Fact-checked by Grok 2 weeks ago

Polarization density

Polarization density, denoted as the vector field \mathbf{P}, is a fundamental concept in representing the per unit volume within a material exposed to an . It quantifies the extent to which the material's atoms or molecules become polarized, either through the alignment of permanent dipoles or the induction of temporary dipoles, resulting in a net separation of positive and negative charges. Mathematically, \mathbf{P} is expressed as \mathbf{P} = N \mathbf{p}, where N is the of polarizable entities and \mathbf{p} is the average per entity, with units of coulombs per square meter in the system. This polarization leads to the formation of bound charges, which are charges tied to the material's structure and do not move freely, distinguishing them from free charges. The volume density of bound charge is given by \rho_b = -\nabla \cdot \mathbf{P}, arising from spatial variations in \mathbf{P}, while surface bound charge density appears at interfaces as \sigma_b = \mathbf{P} \cdot \hat{\mathbf{n}}, where \hat{\mathbf{n}} is the outward normal vector. These bound charges modify the total electric field inside the material, typically reducing it compared to the applied field in linear dielectrics. In the framework of , polarization density is incorporated through the \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P}, where \epsilon_0 is the and \mathbf{E} is the . This leads to in the form \nabla \cdot \mathbf{D} = \rho_f, where \rho_f is the free , simplifying calculations by isolating the effects of free charges from bound ones. For linear isotropic dielectrics, \mathbf{P} = \epsilon_0 \chi_e \mathbf{E}, with \chi_e as the , yielding \mathbf{D} = \epsilon \mathbf{E} where \epsilon = \epsilon_0 (1 + \chi_e) is the . Polarization density plays a crucial role in applications such as capacitors, where it enhances by a factor of the , and in devices like microphones that exploit permanent polarization. It also influences wave propagation in materials, affecting phenomena like and the behavior of in dielectrics. Understanding \mathbf{P} is essential for modeling the response of materials to in fields ranging from to .

Fundamentals

Definition

Polarization density, denoted as the \vec{P}, quantifies the per unit volume within a material exposed to an . The \vec{p} of a localized distribution of charges is defined as \vec{p} = \sum_i q_i \vec{r}_i, where q_i are the charges and \vec{r}_i are their position vectors relative to a chosen origin within the distribution. In s, an applied causes a slight separation of positive and negative charges in atoms or molecules, inducing such dipole moments that align predominantly with the field. Macroscopically, \vec{P}(\vec{r}) at a point \vec{r} is obtained through spatial averaging of the microscopic contributions over a small volume V surrounding \vec{r}, where V is large enough to encompass many dipoles but small compared to variations in the field: \vec{P}(\vec{r}) = \frac{1}{V} \int_V \vec{p}(\vec{r}') \, dV', with \vec{p}(\vec{r}') representing the local density from all charges in the infinitesimal element dV'. This averaging smooths out microscopic fluctuations to yield a continuous describing the material's overall response. The units of \vec{P} are coulombs per square meter (C/m²), reflecting its equivalence to a surface charge density.

Microscopic Interpretation

Polarization density originates at the and molecular scales through the formation of within materials exposed to an external . In , this begins with the displacement of clouds relative to the positively charged nuclei, creating induced dipoles that contribute to the overall . This displacement is a universal mechanism in insulators, where the applied field distorts the symmetric charge distribution, leading to a net per atom. Several distinct mechanisms drive these microscopic dipoles, categorized by the nature of the material and the response to the field. Electronic polarization involves the aforementioned shift of clouds and occurs in all dielectrics, dominating at high frequencies. , or ionic, polarization arises in ionic crystals through the relative displacement of oppositely charged ions in the , such as in (NaCl), where Na⁺ and Cl⁻ ions shift slightly under the field, enhancing the dipole density. Orientational polarization occurs in materials with permanent molecular dipoles, like (H₂O), where the asymmetric charge distribution allows the molecules to rotate and align with the field, overcoming thermal randomization. Interfacial polarization, also known as space-charge polarization, emerges at boundaries or heterostructures due to the accumulation of mobile charges, creating effective dipoles at interfaces rather than within bulk atoms. The total polarization density combines contributions from induced dipoles, which form solely due to the external field distorting charge distributions (as in and mechanisms), and permanent dipoles, which exist intrinsically but require field-induced (as in orientational ). Induced dipoles are temporary and proportional to the field strength, while permanent ones rely on structure for their baseline moment. In , for instance, the permanent of each H₂O (arising from oxygen's ) aligns to produce significant orientational effects at low frequencies, whereas NaCl's response is primarily from induced ionic shifts without inherent dipoles. The microscopic view transitions to the macroscopic polarization density through statistical averaging over large ensembles of these dipoles, particularly in non-uniform fields where local variations must be smoothed to yield bulk properties. This averaging accounts for and spatial inhomogeneities, defining the observable as an ensemble average rather than individual moments. Such an approach highlights limitations in directly equating microscopic behaviors to macroscopic observables without considering these statistical effects.

Mathematical Formulations

Equivalent Expressions

One equivalent mathematical representation of the polarization density arises from microscopic charge distributions in dielectrics. For a localized, neutral charge configuration, the local polarization at position \mathbf{r} can be expressed as an integral over the \rho(\mathbf{r}') within a small volume surrounding \mathbf{r}: \mathbf{P}(\mathbf{r}) = \int \left( \mathbf{r}' - \mathbf{r} \right) \rho(\mathbf{r}') \, dV', where the integral captures the effective dipole contribution from the displaced charges, assuming higher-order multipoles are negligible. This form is equivalent to the standard dipole-per-volume definition under the dipole approximation and is particularly useful for deriving macroscopic properties from atomic-scale charge arrangements./03%3A_Polarization_and_Conduction/3.01%3A_Polarization) An alternative integral expression emphasizes the dynamic origin of polarization through adiabatic charge transport. In this view, the polarization density is given by the space-time integral of the current density \mathbf{j}(\mathbf{r}, \lambda) over an adiabatic parameter \lambda (typically from 0 to 1, representing a transition between reference states): \mathbf{P}(\mathbf{r}) = \int_0^1 \mathbf{j}(\mathbf{r}, \lambda) \, d\lambda. This formulation, rooted in classical electrodynamics, treats polarization as the accumulated charge flow during a reversible deformation of the charge distribution and holds equivalence to static dipole models for slow, quasistatic processes. It clarifies ambiguities in finite systems and extends naturally to non-periodic structures. In linear media, a differential form expresses the polarization density locally through the electric susceptibility \chi, where \mathbf{P}(\mathbf{r}) is proportional to the susceptibility at each point, reflecting spatial variations without explicit field dependence in the expression itself. This local, point-wise relation assumes isotropy and linearity, providing a continuum description suitable for inhomogeneous materials. For periodic structures, such as , Fourier transform representations facilitate computation via Bloch wavefunctions. The electronic contribution to is obtained by integrating the Berry connection over the : P_{el} = \frac{e}{(2\pi)^3} \mathrm{Im} \sum_n \int_{BZ} d\mathbf{k} \, \langle u_{n\mathbf{k}} | \nabla_{\mathbf{k}} u_{n\mathbf{k}} \rangle, where u_{n\mathbf{k}} are periodic parts of Bloch functions, summed over occupied bands n, and the total includes ionic terms. This k-space integral yields a multi-valued modulo lattice vectors, essential for quantifying spontaneous in ferroelectrics. In computational simulations, polarization density is modeled either discretely or continuously. Discrete approaches assign inducible point dipoles to atomic sites, computing \mathbf{P} as a sum over these moments divided by local volumes, which is efficient for molecular dynamics but approximates distributed charge effects. Continuous models, conversely, derive \mathbf{P} from integrated electron or force densities across the simulation cell, offering higher fidelity for delocalized systems like solids, though at greater computational cost. These methods converge under refined grids, with discrete variants suiting large-scale polarizable force fields.

Associated Bound Charges

In dielectrics, spatial variations in the density \mathbf{P} give rise to bound charges, which are immobile charges resulting from the of electrons relative to positive nuclei within the . These bound charges effectively act as sources of the produced by the polarization itself, distinct from free charges that can move through the . The volume bound charge density \rho_b is defined as \rho_b = -\nabla \cdot \mathbf{P}, where the negative sign arises because a positive of \mathbf{P} corresponds to a net outflow of , leaving an effective negative . To derive this, consider the \phi_P generated by the dipole distribution equivalent to \mathbf{P}: \phi_P(\mathbf{r}) = \frac{1}{4\pi\epsilon_0} \int_V \frac{\mathbf{P}(\mathbf{r}') \cdot (\mathbf{r} - \mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|^3} \, dV'. This integral can be transformed using the identity \mathbf{P} \cdot \nabla' (1/|\mathbf{r} - \mathbf{r}'|) = \nabla' \cdot (\mathbf{P}/|\mathbf{r} - \mathbf{r}'|) - (\nabla' \cdot \mathbf{P})/|\mathbf{r} - \mathbf{r}'| and , yielding a volume term \frac{1}{4\pi\epsilon_0} \int_V \frac{-\nabla' \cdot \mathbf{P}(\mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|} \, dV' plus surface contributions. The volume term matches the potential from a \rho_b = -\nabla \cdot \mathbf{P}, so \nabla^2 \phi_P = -\rho_b / \epsilon_0. The associated \mathbf{E}_P = -\nabla \phi_P then satisfies in the form \nabla \cdot \mathbf{E}_P = \rho_b / \epsilon_0. At the surface of the , a surface bound \sigma_b forms due to the abrupt termination of \mathbf{P}, given by \sigma_b = \mathbf{P} \cdot \hat{\mathbf{n}}, where \hat{\mathbf{n}} is the outward-pointing unit normal vector. This term originates from the surface integral in the of the potential derivation, representing the component of \mathbf{P} normal to the . Physically, the divergence \nabla \cdot \mathbf{P} quantifies a local imbalance in the density of aligned : a positive divergence implies more leaving a volume than entering, equivalent to a net negative bound charge accumulation, while a negative divergence indicates a net positive bound charge accumulation. This imbalance mimics the effect of separated positive and negative charges without actual charge transfer. For uniform \mathbf{P}, \nabla \cdot \mathbf{P} = 0, so \rho_b = 0 everywhere, but \sigma_b = P on one surface and \sigma_b = -P on the opposite surface, as seen in a uniformly polarized slab or . In contrast, non-uniform \mathbf{P} in electrets—materials exhibiting quasi-permanent due to frozen alignments—produces both volume and surface bound charges, resulting in internal that persist over time.

Field Relationships

Isotropic Dielectrics

In isotropic dielectrics, the polarization density \mathbf{P} exhibits a linear and direction-independent response to the applied \mathbf{E}, characteristic of homogeneous materials where properties are uniform in all directions. This scalar relationship arises in media such as gases, liquids, and amorphous solids, where microscopic dipoles align without preferred orientations. The constitutive relation for such materials is given by \mathbf{P} = \epsilon_0 \chi \mathbf{E}, where \epsilon_0 is the and \chi is the , a scalar quantity that quantifies the material's ability to become polarized. This equation assumes a linear response, valid for sufficiently weak fields where higher-order effects are negligible. This macroscopic relation derives from microscopic averaging in uniform media, where the polarization density \mathbf{P} is the volume average of individual dipole moments \mathbf{p}_i from atoms or molecules: \mathbf{P} = n \langle \mathbf{p} \rangle, with n as the number density of polarizable units and \langle \mathbf{p} \rangle the ensemble average dipole moment induced by the local field. In isotropic cases, each dipole responds linearly as \mathbf{p} = \alpha \mathbf{E}_\text{loc}, with \alpha the atomic polarizability, leading to \chi = n \alpha / \epsilon_0 after averaging over a volume much larger than molecular scales but smaller than macroscopic variations. The permittivity \epsilon of the material connects directly to this relation via \epsilon = \epsilon_0 (1 + \chi), where the dielectric constant \kappa = 1 + \chi measures the enhancement of capacitance relative to vacuum. For example, distilled water has \kappa \approx 80 at room temperature, implying a high \chi \approx 79, while dry air has \kappa \approx 1.0006 with \chi \approx 0.0006, reflecting weak polarization in dilute gases. In practice, \chi exhibits frequency dependence due to resonant responses of bound charges, causing dispersion where the effective permittivity varies with the field's oscillation rate, though the linear form holds quasi-statically for non-resonant frequencies. This model is limited to linear, homogeneous conditions, failing in nonlinear regimes at strong fields or in inhomogeneous media where spatial variations disrupt uniform averaging.

Anisotropic Dielectrics

In anisotropic dielectrics, the polarization density \mathbf{P} does not align uniformly with the \mathbf{E} as in isotropic materials, where a scalar \chi yields \mathbf{P} = \epsilon_0 \chi \mathbf{E}; instead, the response depends on the direction of \mathbf{E} relative to the material's , leading to direction-dependent polarization. This anisotropy arises from the lack of spherical in the atomic or molecular arrangement, requiring a tensor description to capture the varying along different axes. The relationship between polarization and electric field is expressed using the electric susceptibility tensor \boldsymbol{\chi}, a 3×3 matrix, as P_i = \epsilon_0 \chi_{ij} E_j, where summation over repeated indices j is implied, \epsilon_0 is the vacuum permittivity, and the Einstein summation convention applies. The tensor \boldsymbol{\chi} is symmetric (\chi_{ij} = \chi_{ji}) due to the real nature of the dielectric response in non-absorbing media, and its components determine how \mathbf{P} components arise from \mathbf{E}. The corresponding relative permittivity tensor is then \epsilon_{ij}/\epsilon_0 = \delta_{ij} + \chi_{ij}, where \delta_{ij} is the Kronecker delta, linking the displacement field \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P} to \mathbf{E} via D_i = \epsilon_{ij} E_j. For practical analysis, the susceptibility tensor can be diagonalized by choosing coordinates aligned with the material's principal axes, simplifying \boldsymbol{\chi} to a diagonal form with elements \chi_1, \chi_2, \chi_3 along these axes. In uniaxial crystals, two principal susceptibilities are equal (\chi_1 = \chi_2 \neq \chi_3), corresponding to a single optic axis, while biaxial crystals have three distinct values (\chi_1 \neq \chi_2 \neq \chi_3), with two optic axes. A representative example is calcite (\ce{CaCO3}), a uniaxial crystal where the differing principal permittivities cause double refraction: an incident light ray splits into ordinary and extraordinary rays with orthogonal polarizations, experiencing refractive indices n_o \approx 1.658 and n_e \approx 1.486 at visible wavelengths, respectively. Techniques like measure the tensor components by analyzing changes in light upon or from the material surface, providing insights into the anisotropic response without requiring invasive probes. This direction-dependent fundamentally distinguishes anisotropic dielectrics from isotropic ones, enabling applications in waveguiding, polarization control, and where field orientation critically influences material behavior.

Role in Electrodynamics

Static Maxwell's Equations

In the static regime of electrodynamics, the polarization density \mathbf{P} enters Maxwell's equations through the introduction of the electric displacement field \mathbf{D}, defined as \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P}, where \epsilon_0 is the permittivity of free space and \mathbf{E} is the electric field strength. This definition encapsulates the contribution of both the vacuum response and the material's polarization to the total displacement. The static Maxwell's equations in matter then take the form \nabla \cdot \mathbf{D} = \rho_f, \quad \nabla \times \mathbf{E} = 0, \quad \nabla \cdot \mathbf{B} = 0, \quad \nabla \times \mathbf{H} = \mathbf{J}_f, where \rho_f and \mathbf{J}_f are the free charge density and free current density, respectively, \mathbf{B} is the magnetic flux density, and \mathbf{H} is the magnetic field strength. The first equation, a modified form of Gauss's law, isolates the effects of free charges by incorporating the bound charges arising from \mathbf{P} (with volume bound charge density \rho_b = -\nabla \cdot \mathbf{P}) into \mathbf{D}, while the remaining equations are unchanged from their vacuum counterparts, reflecting the absence of time-varying fields. At interfaces between different , boundary conditions derived from these equations ensure consistent with the physics of charges and fields. The normal component of \mathbf{D} is discontinuous by the \sigma_f: \hat{n} \cdot (\mathbf{D}_2 - \mathbf{D}_1) = \sigma_f, where \hat{n} is pointing from medium 1 to 2; if \sigma_f = 0, the \mathbf{D} is continuous. The tangential component of \mathbf{E} remains continuous: \hat{n} \times (\mathbf{E}_2 - \mathbf{E}_1) = 0. For \mathbf{P}, the discontinuity in its component corresponds to the bound surface \sigma_b: \hat{n} \cdot (\mathbf{P}_2 - \mathbf{P}_1) = \sigma_b. These conditions arise from applying the integral forms of and Faraday's law to infinitesimal pillboxes and loops straddling the interface. This formulation highlights the role of \mathbf{P} in describing the material's dielectric response to applied fields, permitting the equations to depend solely on externally controllable free charges and currents rather than the induced bound charges.

Time-Varying Fields

In time-varying electromagnetic fields, the polarization density \mathbf{P} becomes time-dependent, introducing dynamic effects that modify to account for the motion of bound charges within s. Specifically, in the Ampère-Maxwell law, the curl of the magnetic field \mathbf{B} relates to the free current density \mathbf{J}_f and the time derivative of the \mathbf{D}: \nabla \times \mathbf{B} = \mu_0 \left( \mathbf{J}_f + \frac{\partial \mathbf{D}}{\partial t} \right), where \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P}, so \frac{\partial \mathbf{D}}{\partial t} = \epsilon_0 \frac{\partial \mathbf{E}}{\partial t} + \frac{\partial \mathbf{P}}{\partial t}. The term \frac{\partial \mathbf{P}}{\partial t} represents the contribution from the changing polarization, effectively acting as a bound current that influences magnetic field generation alongside free currents. This time-dependent polarization gives rise to the polarization current density \mathbf{J}_p = \frac{\partial \mathbf{P}}{\partial t}, which describes the flow of bound charges due to the reorientation or displacement of dipoles in response to the varying \mathbf{E}. In the Ampère-Maxwell law, \mathbf{J}_f + \frac{\partial \mathbf{D}}{\partial t} can thus be rewritten as \mathbf{J}_f + \mathbf{J}_p + \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}, highlighting how \mathbf{J}_p supplements the vacuum displacement current. Meanwhile, Faraday's law, \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, remains formally unchanged by polarization, but the induced electric field it produces can alter \mathbf{P} through mechanisms in materials, coupling the electric and magnetic dynamics. For alternating current (AC) fields, analysis often shifts to the frequency domain, where time-harmonic fields are assumed with angular frequency \omega. Here, the polarization responds via a frequency-dependent electric susceptibility \chi(\omega), leading to a complex permittivity \epsilon(\omega) = \epsilon_0 [1 + \chi(\omega)]. The imaginary part of \epsilon(\omega) accounts for energy dissipation, while the real part governs energy storage, enabling compact descriptions of wave propagation and material response in dynamic scenarios. A practical example occurs in capacitors with dielectric fillers under AC voltage, where the lagging response of \frac{\partial \mathbf{P}}{\partial t} behind \mathbf{E}—due to finite reorientation times—results in losses, manifesting as heat from the phase difference between and voltage. This lag, quantified by the loss tangent \tan \delta = \frac{\epsilon''(\omega)}{\epsilon'(\omega)} where \epsilon = \epsilon' - i \epsilon'', reduces efficiency but is essential for understanding power dissipation in devices like electrolytic capacitors.

Special Considerations

Crystalline Materials

In crystalline materials, the polarization density \mathbf{P} arises from the ordered arrangement of charges within the periodic lattice, but its definition encounters fundamental ambiguities due to the inherent in descriptions. The modern theory of addresses the electronic contribution in insulators through the Berry phase approach, which computes changes in \mathbf{P} as the system evolves adiabatically under variations in the , resolving issues with the multi-valued nature of polarization in periodic systems. This formulation, developed in the early 1990s, expresses \mathbf{P} in terms of the geometric phase accumulated by Bloch wavefunctions over the , providing a gauge-invariant measure for insulating crystals. A key ambiguity in defining \mathbf{P} stems from the , where the is only determined up to multiples of the vectors, reflecting the indeterminacy in assigning charge ownership across boundaries in an infinite crystal. This "quantum of polarization" implies that absolute values of \mathbf{P} cannot be uniquely fixed without additional conventions, such as choosing a specific branch of the . In ferroelectric crystals, this manifests practically; for instance, (BaTiO₃) exhibits spontaneous of approximately 0.26 C/m² in its tetragonal phase below the , arising from the displacement of Ti ions relative to the oxygen octahedra. The can be switched by external through the reorientation of ferroelectric domains, enabling applications in memory devices, though domain walls introduce additional complexities in achieving uniform switching. Measuring polarization density in crystals presents challenges, particularly in distinguishing intrinsic \mathbf{P} from extrinsic effects like domain contributions or surface charges. Piezoelectric effects provide an indirect link, as the converse piezoelectric response—strain induced by an electric field—couples to \mathbf{P} via the piezoelectric tensor, allowing local probing of polarization orientation through techniques like . However, in multi-domain crystals, averaging over domains complicates quantitative assessment, often requiring complementary methods such as to map domain-specific \mathbf{P}. Quantum mechanically, the underpinnings of \mathbf{P} in crystals rely on , which represent localized charge distributions obtained by Fourier transforming Bloch states, enabling the decomposition of electronic into contributions from Wannier centers within each . These functions resolve the delocalized nature of electrons in periodic potentials, allowing \mathbf{P} to be viewed as the sum over ionic and electronic Wannier charge densities, though ambiguities persist in maximally localizing them across the . This approach underpins computational predictions of \mathbf{P} in materials design, emphasizing the role of crystal symmetry in constraining possible states.

Amorphous Materials

In amorphous materials, the polarization density \mathbf{P} is defined as the statistical average of local electric dipole moments over a volume element, reflecting the random spatial arrangement and orientations of atoms or molecules without the periodic lattice structure found in crystals. This averaging process arises from the inherent structural disorder, where dipoles form due to local asymmetries in charge distribution, such as in polar organic solids or oxide glasses. Unlike crystalline counterparts, this approach lacks reliance on long-range order, making the macroscopic \mathbf{P} more straightforward to conceptualize as an ensemble property. Local variations in amorphous structures introduce ambiguities in \mathbf{P}, primarily through heterogeneous charge distributions that depend on the material's preparation history, such as cooling rates or deposition conditions, rather than global topological features like branch cuts in periodic systems. For instance, thermal aging in amorphous phase-change materials can alter alignments over time, leading to drifts in that complicate consistent measurements. These history-dependent effects stem from frozen-in defects or stress-induced distortions during solidification, resulting in non-uniform local fields that influence the overall response to applied . Representative examples include glasses, where trapped charges from manufacturing impurities contribute to persistent , and amorphous polymers like fluorinated variants used in electrets, which exhibit quasi-permanent \mathbf{P} due to deep traps in regions. In electrets, such as corona-charged amorphous fluoropolymers, space charges accumulate at interfaces or within bulk voids, stabilizing orientations for extended periods and enabling applications in or air filters. These materials demonstrate how facilitates charge trapping without the need for crystalline defects, though the randomness limits precise control over \mathbf{P}. Compared to crystalline materials, the macroscopic definition of \mathbf{P} in amorphous solids is simpler, as it avoids complications from lattice periodicity and allows direct volume averaging of dipoles, but microscopic modeling proves more challenging due to the need to account for extensive configurational variability and lack of . Computational simulations of amorphous dielectrics, for example, require large ensembles to capture statistical fluctuations, contrasting with the symmetry-reduced calculations feasible for . This trade-off highlights why amorphous systems are often characterized empirically rather than through derivations. Experimentally, effects in non-crystalline semiconductors, such as or organic photoconductors, manifest as interfacial that dominates low-frequency responses, arising from mobile carriers trapped at disorder-induced barriers. These effects are probed via techniques like thermally stimulated depolarization currents, revealing how accumulated charges near electrodes or grain boundaries enhance effective \mathbf{P} and contribute to in current-voltage characteristics. In such systems, polarization can exceed contributions, underscoring the role of structural inhomogeneities in practical device performance.

References

  1. [1]
    [PDF] Electromagnetic Fields and Energy - Chapter 6: Polarization
    That is, within such a polarizable body placed into an electric field, a polarization charge density is produced which, in turn, modifies the electric field. In ...
  2. [2]
    Polarization - Richard Fitzpatrick
    Polarization. The terrestrial environment is characterized by dielectric media (e.g., air and water) that are, for the most part, electrically neutral, because ...
  3. [3]
  4. [4]
    The Feynman Lectures on Physics Vol. II Ch. 10: Dielectrics - Caltech
    The surface density of charge is equal to the polarization inside the material. ... The answer has to do with the polarization of a dielectric when it is placed ...
  5. [5]
    Introducing polarization and magnetization into Maxwell's equations
    Jan 1, 2014 · Here ΔV denotes a small volume centered on the position r where the polarization is to be determined. The sum accounts for all dipole moments pi ...
  6. [6]
    6.1.1: Dielectric Polarization - Engineering LibreTexts
    Jul 28, 2022 · Interfacial or space charge polarization occurs when there is an accumulation of charge at an interface between two materials or between two ...Introduction · Ionic Polarization · Orientational Polarization · Interfacial Polarization
  7. [7]
    Polarisation mechanisms - DoITPoMS
    There are three main polarisation mechanisms that can occur within a dielectric material: electronic polarisation, ionic polarisation (sometimes referred to as ...
  8. [8]
    The Feynman Lectures on Physics Vol. II Ch. 11: Inside Dielectrics
    We will first discuss the polarization of nonpolar molecules. We can start with the simplest case of a monatomic gas (for instance, helium).
  9. [9]
    2 Electromagnetism: microscopic and macroscopic fields
    The macroscopic field (and other related quantities such as charge density ρ) must be the result of space-averaging (in some way) the microscopic field, with ...
  10. [10]
    [PDF] Theory of Polarization: A Modern Approach - Rutgers Physics
    In a crystalline system the CM macroscopic polarization PCM is defined as the sum of the dipole moments in a given cell divided by the cell volume. We shall ...
  11. [11]
    Polarization of arbitrary charge distributions: The classical ...
    Polarization can then be rigorously defined as the integral of the current density over that transition.
  12. [12]
    6.1 Polarization Density - MIT
    Also by definition, the net charge in V can be determined by integrating the polarization charge density over its volume.
  13. [13]
    Computing three-dimensional densities from force ... - AIP Publishing
    Aug 14, 2019 · This force density route to compute number, charge, and polarization densities is applied to a series of test cases in order to quantify the ...
  14. [14]
    [PDF] Electrostatics in material
    This polarization is parameterized by a polarization density P(r) giving the dipole moment per unit volume which is a function of position in the matter.
  15. [15]
    [PDF] DIELECTRIC POLARIZATION AND BOUND CHARGES - UT Physics
    Physically, we identify the σb = P·n as the net surface density of the bound charges and the ρb = −∇ · P as the net volume density of the bound charges.
  16. [16]
    6.4 Constitutive Laws of Polarization - MIT
    All isotropic media behave as linear media and obey (2) if the applied E field is sufficiently small. As long as E is small enough, any continuous function P(E ) ...
  17. [17]
    [PDF] Section 4: Electrostatics of Dielectrics
    The electric polarization P(r) is defined as the dipole moment per unit ... This leads to a polarization charge density on the surfaces of the dielectric:.
  18. [18]
    [PDF] Chapter 2 Classical Electromagnetism and Optics
    susceptibility of the classical oscillator model for the dielectric polarizability. ... The tensor χ is called the electric susceptibility tensor. As shown ...Missing: Jackson | Show results with:Jackson
  19. [19]
    The Feynman Lectures on Physics Vol. II Ch. 31: Tensors
    In other words, the dielectric property of a crystal is described completely by giving the components of the polarization tensor αij with respect to any ...
  20. [20]
    [PDF] A Companion to Classical Electrodynamics 3 Edition by J.D. Jackson
    Sep 30, 2008 · Find the polarization charge density by subtracting the effective charge den- sity from the total contained charge density: Qeff = Q + Qpol ...
  21. [21]
    52. Plane Waves in Anisotropic Media - Galileo and Einstein
    A common anisotropic crystal found in elementary labs is calcite, CaCO3, whose strange optical properties are readily observed by putting a crystal on top of ...
  22. [22]
    Ellipsometry: dielectric functions of anisotropic crystals and symmetry
    Dec 1, 2022 · The optical functions of anisotropic materials can be determined using generalized ellipsometry, which can measure the cross-polarization
  23. [23]
  24. [24]
  25. [25]
  26. [26]
    Boundary conditions for and - Richard Fitzpatrick
    In conclusion, the normal component of the electric displacement, and the tangential component of the electric field, are both continuous across any interface ...
  27. [27]
  28. [28]
    6.2 Laws and Continuity Conditions with Polarization - MIT
    If the polarization density varies with time, then the flow of charge across the surface S described in Sec. 6.1 comprises an electrical current. Thus, we need ...
  29. [29]
    [PDF] Section 1: Maxwell Equations
    static case that an electric polarization P produces a bound (polarization) charge density. P ρ = -∇∙P . (1.51). Likewise, a magnetic polarization (or ...
  30. [30]
    [PDF] Electrodynamics II [lln15]
    The bound current density is divergence free. It follows that the bound charge density and the polarization current density satisfy a separate continuity ...
  31. [31]
    [PDF] Lectures on Electromagnetic Field Theory
    May 2, 2020 · ... Polarization ... Density. 91. 10.1 Spin Angular Momentum and Cylindrical Vector Beam . . . . . . . . . . . . 91. 10.2 Complex Poynting's ...
  32. [32]
    Dielectric Relaxation - Experiment of The Month | Millersville University
    Dielectric relaxation in electrolytic capacitors occurs when rotating dipoles' orientation lags behind the electric field, causing lower capacitance at higher ...
  33. [33]
    [PDF] Origin of the Curie-von Schweidler law and the fractional capacitor ...
    Jun 10, 2020 · However, a lag in the polarization leads to dielectric relaxation. Since an insight into the optical and electrical properties of a dielectric.
  34. [34]
    Theory of polarization of crystalline solids | Phys. Rev. B
    Jan 15, 1993 · We consider the change in polarization ΔP which occurs upon making an adiabatic change in the Kohn-Sham Hamiltonian of the solid.
  35. [35]
    [PDF] Theory of Polarization: A Modern Approach - Rutgers Physics
    May 2, 2025 · The modern theory of polarization provides the correct definition of Ps, as well as the theoretical framework allowing one to compute it from.
  36. [36]
    Barium titanate ground- and excited-state properties from first ...
    Feb 23, 2011 · At 278 K, the structure becomes tetragonal, with spontaneous polarization directed along the [100] direction. Finally, at 393 K, BaTiO 3 ...
  37. [37]
    Ferroelectric domain architecture and poling of on Si
    Mar 18, 2020 · We investigate the ferroelectric domain architecture and its operando response to an external electric field in B a T i O 3 -based electro-optic ...
  38. [38]
    Wannier-function description of the electronic polarization and ...
    Dec 15, 2000 · We find that at megabar pressures the contributions to the dipoles arising from the overlapping tails of the Wannier functions are very large.
  39. [39]
    Light in correlated disordered media | Rev. Mod. Phys.
    Nov 15, 2023 · In Sec. II.A , we first focus on the “constitutive” linear relation between the average electric field and the average polarization density in ...
  40. [40]
    Aging mechanisms in amorphous phase-change materials - Nature
    Jun 24, 2015 · In the case of phase-change materials, it leads to a drift in the electrical resistance, which hinders the development of ultrahigh density ...
  41. [41]
    Exploring dielectric properties in atomistic models of amorphous ...
    In disordered systems, even in this case, such states often do not actually contribute to DC conduction due to localization effects [30]. However, likely since ...
  42. [42]
    Dielectric constant boost in amorphous sesquioxides - AIP Publishing
    This dielectric constant boost is due to (a) hardly any loss in dynamical polarizability and (b) disorder-induced IR activation of nonpolar low-energy modes ...
  43. [43]
    Is the Future of Materials Amorphous? Challenges and ...
    Due to high computational costs, a significant disparity exists between the system sizes accessible through simulations and those investigated experimentally.Figure 1 · Molecular Amorphous Solids · Figure 2
  44. [44]
    Study of dielectric relaxation and thermally activated a. c. conduction ...
    Between frequency range 1–103 Hz space charge polarization occurs and it arises because of impedance mobile charge carriers by interfaces. ... amorphous ...
  45. [45]
    Theory of ac Space-Charge Polarization Effects in Photoconductors ...
    A linear theory is developed of the ac behavior of solid or liquid materials containing charge carriers which can move freely within the material but cannot ...Missing: non- | Show results with:non-