereas ionic solids rely on localized charges with lower polarizability. In , a prototypical covalent solid, the effective atomic polarizability per carbon atom is approximately 0.85 ų, reflecting the extended sp³ bonding network. This contrasts with ionic solids like NaCl, where per-formula-unit polarizability is smaller owing to rigid positions. Modern applications leverage polarizability gradients—spatial variations in the tensor components across the —to predict piezoelectric responses, particularly in molecular and where in distribution induces stress-dependent . Seminal studies have established linear correlations between piezoelectric coefficients and molecular polarizabilities, guiding the of high-performance materials for sensors and actuators.

In Optics and

Polarizability at optical frequencies, denoted as \alpha(\omega), exhibits a strong dependence on the frequency \omega of the incident light, governed by the Kramers-Kronig relations that connect the real and imaginary parts of the complex polarizability through causality principles. These relations arise from the analyticity of the response function in the complex frequency plane, ensuring that the dispersive (real) part reflects the absorptive (imaginary) behavior across all frequencies. Near electronic resonances, such as UV absorption lines, \alpha(\omega) peaks dramatically, enhancing light-matter interactions; for instance, in alkali metals like sodium and potassium, the dynamic polarizability shows pronounced resonances around their atomic transition wavelengths in the UV range, leading to anomalous dispersion. In dilute gases, this frequency-dependent polarizability directly influences optical , manifesting in the n(\omega). For low densities where effects are negligible, the Clausius-Mossotti approximation simplifies to n(\omega) \approx 1 + \frac{N \alpha(\omega)}{2\epsilon_0}, where N is the and \epsilon_0 is the . This relation is empirically captured by the , which models n(\omega) using oscillator strengths tuned to features, enabling precise predictions of in gases like air or over visible and near-IR wavelengths. Such is critical for applications in precision and pulse propagation. At higher light intensities, nonlinear polarizability effects emerge, extending the linear response to higher-order terms in the expansion \mathbf{P} = \epsilon_0 \left( \chi^{(1)} \mathbf{E} + \chi^{(2)} \mathbf{E}^2 + \chi^{(3)} \mathbf{E}^3 + \cdots \right), where \chi^{(3)} relates to the second hyperpolarizability \gamma at the molecular level. The term involving \gamma enables third-order processes like third-harmonic generation, where input light at \omega produces output at [3](/page/3)\omega, as the nonlinear oscillates at sum and difference of the driving field. This phenomenon is pivotal in conversion devices, with \gamma values enhanced in conjugated organic molecules or nanostructures for efficient . In Raman spectroscopy, polarizability plays a central role in inelastic light scattering, where vibrational modes modulate \alpha, leading to Stokes and anti-Stokes shifts. The change in polarizability with normal coordinate Q, \Delta \alpha = \left( \frac{\partial \alpha}{\partial Q} \right) \Delta Q, determines the scattering efficiency; specifically, the intensity of Raman lines is proportional to (\Delta \alpha)^2, making modes with large polarizability derivatives—such as symmetric stretches in symmetric molecules—prominent. This sensitivity allows Raman to probe molecular vibrations non-destructively, with depolarization ratios revealing tensor anisotropy in \alpha. Tuned polarizability underpins advanced optical applications, including photonic crystals and metamaterials, where engineered resonances in \alpha(\omega) yield effective permittivities \epsilon < 0, facilitating . In photonic crystals, periodic structures with high-index contrasts create bandgaps and flat equifrequency contours that bend light oppositely to , enabling superlensing beyond the diffraction limit. Metamaterials extend this by subwavelength resonators, such as split-ring arrays, whose electric polarizability mimics negative \epsilon, paired with magnetic responses for double-negative indices and applications in or .

Magnetic Polarizability

Definition and Basic Properties

Magnetic polarizability quantifies the linear response of a to an external , characterized by the of a moment \mathbf{m} proportional to the applied magnetic field strength \mathbf{H}, given by \mathbf{m} = \alpha_m \mathbf{H}, where \alpha_m is the magnetic polarizability. This property arises primarily from the perturbation of orbits and spins by the field, leading to circulating currents that generate the opposing or aligning moment. Unlike electric polarizability, which measures response to , magnetic polarizability reflects the material's tendency to modify the local magnetic environment through induced moments. The units of \alpha_m are cubic meters (m³), reflecting its volume-like nature for atomic-scale systems, in contrast to the units of electric polarizability \alpha_e (C² m² J⁻¹). For dilute systems or isolated atoms, \alpha_m relates to the volume magnetic susceptibility \chi_v via \chi_v = n \alpha_m, where n is the of particles. In isotropic systems, such as free atoms, \alpha_m is a scalar , assuming uniform response in all directions; however, in oriented or anisotropic systems like molecules or crystals, it becomes a second-rank tensor \boldsymbol{\alpha}_m, with components \alpha_{m,ij} describing direction-dependent . The physical basis of magnetic polarizability stems from the interaction of the external field with motion: orbital contributions induce diamagnetic currents via , where electrons orbit to oppose the field, while spin contributions can lead to paramagnetic alignment. For closed-shell atoms, the dominant effect is diamagnetic, with the Larmor theorem providing a classical approximation for the induced moment per \alpha_m \approx -\frac{\mu_0 e^2 r^2}{6 m_e}, where \mu_0 is the , e and m_e are the charge and , and r is a characteristic orbital radius. Representative examples include atoms like and , which exhibit small negative \alpha_m values on the order of -10^{-31} m³, reflecting their purely diamagnetic response due to filled shells.

Relation to Magnetic Susceptibility

In dilute systems, the magnetic susceptibility \chi_m is related to the magnetic polarizability \alpha_m by the expression \chi_m = N \alpha_m, where N is the of atoms or molecules. This relation connects the microscopic response of individual particles to the macroscopic M = \chi_m H, assuming the local approximates the applied and interactions between particles are negligible. Diamagnetism arises from a universal negative contribution to the magnetic polarizability, \alpha_{m,\mathrm{dia}} \approx -\frac{\mu_0 e^2 Z \langle r^2 \rangle}{6 m_e}, derived from the Larmor theorem, which accounts for the induced orbital currents opposing the applied field. This term reflects the quantum mechanical expectation value of r^2 over the electron distribution, leading to a weak, negative response inherent to all materials. In contrast, paramagnetism provides a positive contribution to \alpha_{m,\mathrm{para}} due to unpaired electron spins aligning with the field, following the Curie law \chi_m = C/T where the Curie constant C = N \mu^2 / 3k_B and \mu is the effective magnetic moment. This behavior dominates in materials with permanent magnetic moments, such as transition metal ions. The temperature dependence differs markedly: \alpha_m for paramagnets decreases inversely with T as thermal disorder randomizes spin orientations, while the diamagnetic \alpha_m remains nearly constant, independent of temperature. A striking example of extreme occurs in superconductors, where the expels magnetic fields completely below the critical temperature, yielding perfect diamagnetism with \chi_m = -1. In terms of polarizability, this corresponds to an effective response where N \alpha_m = -1.

References

  1. [1]
    The Feynman Lectures on Physics Vol. II Ch. 11: Inside Dielectrics
    The displacement of the electron distribution which produces this kind of induced dipole moment is called electronic polarization. We have already discussed the ...
  2. [2]
    [PDF] Discussion Class 7
    Oct 19, 2007 · The induced dipole moment of a neutral atom is approximately proportional to the external electric field: p = αE. The constant of ...
  3. [3]
    29. Molecular Polarizability and Electric Susceptibility
    Our simple model trivially gives the molecular polarization if the molecule is placed in a field of given strength. But real life is more complicated.
  4. [4]
  5. [5]
    [PDF] Chemical Science - University of Minnesota Twin Cities
    Polarizability is a key molecular property controlling induction and dispersion forces in molecules, and atomic polarizabilities in molecules are widely ...Missing: definition | Show results with:definition
  6. [6]
    London Dispersion Forces
    The ease with which the electron distribution around an atom or molecule can be distorted is called the polarizability. London dispersion forces tend to be:.
  7. [7]
    Development of Polarizable Models for Molecular Mechanical ... - NIH
    Mar 10, 2011 · The induced dipole of an atom due to other point charges and other point dipoles are computed by the i_RESP program. Consistent with the ...
  8. [8]
    polarizability (P04711) - IUPAC Gold Book
    Polarizability is the ease of distortion of a molecule's electron cloud by an electric field, measured as the ratio of induced dipole moment to the field.
  9. [9]
  10. [10]
    VIII. A dynamical theory of the electromagnetic field
    Oct 27, 2025 · (1) The most obvious mechanical phenomenon in electrical and magnetical experiments is the mutual action by which bodies in certain states ...
  11. [11]
    [PDF] On the Clausius-Mossotti Relation - Kirk T. McDonald
    May 20, 2025 · In this note we review the derivation of the Clausius-Mossotti relation [1, 3] between the molecular polarizability α and the relative ...
  12. [12]
    Peter J. W. Debye – a whole life devoted to science - ResearchGate
    Aug 7, 2025 · This paper re-examines briefly Debye's works on the origin and ... Debye (1912) at the age of 28 (Authier, 2013). …
  13. [13]
    Electronic Polarizabilities of Ions from the Hartree-Fock Wave ...
    Electronic Polarizabilities of Ions from the Hartree-Fock Wave Functions. R. M. Sternheimer. Brookhaven National Laboratory, Upton, New York. PDF Share. X ...
  14. [14]
    COMPENDIUM OF AB INITIO CALCULATIONS OF MOLECULAR ...
    The number of ab initio molecular electronic calculations has increased dramatically in the last few years. Both the practitioners and other interested ...
  15. [15]
  16. [16]
    [PDF] Infrared Refractive Index and Thermo-optic Coefficient Measurement ...
    REFRACTIVE INDEX. The refractive index arises from the molecular polarizability a according to the Lorentz–Lorenz formula: n n. T. T. 2. 2. 0. 1. 2. 3. −. +. =.
  17. [17]
    Electronic polarizability of light crude oil from optical and dielectric ...
    Molar refractive index and the molar volume are evaluated through Lorentz-Lorenz equation. The function of the refractive index, FRI, divided by the mass ...
  18. [18]
    Alkali Polarizabilities by the Atomic Beam Electrostatic Deflection ...
    The electric dipole polarizabilities of the alkali atoms have been measured by observing the deflection of a collimated beam of neutral atoms in a ...
  19. [19]
    Phys. Rev. A 10, 1141 (1974) - Measurement of alkali-metal ...
    Oct 1, 1974 · The electric dipole polarizability of the alkali-metal atoms Na, K, Rb, and Cs are determined by measuring the deflection of a velocity-selected ...
  20. [20]
    [PDF] Atom interferometry measurement of the electric polarizability of lithium
    Dec 7, 2005 · For alkali atoms, all the accurate experiments were based on the deflection of an atomic beam by an inhomogeneous electric field and, in the ...
  21. [21]
    Measurement of the Stark shift of the transitions in atomic cesium
    In this report, we discuss our recent measurements of the scalar static polarizabilities α 0 of the 7 p 2 P 1 / 2 and 7 p 2 P 3 / 2 states of atomic cesium and ...<|separator|>
  22. [22]
    Measurement of dynamic Stark polarizabilities by analyzing spectral ...
    Dec 8, 2010 · The dc Stark shift is given by E dc ( ℓ ) = - ( 1 / 2 ) α ℓ dc E dc 2 , where E dc = | E dc | is the magnitude of the dc electric field, and α ...
  23. [23]
    [PDF] Cavity ring-down spectrometer for high-fidelity molecular absorption ...
    We present a cavity ring-down spectrometer which was developed for near-infrared measurements of laser absorption by atmospheric greenhouse gases.Missing: polarizability | Show results with:polarizability
  24. [24]
    [PDF] Ellipsometry in the measurement of surfaces and thin films
    review, theory, computational techniques, measurement techniques, and the use of ellipsometry in measuring metal surface oxide films and organic films. Library ...
  25. [25]
    [PDF] 2023 Table of Polarizabilities
    Jul 17, 2023 · Figure 1. Recommended values from Table 1 for the atomic polarizabilities (atomic units; estimated uncertainties in parentheses) of elements Z ...
  26. [26]
    2018 Table of static dipole polarizabilities of the neutral elements in ...
    In very nearly all cases, polarizabilities decrease with increase in atomic number for each nl group of elements (e.g., the 2p elements B-Ne).
  27. [27]
  28. [28]
    The quantum theory of atomic polarization I - Journals
    Though the perturbation method, strictly applied, is the more accurate, since it takes into account the possible excited states of the system, its usefulness is ...
  29. [29]
    Variational Approach to Perturbation Theory. I. Application to the ...
    I. Application to the Calculation of Atomic Polarizabilities. A variational approach to perturbation theory is derived which establishes the relationship ...
  30. [30]
    Additivity methods in molecular polarizability - ACS Publications
    Derivation of Distributed Models of Atomic Polarizability for Molecular Simulations. Journal of Chemical Theory and Computation 2007, 3 (6) , 1901-1913 ...
  31. [31]
    Measurement of the electric polarizability of sodium with an atom ...
    Aug 6, 2025 · Using laser-atomic-beam spectroscopy and a special field arrangement the Stark shift of the sodium D1-line was determined to be 48.986 (112) ...
  32. [32]
    Static and dynamic polarizability for C 2+ in Rydberg states
    This work presents results from a non-perturbative calculation of dynamic polarizability of C III ions in 1s22sns (1Se) Rydberg states.
  33. [33]
    [PDF] Static and dynamic dipole polarizability of the helium atom using ...
    Jul 23, 2003 · We present a calculation of the static and dynamic dipole polarizability of the helium atom using a varia- tionally stable treatment that ...
  34. [34]
    Additive models for the molecular polarizability and volume - ADS
    Additive models for molecular polarizabilities and volumes are created by fitting to data for 298 molecules. Tests on data for the 1641 organic molecules in ...Missing: Applequist | Show results with:Applequist
  35. [35]
    Atom dipole interaction model for molecular polarizability ...
    Refinement of atomic polarizabilities for a polarizable Gaussian multipole force field with simultaneous considerations of both molecular polarizability ...
  36. [36]
    Measuring polarizability anisotropies of rare gas diatomic molecules ...
    Jun 6, 2011 · The polarizability anisotropies of homonuclear rare gas diatomic molecules, Ar2, Kr2, and Xe2, are investigated by utilizing the interaction ...
  37. [37]
    Density functional calculations of molecular polarizabilities and ...
    The corresponding average error is 5.5% for the Sadlej basis set and 2.6% for the TZVP-FIP1 basis set. This is an indication of the importance of a DFT ...
  38. [38]
    A systematic development of a polarizable potential of water
    May 31, 2013 · The polarizability of the gas-phase molecule is close to isotropic, the values of its components are60 αxx = 1.4146 Å3, αyy = 1.5284 Å3, and ...
  39. [39]
    Experimental data for C 6 H 6 (Benzene) - CCCBDB
    Calculated electric dipole polarizability for C6H6 (Benzene). References ... molecular polarizability. An analysis of the experimental values" J. Mol ...
  40. [40]
    Hyperpolarizability - an overview | ScienceDirect Topics
    According to the theory, the molecular second- order hyperpolarizability can be decomposed into two parts. [25.62] β = β add + β ct. where βadd is an additive ...
  41. [41]
    Clausius-Mossotti Relation - Richard Fitzpatrick
    $$ \alpha$ is called the molecular polarizability. If $ N$ is the number density of such molecules then the polarization of the medium is. $\displaystyle {\bf ...
  42. [42]
    The electric field in crystals III. The refractivity of α-quartz - IOPscience
    A calculation has been performed by using dipolar lattice sums in α-quartz to interpret the refractivities in terms of the point polarizabilities of silicon and ...
  43. [43]
    30 The Internal Geometry of Crystals - Feynman Lectures - Caltech
    For instance, a crystal will, in general, have a tensor electric polarizability. If we describe the tensor in terms of the ellipsoid of polarization, we should ...
  44. [44]
    Polarizable Atomic Multipole X-Ray Refinement - PubMed Central
    We recently developed a polarizable atomic multipole refinement method assisted by the AMOEBA force field for macromolecular crystallography.
  45. [45]
    Ab initio simulations of color centers in diamond - ScienceDirect.com
    This chapter covers ab initio modeling of color centers in diamonds, including atomic structure, defect formation energy, charge states, and different color ...
  46. [46]
    [PDF] Optical Properties of Solids Over a Wide Frequency Range
    The Kramers–Kronig relations relate ε1(ω) and ε2(ω) so that if either of these functions is known as a function of ω the other is completely determined. ...
  47. [47]
    Frequency-dependent polarizabilities of alkali-metal atoms from ...
    Feb 9, 2006 · We present results of first-principles calculations of the frequency-dependent polarizabilities of all alkali-metal atoms for light in the wavelength range 300 ...Missing: absorption | Show results with:absorption
  48. [48]
    polarizability - The Nonlinear Optics Home Page
    where p0 is the dipole moment with no field applied, α is called the polarizability, β the hyperpolarizability, γ the second hyperpolarizability, etc.
  49. [49]
    Negative refractive index metamaterials - ScienceDirect.com
    Metamaterials that display negative refractive index – a property not found in any known naturally occurring material – have drawn significant scientific ...
  50. [50]
    [PDF] A rigorous derivation of the Larmor and Van Vleck contributions. - HAL
    Apr 24, 2014 · Abstract. The purpose of this paper is to rigorously investigate the orbital magnetism of core elec- trons in 3-dimensional crystalline ...
  51. [51]
    [PDF] 3 Molecules in Electric and Magnetic Fields
    The magnetic polarizability, defined in analogy to electric polarizability, cf. equ. (3.06), is mind = ξ B. (3.55) with the magnetizability ξ (zeta) ...
  52. [52]
    Explorations of Magnetic Properties of Noble Gases - MDPI
    The calculations of diamagnetic effects, also known as shielding effects present in any atom or molecule, are crucial for precise calculations. In this case ...
  53. [53]
    [PDF] arXiv:1803.09555v1 [physics.optics] 26 Mar 2018
    Mar 26, 2018 · The contribution of the magnetic polarizability is in general neglected because it is small, and the magnetic susceptibility is set equal to ...
  54. [54]
    [PDF] classical Langevin theory of diamagnetism
    Classical Langevin theory describes diamagnetism where electron magnetic moments develop opposite to an applied field, and the susceptibility is negative and ...Missing: formula polarizability
  55. [55]
    [PDF] Physics of Modern Materials Chapter 8: Magnetic Materials
    Experimentally, diamagnetism can be observed by placing a diamagnetic material in a strong magnetic field gradient. The material will experience a force toward ...Missing: formula polarizability
  56. [56]
    Tutorial: a beginner's guide to interpreting magnetic susceptibility ...
    Apr 19, 2022 · In this tutorial, we provide a guide to the interpretation of magnetic susceptibility data with a special emphasis on the Curie–Weiss law, a simple but ...Missing: polarizability | Show results with:polarizability
  57. [57]
    [PDF] Magnetic Susceptibility of Superconductors
    Meissner effect, magnetic units, and formula conversions are discussed. There is a comprehensive summary of critical-state formulas for slabs and cylinders ...Missing: polarizability | Show results with:polarizability