Fact-checked by Grok 2 weeks ago

Relative permittivity

Relative permittivity, often denoted as ε_r and commonly referred to as the dielectric constant, is a that measures a material's ability to store electrical energy in an relative to . It is defined as the ratio of the of the material (ε) to the permittivity of free space (ε_0), expressed as ε_r = ε / ε_0, where ε_0 ≈ 8.85 × 10^{-12} F/m. In physical terms, relative permittivity quantifies the reduction in strength within a compared to for the same free , arising from the of bound charges in the . For non-lossy materials, ε_r is a greater than or equal to 1, with having ε_r = 1 and typical engineering materials ranging from about 2 (e.g., Teflon) to over 60 (e.g., certain ceramics). In lossy dielectrics, it becomes complex, ε_r = ε_r' - j ε_r'', where the imaginary part accounts for dissipation as . Relative permittivity plays a fundamental role in electromagnetics and , influencing the of devices—where scales directly with ε_r—and the propagation speed and attenuation of electromagnetic waves in materials. It is critical for designing components such as capacitors, antennas, circuits, and transmission lines, as well as in applications like high-frequency dielectrics and sensing. Accurate measurement of ε_r is essential across frequencies, from static fields to s, using techniques like cavity resonators or transmission lines to ensure performance in technologies ranging from to .

Fundamentals

Definition

Relative permittivity, denoted as \epsilon_r, is defined as the ratio of the permittivity of a (\epsilon) to the permittivity of (\epsilon_0), mathematically expressed as \epsilon_r = \frac{\epsilon}{\epsilon_0}. This quantity characterizes a 's response to an in terms of charge storage capacity relative to . As a dimensionless scalar, \epsilon_r quantifies the degree of induced in the by an external , which effectively reduces the net inside the compared to that in . with \epsilon_r > 1 exhibit enhanced , allowing them to support greater electric displacement for a given field intensity. The concept emerged in the 19th century from investigations into dielectrics, with coining the term "specific inductive capacity" in his 1837 studies to describe this property. By definition, \epsilon_r = 1 for ; typical values include 5–10 for and approximately 80 for at .

Terminology

Relative permittivity is also known as the dielectric constant, relative dielectric constant, and—particularly in historical contexts—specific inductive capacity. The term dielectric constant, first documented in scientific literature around 1875, became a standard descriptor for the property in the late . Specific inductive capacity, introduced by in the 1830s, referred to the material's ability to store relative to and laid the groundwork for modern understandings. In notation, the SI standard uses the symbol ε_r to denote relative permittivity, distinguishing it clearly from the absolute permittivity ε, defined as ε = ε_r ε_0 where ε_0 is the (8.854 × 10^{-12} F/m). Older texts frequently employ κ for the dielectric constant, while in German-language literature, the equivalent term Dielektrizitätskonstante is abbreviated as DK. Contextually, relative permittivity is sometimes called simply "" in discussions, though this can lead to ambiguity with absolute permittivity. In optics, for non-magnetic materials where the relative permeability μ_r ≈ 1, it relates to the n through the approximation n ≈ √ε_r, connecting electromagnetic properties across disciplines. The evolution of terminology reflects a shift toward precision: while "dielectric constant" remains in widespread use, IUPAC recommendations favor "relative permittivity" to emphasize its dimensionless, relative nature and avoid implying a true constant independent of or conditions.

Physical Principles

Relation to and Fields

In the context of , the relative permittivity \epsilon_r describes the response of a material to an applied \mathbf{E} at the macroscopic level. In , where no material is present, \epsilon_r = 1, and the \mathbf{D} is simply \mathbf{D} = \epsilon_0 \mathbf{E}, with \epsilon_0 being the . When a material is introduced, the material polarizes, inducing a \mathbf{P} that modifies the displacement field according to \mathbf{D} = \epsilon_0 \mathbf{E} + \mathbf{P}. For linear isotropic dielectrics, the polarization is proportional to the , \mathbf{P} = \epsilon_0 \chi_e \mathbf{E}, where \chi_e is the of the material. Substituting this into the expression for \mathbf{D} yields \mathbf{D} = \epsilon_0 (1 + \chi_e) \mathbf{E} = \epsilon \mathbf{E}, where \epsilon = \epsilon_0 \epsilon_r is the permittivity and \epsilon_r = 1 + \chi_e. This highlights how \epsilon_r quantifies the enhancement of the displacement field beyond the case due to material . The role of \epsilon_r becomes evident in practical devices such as parallel-plate capacitors. The capacitance C of such a device filled with a dielectric is given by C = \epsilon_r \epsilon_0 A / d, where A is the plate area and d is the separation distance. Compared to the vacuum capacitance C_0 = \epsilon_0 A / d, the factor \epsilon_r > 1 increases the stored charge for a given voltage, enhancing capacity. For example, in materials like (\epsilon_r \approx 80), this effect dramatically boosts relative to air (\epsilon_r \approx 1). At interfaces between dielectrics with different \epsilon_r, boundary conditions from govern field behavior. The tangential component of \mathbf{E} is continuous across the , while the normal component of \mathbf{D} is continuous in the absence of charge. These conditions lead to refraction of lines, analogous to in but involving the tangents of the angles \theta with the normal: \frac{\tan \theta_1}{\tan \theta_2} = \frac{\varepsilon_{r1}}{\varepsilon_{r2}}. Thus, field lines bend away from the normal when entering a higher-\varepsilon_r medium, reflecting the material's greater ability to support displacement.

Microscopic Origins

The relative permittivity of a material arises from the polarization response of its atoms and molecules to an applied electric field, where polarization P represents the dipole moment per unit volume induced by the field. This response originates at the microscopic level through several distinct mechanisms that shift or align charges within the material. The primary types of polarization include electronic, atomic (also known as ionic), orientational, and interfacial. Electronic polarization occurs due to the displacement of electron clouds relative to atomic nuclei, creating induced dipoles in all materials; this mechanism is fast and dominates at high frequencies up to the optical range (~10^{15} Hz). Atomic polarization involves the relative displacement of positively and negatively charged ions in crystalline lattices, such as in ionic solids, and is effective up to infrared frequencies (~10^{13} Hz). Orientational polarization arises from the alignment of permanent electric dipoles in polar molecules, like water, under the field; this is prevalent in liquids and gases. Interfacial polarization, also called space-charge polarization, results from the accumulation of free charges at material boundaries or defects, leading to charge separation; it is the slowest mechanism and significant at low frequencies below the kilohertz range. These microscopic polarizabilities link to the macroscopic relative permittivity \epsilon_r through relations like the Clausius-Mossotti equation, which accounts for the local field experienced by molecules in a dense medium: \frac{\epsilon_r - 1}{\epsilon_r + 2} = \frac{N \alpha}{3 \epsilon_0} Here, N is the of molecules, \alpha is the molecular , and \epsilon_0 is the . This equation connects the bulk response to atomic-scale properties, assuming non-interacting induced dipoles, and is particularly applicable to nonpolar dielectrics where electronic polarization dominates. The temperature dependence of relative permittivity is primarily influenced by orientational , which decreases with increasing temperature due to thermal agitation randomizing alignments, as described by Debye's theory of polar molecules (1929). In this model, the orientational contribution to follows P \propto 1/T, leading to a Curie-like behavior in the static constant for polar liquids. Electronic and atomic polarizations are largely temperature-independent. Frequency effects cause in \epsilon_r, where the value drops at higher frequencies because slower mechanisms—such as orientational and interfacial—cannot follow rapid field oscillations, while persists. For instance, orientational effects cease beyond frequencies due to molecular rotational . Detailed analysis of this frequency-dependent complex permittivity, including losses, is covered in the section on complex permittivity.

Measurement

Experimental Techniques

The measurement of relative permittivity, denoted as \epsilon_r, has evolved from early electrostatic experiments to precise modern techniques. In 1837, conducted pioneering work using Leyden jars to investigate the inductive capacity of materials, effectively measuring how insulators enhanced charge storage in capacitors, laying the foundation for quantitative assessments of dielectric properties. These historical methods relied on observing charge accumulation and discharge, but lacked the standardization seen today. Modern protocols, such as ASTM D150 established in 1922, provide standardized procedures for AC-based permittivity measurements on solid insulators, ensuring reproducibility across laboratories. Capacitance methods remain the cornerstone for low-frequency and static measurements of \epsilon_r. In the parallel-plate setup, a sample is inserted between two conductive plates separated by distance d, with plate area A. The C is measured using an , and \epsilon_r is computed via the formula \epsilon_r = \frac{C d}{\epsilon_0 A}, where \epsilon_0 is the ($8.85 \times 10^{-12} F/m). This approach assumes uniform fields and negligible fringing effects, achieving accuracies suitable for many engineering applications. Variations, such as the guarded , incorporate a surrounding guard ring to minimize and stray , enhancing precision for thin or irregular samples. For instance, in ASTM D150 procedures, samples are typically thin films or sheets under controlled voltage (e.g., 1 V), with measurements at frequencies from 60 Hz to 1 MHz. At higher frequencies, resonance techniques like cavity perturbation are employed to probe \epsilon_r without direct contact in some setups. A small dielectric sample is introduced into a resonant microwave cavity, perturbing its resonant frequency f_0. The frequency shift \Delta f relates to \epsilon_r approximately through \frac{\Delta f}{f_0} \propto (\epsilon_r - 1) times a geometric factor involving sample and cavity volumes, allowing extraction of \epsilon_r from calibration curves. This method excels for frequencies in the GHz range, such as 1-10 GHz using rectangular or cylindrical cavities, and is particularly useful for low-loss materials where broadband coverage is not required. Systems often automate the perturbation via vector network analyzers for rapid, non-destructive testing. For broadband characterization spanning DC to GHz, time-domain methods utilize dielectric spectroscopy with pulse propagation. Short electrical pulses are launched into a (e.g., or stripline) loaded with the sample, and the reflected or transmitted waveform is analyzed to derive time-dependent responses, from which frequency-domain \epsilon_r is obtained via . This approach, often implemented with time-domain reflectometry, captures dispersive effects over wide frequency bands (up to 100 GHz) by measuring pulse delay and , making it ideal for heterogeneous or samples. Common error sources in these techniques include poor electrode-sample contact, which introduces air gaps altering effective d, and non-uniform sample thickness, leading to averaged \epsilon_r values that deviate from properties. Fringing fields and exacerbate inaccuracies in setups, while cavity methods are sensitive to sample positioning. With careful preparation, such as vapor-deposited electrodes and micrometer-controlled spacing, precision can reach 0.1% for solid samples at low frequencies. These methods primarily yield the real part of ; complex aspects, including losses, are addressed in specialized frequency-domain analyses.

Complex Permittivity and Losses

In materials subjected to alternating , the relative permittivity ε_r becomes frequency-dependent and is generally represented as a , ε_r(ω) = ε_r' - j ε_r'', where ω is the , ε_r' is the real part associated with and , and ε_r'' is the imaginary part representing or losses. The real part ε_r' determines the material's ability to store electric , while the imaginary part ε_r'' quantifies the of into heat through various relaxation and conduction mechanisms. The extent of these losses is often characterized by the loss tangent, defined as tan δ = ε_r'' / ε_r', which provides a dimensionless measure of the of dissipated to stored ; low values of tan δ (e.g., < 0.01) indicate low-loss dielectrics suitable for high-frequency applications. Frequency dispersion in the complex relative permittivity arises from the delayed response of molecular dipoles or charges to the oscillating field, leading to variations in ε_r' and ε_r'' across different frequencies. A foundational model for this behavior is the , which describes a single relaxation process as \varepsilon_r(\omega) = \varepsilon_\infty + \frac{\varepsilon_s - \varepsilon_\infty}{1 + j \omega \tau}, where ε_s is the static (low-frequency) relative permittivity, ε_∞ is the high-frequency relative permittivity (approaching the optical limit), and τ is the relaxation time constant governing the transition between these limits. In this model, ε_r' decreases from ε_s to ε_∞ with increasing frequency, while ε_r'' peaks near the relaxation frequency ω = 1/τ, reflecting maximum energy loss. This single-relaxation approximation is particularly effective for polar liquids like water, though more complex materials may require multi-relaxation extensions. Losses in dielectrics are quantified through several methods that relate measurable quantities to the complex permittivity. Power dissipation, derived from the time-averaged Poynting theorem, is proportional to ω ε_0 ε_r'' |E|^2 / 2, where ε_0 is the vacuum permittivity and |E| is the electric field magnitude, allowing losses to be inferred from temperature rise or calorimetric measurements in the material. In resonant structures, such as cavities or transmission lines, the quality factor Q (ratio of stored to dissipated energy) provides another metric, with dielectric losses contributing to 1/Q_d = ε_r'' / ε_r' = tan δ at the resonant frequency. Impedance analysis, often using vector network analyzers with coaxial or waveguide fixtures, extracts ε_r' and ε_r'' from reflection or transmission coefficients via fitting to transmission line models, enabling broadband characterization of lossy media. Representative values of ε_r' and tan δ for common materials illustrate the range of behaviors, from high-loss polar substances to low-loss insulators. The following table summarizes data at room temperature (~20–25°C) for water (a high-permittivity, lossy dielectric) and polytetrafluoroethylene (PTFE, a low-loss polymer), highlighting frequency dependence:
MaterialFrequencyε_r'tan δNotes
Water~1 Hz (static)~80~0.0001High storage, negligible conduction loss at DC.
Water10 GHz~61~0.53Significant dispersion near relaxation frequency.
Water2.45 GHz~78~0.16Microwave regime, balanced storage and dielectric heating potential.
PTFE10 GHz~2.1~0.0002Stable low-loss behavior across microwaves.
These values underscore how polar materials like water exhibit strong frequency dispersion due to dipole reorientation, while non-polar PTFE maintains consistent low losses. The losses captured by ε_r'' enable practical applications, notably dielectric heating in microwave systems, where absorbed power P = (1/2) ω ε_0 ε_r'' ∫ |E|^2 dV generates uniform volumetric heating in materials with moderate tan δ, such as foods or polymers, without surface overheating common in conductive heating. This process, governed by the penetration depth δ ≈ 1 / (ω √(μ_0 ε_0 ε_r' / 2) √(√(1 + tan² δ) - 1)), is widely used in industrial drying, chemical synthesis, and medical therapies, with water's high ε_r'' at 2.45 GHz making it ideal for microwave ovens.

Material-Specific Behavior

Dielectrics and Lossy Media

Dielectrics are broadly classified into linear and nonlinear types based on their polarization response to an applied electric field. Linear dielectrics exhibit a constant relative permittivity independent of field strength, resulting in a proportional relationship between electric displacement and field, with minimal energy dissipation. In contrast, nonlinear dielectrics display field-dependent permittivity, often characterized by hysteresis in the displacement-electric field loop due to reversible domain reorientation or phase changes. Ferroelectric materials, a subset of nonlinear dielectrics, such as , achieve exceptionally high relative permittivities exceeding 1000, enabling applications in high-capacitance devices but introducing losses from hysteresis during polarization switching. In lossy dielectrics, energy dissipation mechanisms contribute to the imaginary component of the complex relative permittivity, particularly through finite conductivity. The conductivity σ adds to the effective dielectric loss via the term \epsilon_r'' = \frac{\sigma}{\omega \epsilon_0}, where \omega is the angular frequency and \epsilon_0 is the vacuum permittivity, representing conduction currents that convert electromagnetic energy to heat. Moisture absorption exacerbates losses in polymers by introducing polar water molecules that enhance both real and imaginary permittivity components, often increasing overall dielectric losses due to interfacial polarization and ionic conduction./03%3A_Wave_Propagation_in_General_Media/3.04%3A_Complex_Permittivity) Environmental factors significantly influence relative permittivity in dielectrics. Temperature variations alter molecular alignment and vibrational modes; for instance, the relative permittivity of water reaches a maximum of approximately 88 at 0°C, decreasing monotonically to about 55 at 100°C due to reduced hydrogen bonding strength. Humidity absorption in polymer composites can elevate relative permittivity by 10-20% through water ingress, which boosts polarizability while simultaneously raising losses via enhanced conductivity pathways. Representative examples illustrate the range of behaviors in dielectrics and lossy media. Ceramics, such as titanates and zirconates, typically exhibit relative permittivities spanning 10 to 10,000, with high-ε_r variants like suited for capacitors despite elevated losses. Insulating liquids, including mineral and vegetable transformer oils, possess low relative permittivities of 2-4, minimizing losses in high-voltage insulation while providing thermal management. Recent advances since 2020 have focused on polymer-ceramic hybrids for 5G applications, achieving tunable relative permittivities of 2-10 with ultralow losses below 0.001 through optimized filler dispersion and core-shell structures that suppress interfacial polarization.

Conductors and Metals

In conductors and metals, the relative permittivity \epsilon_r is generally complex and frequency-dependent due to the presence of free charge carriers, which dominate the response over bound charges. Unlike dielectrics, where \epsilon_r is typically positive and real at low frequencies, in highly conductive materials, the imaginary part \epsilon_r'' becomes significant, reflecting energy dissipation through conduction. This behavior arises from the collective motion of free electrons, modeled classically by the , which treats electrons as a gas subject to damping collisions. The Drude model yields the complex relative permittivity as \epsilon_r(\omega) = 1 - \frac{\omega_p^2}{\omega^2 + j \gamma \omega}, where \omega_p = \sqrt{n e^2 / (\epsilon_0 m)} is the plasma frequency, with n the free electron density, e the electron charge, \epsilon_0 the vacuum permittivity, m the electron mass, \omega the angular frequency, and \gamma the damping rate (inverse collision time). For metals like copper or silver, \omega_p typically ranges from $10^{15} to $10^{16} rad/s, corresponding to ultraviolet wavelengths. Below \omega_p, the real part \operatorname{Re}(\epsilon_r) < 0, leading to evanescent waves and high reflectivity, as electromagnetic waves cannot propagate into the material; this explains the metallic sheen observed at visible frequencies. The negative \operatorname{Re}(\epsilon_r) results from the inertial response of free electrons overpowering the positive background lattice contribution. At low frequencies, well below \omega_p, the high conductivity \sigma = n e^2 / (m \gamma) dominates, making \epsilon_r effectively infinite in magnitude due to the large imaginary component \epsilon_r'' \approx \sigma / (\epsilon_0 \omega). This regime manifests as the skin effect, where alternating currents concentrate near the conductor surface, with penetration depth \delta = \sqrt{\frac{2}{\omega \mu \sigma}}, \mu being the permeability (approximately \mu_0 for non-magnetic metals). For copper at 60 Hz, \delta \approx 8.5 mm, limiting field penetration and increasing effective resistance. Such behavior is crucial for understanding power transmission losses and RF shielding. In superconductors, the Drude-like model applies with zero damping (\gamma = 0) below the critical temperature, yielding \epsilon_r \approx 1 - (\omega_p / \omega)^2 for frequencies above the superconducting energy gap, with negligible imaginary part indicating lossless response. The plasma frequency remains similar to normal metals (\omega_p \sim 10^{15} rad/s), but the absence of scattering enables perfect conductivity. The Meissner effect further expels magnetic fields from the interior, equivalent to perfect diamagnetism (\mu_r = 0), preventing static field penetration and reinforcing the reflective properties at low frequencies. This is observed in type-I superconductors like lead, where fields are completely screened. Engineered metamaterials extend negative \epsilon_r into desired bands, including the visible spectrum, via plasmonic nanostructures such as metal-dielectric multilayers or nanorods. Since the 2000s, these have enabled applications like electromagnetic cloaking by achieving \operatorname{Re}(\epsilon_r) < 0 without natural plasma frequencies, using subwavelength arrays to mimic effective media with hyperbolic dispersion. For instance, vertically stacked silver-silica hyperbolic metamaterials demonstrate broadband negative refraction from 400 to 700 nm, with low loss and transmittance up to 18% at 600 nm, far beyond bulk metals. Representative examples illustrate these behaviors. For silver, spectroscopic ellipsometry reveals \epsilon_r \approx -11.8 + 0.37i at 532 nm (visible), with \operatorname{Re}(\epsilon_r) highly negative across optical frequencies due to free-electron plasmons near 9 eV. Electrolytes, such as aqueous NaOH solutions, are modeled as lossy conductors with ionic free carriers, exhibiting \operatorname{Re}(\epsilon_r) > 100 at low frequencies (e.g., 60.7 at 1.5 M concentration, 0.2 GHz), attributed to interfacial and bulk ionic enhancing the effective before conduction dominates.

Applications

Energy Storage and Capacitors

The energy density stored in a dielectric capacitor is fundamentally enhanced by materials with high relative permittivity ε_r, as described by the expression for energy U = \frac{1}{2} \epsilon_r \epsilon_0 E^2 V, where \epsilon_0 is the vacuum permittivity, E is the applied electric field strength, and V is the volume. This relationship underscores how increasing ε_r allows for greater charge storage at a given voltage and geometry, making high-ε_r dielectrics essential for compact, efficient energy storage devices. In practice, multilayer ceramic capacitors (MLCCs) leverage ferroelectric ceramics with ε_r exceeding 1000 to achieve high capacitance in high-voltage applications, such as power electronics and pulse discharge systems, where space constraints demand elevated energy density without excessive size. A key challenge in capacitor design lies in the inherent trade-off between achieving high ε_r and maintaining sufficient dielectric breakdown strength, as materials with elevated permittivity often exhibit reduced resistance to high electric fields. For instance, mica-based capacitors, with a modest ε_r of approximately 6–8, offer exceptional breakdown strength up to 1000 V/μm, making them suitable for high-reliability, high-voltage environments like RF and systems, where stability under stress is paramount over maximum . This balance ensures reliable operation, as exceeding the breakdown field leads to irreversible dielectric failure and energy loss. In advanced energy storage, pseudocapacitors incorporate high-ε_r electrolytes like s, which typically exhibit relative permittivities of 10–20, to facilitate faradaic charge transfer and enhance overall device performance. These electrolytes enable pseudocapacitors to achieve cycle lives surpassing 10^6 charge-discharge cycles while maintaining high , as demonstrated in systems combining conducting polymers or metal oxides with ionic liquid media for applications in hybrid vehicles and buffering. The historical evolution of capacitors highlights the role of high-ε_r materials in scaling ; electrolytic capacitors, pioneered in the with wet electrolytes, achieve high through very thin layers (ε_r ≈ 7–10) despite the low , enabling compact designs for early radio and . Today, MLCCs dominate approximately 93% of the market since the , driven by their scalability and integration in and automotive systems. For low-loss , a critical is ε_r / tan δ, where tan δ represents the tangent, prioritizing materials that maximize stored energy while minimizing dissipative heating during charge-discharge cycles.

Electromagnetic Communications

In electromagnetic communications, relative permittivity fundamentally governs the propagation characteristics of signals across various media, from radio frequencies to optical regimes. The speed of electromagnetic waves in a non-magnetic dielectric (μ_r ≈ 1) is reduced from the vacuum speed c by the factor 1/√ε_r, expressed as v = c / √ε_r, which directly impacts delay and phase synchronization in transmission lines and waveguides. Similarly, the characteristic impedance of the medium, Z = Z_0 / √ε_r where Z_0 ≈ 377 Ω is the free-space impedance, influences reflection coefficients, power transfer efficiency, and matching requirements in cables and antennas. In practical coaxial or microstrip cables, frequency-dependent variations in ε_r introduce dispersion, where different frequency components travel at slightly varying speeds, leading to pulse broadening and inter-symbol interference over long distances. For antennas and waveguides, the choice of substrate ε_r critically affects physical dimensions, radiation efficiency, and operational bandwidth. Higher ε_r values concentrate the electromagnetic fields within the substrate, enabling antenna miniaturization by a factor of approximately 1/√ε_r while maintaining resonance at a given frequency, though this often narrows bandwidth due to increased quality factor Q and stored reactive energy. For instance, Rogers RO4000 series hydrocarbon ceramic laminates, with a stable ε_r of 3.38 across broad frequencies, are widely adopted in printed circuit board (PCB) substrates for microwave antennas, supporting compact designs in radar and wireless systems with minimal dispersion. In waveguides, low-ε_r substrates enhance mode confinement and reduce ohmic losses, broadening usable bandwidth for high-data-rate links. Optical communications leverage relative permittivity contrasts for light guidance in fibers and photonic structures. In step-index optical fibers, total internal reflection occurs at the core-cladding interface due to the core's higher ε_r; for standard silica fibers, the core ε_r ≈ 2.13 (refractive index n ≈ 1.46) exceeds the cladding's by a small margin (Δε_r ≈ 0.01–0.02), enabling low-loss propagation over kilometers. Photonic crystals, engineered with periodic ε_r modulations (e.g., alternating high- and low-ε_r regions in silicon or polymer lattices), create photonic bandgaps that inhibit light propagation in specific bands, facilitating compact waveguides, multiplexers, and dispersion compensators for wavelength-division multiplexing (WDM) systems. In millimeter-wave (mm-wave) applications for and beyond, low-ε_r materials are essential to mitigate and surface wave losses at frequencies above 24 GHz. Foamed or porous substrates with ε_r < 1.5, such as lightweight silica aerogels (ε_r ≈ 1.03) or expanded polypropylene foams, serve as low-loss backings for phased-array antennas, improving gain and efficiency while reducing weight in base stations and handsets. For terahertz (THz) communications, emerging since 2022, ε_r-engineered polymers like photopolymer composites enable flexible, low-loss waveguides and antennas operating at 0.1–1 THz, supporting ultra-high data rates with minimal dispersion through tailored permittivity gradients. These materials address THz-specific challenges, including high atmospheric absorption, by optimizing waveguiding with ε_r values near 2–3. Signal integrity in high-speed integrated circuits benefits from strategic ε_r placement to suppress crosstalk between adjacent interconnects. High-ε_r barriers, such as silicon nitride (ε_r ≈ 7) layers or capacitors integrated near signal lines, enhance capacitive shunting to ground, isolating fields and reducing inductive/capacitive coupling by up to 20 dB in multi-Gbps links. This approach complements low-ε_r inter-metal dielectrics in back-end-of-line (BEOL) processing, ensuring reliable eye diagrams and bit error rates below 10^{-12} in 5G modems and optical transceivers. As of November 2025, advancements in high-ε_r low-loss dielectrics are enabling enhanced performance in , where materials with tunable ε_r support reconfigurable intelligent surfaces for improved beamforming and coverage in sub-THz bands.

Chemical and Environmental Sensing

Dielectric spectroscopy leverages variations in relative permittivity (ε_r) to detect changes in chemical composition and environmental conditions, enabling non-invasive sensing of analytes through shifts in the dielectric response of materials. In soil moisture monitoring, ε_r exhibits a dramatic increase from approximately 4 for dry soil to 80 for water-saturated conditions, allowing sensors to quantify water content accurately via electromagnetic probes. This principle underpins techniques, where the propagation velocity of electromagnetic waves correlates with ε_r, providing real-time data for agricultural and hydrological applications. Biosensors exploit ε_r shifts in microfluidic channels for biomolecular detection, such as proteins or nucleic acids, where binding events alter the local dielectric environment. For instance, dry DNA exhibits an ε_r of about 2.5, while hydration increases it to around 70 due to water molecule orientation, enabling label-free detection of conformational changes or analyte binding through impedance measurements. These shifts are measured using interdigitated electrodes integrated into microchannels, offering high sensitivity for point-of-care diagnostics without optical labels. In protein detection, specific binding induces measurable capacitance changes, distinguishing target molecules from non-specific interactions at concentrations as low as nanomolar levels. Environmental monitoring employs permittivity probes to assess soil salinity and atmospheric conditions, where ε_r correlates with ionic content and moisture. For soil salinity, increased salt concentrations elevate the imaginary part of the complex permittivity, affecting the real ε_r and allowing indirect estimation of electrical conductivity (EC) alongside water content; this relationship deviates from low-salinity models but can be calibrated for accurate profiling in coastal or irrigated areas. The seminal Topp equation from 1980 relates bulk ε_r to volumetric water content across soil types, though salinity adjustments are necessary for high-EC environments. In atmospheric sensing, variations in air's ε_r (typically near 1 but influenced by humidity and aerosols) aid lightning prediction models by informing dielectric breakdown thresholds in thunderclouds. Chemical analysis utilizes impedance spectroscopy to probe solution properties, such as pH, where ε_r of aqueous media ranges from 50 to 80 depending on ion dissociation and hydrogen bonding. At neutral pH, pure water's ε_r approaches 80, but acidic or basic conditions alter dipole moments, leading to detectable frequency-dependent losses that correlate with proton activity. Nanoscale sensors incorporating since 2018 enable volatile organic compound (VOC) detection by tuning the effective ε_r through adsorption-induced charge transfer, shifting resonance frequencies in metasurface or field-effect configurations for selective gas identification at parts-per-billion levels. Loss mechanisms, such as interfacial polarization from ionic mobility, briefly influence these measurements but are secondary to real ε_r changes in dilute solutions. Calibration in these systems often involves frequency sweeps across broadband ranges (e.g., 1 MHz to 10 GHz) to distinguish (relaxation ~100 MHz, lower ε_r due to restricted mobility) from (~20 GHz, higher ε_r akin to bulk liquid). This differentiation is critical for accurate quantification in heterogeneous media like soils or biological fluids, where bound layers contribute minimally to overall permittivity compared to free water fractions. Such sweeps ensure sensor reliability by isolating dispersion regions, mitigating errors from temperature or salinity variations.

Advanced Materials and Devices

In metamaterials, negative relative permittivity (ε_r < 0) enables extraordinary electromagnetic properties, such as superlensing, where subwavelength imaging beyond the diffraction limit is achieved. The theoretical foundation for such materials with simultaneous negative ε_r and permeability (μ_r < 0) was laid by Veselago in 1968, with Pendry's 2000 proposal of a "perfect lens" using a slab with ε_r = -1 focusing evanescent waves. Experimental realization at microwave frequencies, demonstrating negative refraction and superlensing with ε_r ≈ -1, was reported in 2006 using split-ring resonator arrays. These designs have since influenced applications in cloaking and high-resolution imaging, with ongoing refinements to extend negative ε_r to optical regimes. Ferroelectric materials exhibit tunable relative permittivity under applied electric fields, enabling dynamic control in devices like actuators. In relaxor ferroelectrics such as Pb(Mg_{1/3}Nb_{2/3})O_3-PbTiO_3 (PMN-PT) single crystals, ε_r exceeds 5000 near the morphotropic phase boundary, allowing field-induced enhancements up to 20-30% for piezoelectric actuation. Strain-coupled multiferroic variants, like PMN-PT composites with magnetostrictive layers, further modulate ε_r through mechanical stress, achieving bidirectional tuning for sensors and transducers. Recent advances in thin films maintain high tunability (ε_r > 1000) while reducing losses, as demonstrated in 2022 studies on non-stoichiometric PMN-PT. Nanodielectrics incorporate nanoparticles to enhance relative permittivity in matrices while preserving low dielectric losses, critical for high-performance capacitors. For instance, incorporating 1 wt% TiO_2 nanoparticles into increases ε_r by approximately 50% (from ~3 to ~4.5) at low frequencies, with minimal increase in loss tangent (<0.01), due to interfacial polarization effects. Similar enhancements occur in TiO_2-polypropylene composites, where nanoparticle dispersion boosts ε_r by 20-60% depending on loading, enabling higher energy densities without breakdown. These materials leverage core-shell structures to suppress agglomeration, as optimized in 2022 formulations. Two-dimensional materials like hexagonal boron nitride (h-BN) display anisotropic relative permittivity, with in-plane ε_r ≈ 4-5 and out-of-plane ε_r ≈ 2.5-3, arising from layered ionic bonding. This anisotropy supports its use as a dielectric spacer in van der Waals heterostructures, providing electrical isolation with low leakage. Recent 2023 ab initio calculations confirm these values across layer thicknesses, predicting convergence to bulk anisotropy (ε_∥ / ε_⊥ ≈ 1.5-2) beyond 5 nm. In quantum dot dielectrics, photoinduced modulation of ε_r in interacting quantum dot arrays enables coherent control for qubit operations, with field-dependent shifts up to 10-15% reported in 2023 experiments. Research from 2023-2025 highlights ε_r tuning via external fields to mitigate decoherence in silicon-based qubits, enhancing fidelity in quantum computing prototypes. Microelectromechanical systems (MEMS) utilize relative permittivity switching through dielectric liquids or ferroelectrics for tunable RF filters, achieving bandwidth reconfiguration with insertion losses <1 dB. In dual-liquid MEMS designs, varying ε_r from 2 to 80 via fluid displacement tunes center frequencies by 20-50%, as demonstrated in 2022 prototypes operating up to 10 GHz. For energy harvesting, permittivity gradients in piezoelectrics generate additional voltages from strain inhomogeneities, boosting output power by 15-30% in gradient-structured devices. In PMN-PT-based harvesters, engineered ε_r variations along the poling direction exploit flexoelectric coupling, converting ambient vibrations to usable energy with efficiencies >20 μW/cm².

References

  1. [1]
  2. [2]
    [PDF] High-Frequency Dielectric Measurements
    Permittivity is most often expressed as a relative value r: = 0r = 0(r′ - jr′′), where 0 ≅ 8.854 x 10-12 F/m.
  3. [3]
    A Simplified Measurement Configuration for Evaluation of Relative ...
    Jan 25, 2022 · Relative permittivity of a material, ϵ r , is an important parameter for the design of antennas and microwave circuits because it strongly ...
  4. [4]
    [PDF] Lecture 9
    - The variable εr is known as the relative electric permittivity of a material, or its dielectric constant. It is defined relative to the permittivity of free ...
  5. [5]
    Dielectrics - HyperPhysics Concepts
    The permittivity is a characteristic of space, and the relative permittivity or "dielectric constant" is a way to characterize the reduction in effective ...
  6. [6]
    On the specific inductive capacities of certain electric substances
    Faraday relative to the differences in specific inductive capacity exhibited by different dialectric substances, instituted a series of experiments for ...
  7. [7]
    [PDF] The Electrical Discoveries of Joseph Henry - World Radio History
    Proportioning of Magnet and Battery for Maximum Effect {1830). Henry pointed out that the coils of the magnets constructed by his method could be made up either ...
  8. [8]
    Dielectric Constants - HyperPhysics
    Material, Dielectric Constant. Vacuum. 1. Glass. 5-10. Mica. 3-6. Mylar. 3.1. Neoprene. 6.70. Plexiglas. 3.40. Polyethylene. 2.25. Polyvinyl chloride. 3.18.
  9. [9]
    Relative Permittivity - the Dielectric Constant
    Water - permittivity is 88 at 0 oC (31 oF) and drops with rising temperature. Permittivity is 80 at 20 oC (211 oF) and 55.3 at 100 oC (211 oF) ...
  10. [10]
    relative permittivity (R05273) - IUPAC
    Ratio of the electric field strength in vacuum to that in a given medium. It was formerly called the dielectric constant.
  11. [11]
    Michael Faraday | Biography, Inventions, & Facts - Britannica
    Oct 27, 2025 · In short, why should not every material have a specific inductive capacity? Every material does, and Faraday was the discoverer of this fact ...
  12. [12]
    I. Note on specific inductive capacity - Journals
    History: Manuscript received09/11/1886; Published online01/01/1997; Published in print01/01/1887. License: Scanned images copyright © 2017, Royal Society. Get ...
  13. [13]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    Physicists prefer to use the complex relative permittivity (see Section 2.3, p. 16), bεr = εr. 0 + iεr. 00 = Re bεr + iIm bεr instead of the complex ...
  14. [14]
    [PDF] Basics of Measuring the Dielectric Properties of Materials
    Dielectric constant (k) is equivalent to relative permittivity (εr) or the absolute permittivity. (ε) relative to the permittivity of free space (ε0). The ...
  15. [15]
    The dielectric constant - DoITPoMS
    The dielectric constant (sometimes called the 'relative permittivity') is the ratio of the permittivity of the dielectric to the permittivity of a vacuum, so ...Missing: synonyms | Show results with:synonyms
  16. [16]
    Dielectric Constant & Relative Permittivity - Electronics Notes
    The dielectric constant & relative permittivity are key to the operation of capacitors and the determination of the levels of capacitance achievable.
  17. [17]
    of refraction, index contrast - RP Photonics
    ... the vacuum wavelength. The refractive index can be calculated from the relative permittivity ϵ r and the relative permeability μ r of an optical material:.
  18. [18]
    [PDF] Section 4: Electrostatics of Dielectrics
    ε ε χ. = +. (4.25) is the electric permittivity. εr = ε /ε0 = 1 + χe is called the dielectric constant or relative electric permittivity. Page 6. 6. If the ...
  19. [19]
    The Feynman Lectures on Physics Vol. II Ch. 10: Dielectrics - Caltech
    The phenomenon of the dielectric constant is explained by the effect of the charges which would be induced on each sphere. This is one of the earliest physical ...
  20. [20]
    [PDF] INLINE RELATIVE PERMITTIVITY SENSING USING SILICON ...
    capacitor is [8]:. Cs(εr) = ε0εr. A d. = βεr,. (1) with ε0 the constant vacuum permittivity, εr the relative permittivity, A the surface area of the electrodes ...
  21. [21]
    [PDF] Summary of Electrostatics with Dielectrics
    ϵr=ϵ/ϵ0 : Relative electric permittivity, also called the dielectric constant ... Boundary Conditions: (ϵ2 E2−ϵ1 E1)⋅n12=σ and (E2−E1)×n12=0.
  22. [22]
    The Feynman Lectures on Physics Vol. II Ch. 11: Inside Dielectrics
    If there are N atoms in a unit volume, the polarization P—the dipole moment per unit volume—is given by P=Np=Nαϵ0E. Putting (11.1) and (11.8) together, we get ...
  23. [23]
    [PDF] 1 Fundamentals of Dielectrics - Wiley-VCH
    where the subscripts on the right refer to the four types: electronic, ionic, orientational, and space charge polarization, respectively. 1.1.2 Dispersion of ...<|control11|><|separator|>
  24. [24]
    The Debye's model for the dielectric relaxation of liquid water and ...
    Oct 3, 2023 · Debye arrived to this simple model in 19292 by considering a thermal ensemble of non-interacting rigid dipoles (molecules) subjected to an ...
  25. [25]
    Modeling static charge dissipation on solids: An historical perspective
    In his pioneering 1837 experiments, Michael Faraday (Fig. 2) set the stage for this complexity. Using a charged Leyden jar as a generator, a Coulomb torsion ...
  26. [26]
    Standard Test Methods for AC Loss Characteristics and Permittivity ...
    Oct 18, 2022 · D150 Standard Test Methods for AC Loss Characteristics and Permittivity (Dielectric Constant) of Solid Electrical Insulation.
  27. [27]
    Complex permittivity measurements using cavity perturbation ...
    Jun 11, 2010 · This paper presents the theoretical formulas for the evaluation of complex permittivity of materials using cavity perturbation technique with substrate ...
  28. [28]
    [PDF] Dielectric Characterization by Microwave Cavity Perturbation ...
    Dec 11, 2015 · The nonuniform field decreases the displacement field inside the sample near the boundary between the sample and the cavity, which in practice.
  29. [29]
    Automated Cavity Perturbation Method for Measurement of ...
    Dec 8, 2011 · Cavity perturbation techniques 1,2 have been widely used for the measurement of dielectric properties at microwave frequencies.
  30. [30]
    Time domain dielectric spectroscopy: An advanced measuring system
    The generator produces 200 mV pulses with 10 ms duration and 28 ps rise time. ... The equivalent circuit for a conductive dielectric sample with electrode ...<|separator|>
  31. [31]
    Measuring Spectral Dielectric Properties Using Gated Time Domain ...
    Aug 1, 2003 · The results demonstrate that the gated impulse time domain method has promise for measuring the frequency-dependent dielectric permittivity with ...
  32. [32]
    [PDF] Improving the Accuracy of Dielectric Measurements - Novocontrol
    Feb 22, 2018 · Sample preparation is key for accuracy. Inaccuracies come from measurement system, cables, sample cell, and metallization. Proper metallization ...
  33. [33]
    [PDF] Dielectric Properties and Metamaterials - Lehigh University
    Jan 15, 2008 · (b) The relative permittivity is a complex number with real (εr. ') and imaginary (εr. '') parts that exhibit frequency dependence. Time.
  34. [34]
    A Complex Permittivity Based Sensor for the Electrical ...
    It is a measure of how dissipative or lossy a material can be to an external electric field. According to the theory, ε r″ is always > 0 and is usually much ...
  35. [35]
    [PDF] Dielectric and Conductor-Loss Characterization and Measurements ...
    The relative permittivity is obtained from the resonant frequencies of a cavity. ... The temperature coefficient of permittivity is defined as τ² = 1. ²0r ...<|control11|><|separator|>
  36. [36]
    [PDF] A Model for the Complex Permittivity of Water at Frequencies Below ...
    The frequency dependence of the permittivity of pure water is given by the well- known Debye equation or its modification by Cole-Cole. Up to about 100 GHz, the.
  37. [37]
    [PDF] Model for Dielectric Constant of Seawater based on L-band ...
    The parameters ε∞, εs and τ have the same roles as before but now are functions of both salinity and temperature whereas in (1) they were only functions of ...Missing: formula | Show results with:formula
  38. [38]
    11.5 Electromagnetic Dissipation - MIT
    In the following example, we compute the complex permittivity from a model of the polarizable medium and find the electrical dissipation on a macroscopic basis.
  39. [39]
    [PDF] Dielectric and conductor-loss characterization and measurements ...
    The metal losses in a resonant structure are related to the losses in the dielectric in terms of the quality factor of the dielectric, Q^ and conductor, Q^ as.
  40. [40]
    [PDF] Pub236.pdf - Stanford University
    We apply impedance spectroscopy to measure the dielectric loss tangent as a function of tempera- ture and frequency. From the results, we identify two ...
  41. [41]
    [PDF] PERMITTIVITY AND MEASUREMENTS - Washington State University
    E is the electric field intensity in V/m, f is the frequency in. Hz, Dt is the time duration in seconds, and DT is the tem- perature rise in the material in 1C.Missing: dissipation | Show results with:dissipation
  42. [42]
    [PDF] Dielectric Property Measurements in the Electromagnetic Properties ...
    The reference value for Teflon isgiven as 2.10 for a frequency of 10.0 GHz (Reference. 3). As can be seen from the plots, the waveguide and dielectric probe ...Missing: PTFE | Show results with:PTFE
  43. [43]
    [PDF] Measurements of dielectric constants and loss tangents at e-band ...
    Dielectric constant and loss tangent measurements for polytetrafluoroethylene (teflon), polystyrene, and polymethyl methacrylate (lucite, Plexiglas) were made ...Missing: analysis | Show results with:analysis
  44. [44]
    Review on Microwave-Matter Interaction Fundamentals and Efficient ...
    Microwave heating is commonly known as dielectric heating and its application is usually limited to microwave electronic field heating. The information ...
  45. [45]
    Microwave Dielectric Heating of Drops in Microfluidic Devices† - PMC
    We present a technique to locally and rapidly heat water drops in microfluidic devices with microwave dielectric heating. Water absorbs microwave power more ...
  46. [46]
    An Overview of Linear Dielectric Polymers and Their ... - NIH
    This review summarizes some typical studies on linear dielectric polymers and their nanocomposites, including linear dielectric polymer blends, ferroelectric/ ...
  47. [47]
    Hysteresis and tunability characteristics of Ba(Zr,Ti)O3 ceramics ...
    The tunability of a ferroelectric describes the ability to change its permittivity by the electric field. This property is directly related to the non-linear ...
  48. [48]
    Research on dielectric non-linearity of BaTiO3-based ferroelectric ...
    In this paper, the mechanism of dielectric nonlinearity and relative theory of BaTiO3-based ferroelectric materials in ferroelectric phase, paraelectric ...
  49. [49]
    [PDF] Measuring the permittivity and permeability of lossy materials
    This document discusses measuring permittivity and permeability of lossy materials including solids, liquids, metals, building materials, and negative-index ...
  50. [50]
    The effect of moisture contamination on the relative permittivity of ...
    Sep 15, 2015 · In the most significant case, a 3.7% increase in water content by weight resulted in a 43% increase in relative permittivity for BMI/quartz ...
  51. [51]
    [PDF] Dielectric constant of water from 0° to 100° C
    Determinations of the dielectric constant of water were made in cell A at 5-deg intervals, using two or more samples of water at each temperature. As a further ...
  52. [52]
    The Effects of Humidity on Dielectric Permittivity of Surface-Modified ...
    The measured increase in relative permittivity in the composites is attributed to absorbed water. Surface modification significantly reduces this increase, ...
  53. [53]
    [PDF] evaluation of moisture effects on the Dielectric properties of polymer ...
    Absorbed moisture can have a significant effect on polymer properties, particularly when the end-use form is a thin film.
  54. [54]
    [PDF] Progress in High-Dielectric Constant Materials - OSTI.GOV
    High-dielectric constant materials have a relative permittivity greater than silicon nitride (around 7), enabling miniaturization and energy storage. Examples ...
  55. [55]
    Polymer-based dielectrics with high permittivity and low dielectric ...
    In this work, starting from a brief introduction to the current state of polymer-based dielectrics, polarization principles (atomic polarization, electronic ...
  56. [56]
    [PDF] Broadband Dielectric Characterization of Polymers and Ceramics in ...
    Apr 8, 2021 · dielectric or the metal loss. • Both dielectric and metal losses increase with frequency in the 5G frequency range. • Trade-offs in dielectric ...Missing: hybrids tunable epsilon_r
  57. [57]
    [PDF] Drude Model for Metals | EMPossible
    The Drude model for metals is derived from the Lorentz model by setting 𝜔 0. 𝑁 is now interpreted as free electron density 𝑁 .
  58. [58]
    Dielectric constant of a plasma - Richard Fitzpatrick
    The dielectric constant of a plasma is less than unity above the plasma frequency, and negative below it. The plasma frequency is related to the electron ...
  59. [59]
    Understanding the Relative Permittivity of Copper in PCB Design
    The relative permittivity of copper is said to be infinite. This means copper does not store electric energy like insulators do.
  60. [60]
    Electrodynamics of high- T c superconductors
    Aug 26, 2005 · It is customary to discuss electrodynamics of metals and superconductors in terms of the complex dielectric function ϵ = ϵ 1 ( ω ) + i ϵ 2 ( ω ) ...
  61. [61]
    Molecular Plasmonics with Metamaterials | Chemical Reviews
    Oct 4, 2022 · In this paper, we provide a comprehensive overview of the developments of molecular plasmonics with metamaterials. After a brief introduction to ...
  62. [62]
    Realization of broadband negative refraction in visible range using ...
    Oct 1, 2019 · This paper reports the realization of broadband negative refraction in the visible spectrum by using vertically-stacked metal-dielectric multilayer structures.<|separator|>
  63. [63]
  64. [64]
    [PDF] Complex Permittivity of NaOH Solutions Used in Liquid-Metal Circuits
    Aug 24, 2019 · However, NaOH is an aqueous electrolyte, and thus is lossy at RF. Nevertheless, careful design and simulation can help minimize the effects of ...
  65. [65]
    High-energy-density polymer dielectrics via compositional and ...
    Aug 19, 2022 · For linear dielectrics, the energy density (Ue) equation is described as follows:(Equation 1) U e = 0.5 ε 0 ε r E b 2 where ϵ0 is the vacuum ...<|separator|>
  66. [66]
    Electroceramics for High-Energy Density Capacitors: Current Status ...
    Apr 28, 2021 · Class II ceramic capacitors have a dielectric with a high permittivity. C and C0 are represented capacitance value and capacitance value at 25 ° ...
  67. [67]
    Dielectric Strength of Insulating Materials - The Engineering ToolBox
    The dielectric strength of a material is the ability to the material to act as an insulator. ; Mica (muscovite) · Glass · Mica (phlogopite) ; 39.4 · 35.5 · 31.5 ; 1000.Missing: μm | Show results with:μm
  68. [68]
    Mica Dielectric Constant - The Dielectric Properties of Mica
    Mica's dielectric constant ranges from 6.4 to 9.3, with an average of 8.1, indicating a relatively low value.
  69. [69]
    Dielectric study on mixtures of ionic liquids | Scientific Reports - Nature
    Aug 7, 2017 · We demonstrate that mixing two ionic liquids allows to precisely tune their physical properties, like the dc conductivity.
  70. [70]
    Pseudocapacitance - Wikipedia
    Furthermore, the reversibility on these transition metal electrodes is excellent, with a cycle life of more than several hundred-thousand cycles.
  71. [71]
    History Of The Capacitor – The Modern Era | Hackaday
    Jul 26, 2016 · The first “wet” electrolytic capacitors appeared in radios briefly in the 1920s but had a limited lifespan.
  72. [72]
    Top MLCC Manufacturers in the World - Cytech Systems
    Apr 3, 2024 · ... market share, with ceramic capacitors occupying 56% of the capacitor market, and MLCCs accounting for 93% of the ceramic capacitor market share.
  73. [73]
    Capacitor Fundamentals: Part 8 – Dielectric Classifications
    Jun 12, 2023 · Generally speaking, there is a trade-off such that dielectrics with a higher dielectric constant K have greater losses and less stability in ...
  74. [74]
    [PDF] Mathematical description of EM waves
    The speed of an EM wave in free space is given by: ε0 = permittivity of free ... For most physical media,. N > 1 (i.e., the speed of light is reduced relative to ...
  75. [75]
    Characteristic Impedance | Transmission Lines | Electronics Textbook
    Characteristic impedance is also known as natural impedance, and it refers to the equivalent resistance of a transmission line if it were infinitely long.
  76. [76]
    47. Frequency Dispersion in Dielectrics and Conductors
    "Frequency dispersion" means the permittivity (a.k.a. dielectric constant) is a function of frequency: ε=ε(ω). Think of light going through a prism ...
  77. [77]
  78. [78]
    [PDF] ROGERS CORPORATION: RO4000 High Frequency Materials Data ...
    RO4000 materials exhibit a stable dielectric constant over a broad frequency range. (Chart 2). This makes it an ideal substrate for broadband applications.Missing: bandwidth | Show results with:bandwidth
  79. [79]
    Permittivity of Some Common Materials | Electromagnetics I
    Permittivity of Some Common Materials ; Polyethylene, 2.3, coaxial cable ; Polypropylene, 2.3 ; Silica, 2.4, optical fiber ; Polystyrene, 2.6.
  80. [80]
    Photonic Crystal For Communication Applications
    Photonic crystals can also exhibit strong dispersion properties. These properties may offer many exciting opportunities for optical communication applications.
  81. [81]
    Lightweight porous silica foams with extreme-low dielectric ...
    Jan 5, 2021 · In this work, highly porous (98.9% ± 0.1%) and lightweight silica foams (0.025 ± 0.005 g/cm 3 ), that have extremely low relative permittivity.
  82. [82]
    220–325 GHz all-photopolymer Bragg horn antennas towards eco ...
    Jul 30, 2025 · This paper presents the development of the world's first high-gain, all-photopolymer Bragg horn antennas explicitly designed for the WR-3.4 ...
  83. [83]
    Electrical Characterization of Shielded TSVs with Airgap Isolation for ...
    To this end, the proposed shielded TSVs with airgap isolation enable low-loss and low-coupling integration for. RF/mmWave and other high-frequency applications.
  84. [84]
    Dielectric Permittivity and Moisture Content | Request PDF
    As water has the largest real permittivity value, (close to 80,) as compared to the real permittivity of dry soils, which ranges from 3 to 15, the measurement ...
  85. [85]
    [PDF] Soil Water Content: Dielectric Sensors & Calibration
    Jan 30, 2023 · relative dielectric permittivity is much smaller than that of pure water (mean values of 4 versus 80; Behari, 2005; Malm- berg and Maryott ...Missing: spectroscopy jump
  86. [86]
    Direct measurement of the dielectric polarization properties of DNA
    Measured dielectric constants associated to DNA in the literature are therefore relatively high values in the range of 20–60 (20–23), which account for large ...
  87. [87]
    (PDF) Electrical detection of protein biomarkers using bioactivated ...
    Aug 7, 2025 · The use of bioactivated microfluidic channels for electrical detection of protein biomarkers using impedance sensing has also been investigated ...
  88. [88]
    Impact of Soil Salinity on the Relation Between Soil Moisture and ...
    Measurements of apparent dielectric permittivity (Ka), ECa, and soil temperature were taken under different electrical conductivity of soil moisture solutions.
  89. [89]
    Electromagnetic determination of soil water content: Measurements ...
    Topp, A. P. Annan, Electromagnetic detection of soil water content, Proceedings of the Workshop on Remote Sensing of Soil Moisture and Groundwater, Toronto, ...
  90. [90]
    Advanced numerical model of lightning development: Application to ...
    May 16, 2017 · An advanced 3-D numerical model of lightning development is presented. The key features of the model include the probabilistic branching, ...
  91. [91]
    Detection of medically relevant volatile organic compounds with ...
    Dec 12, 2023 · In this paper, we propose a device for VOC detection that was tested inside a controlled gas flow setup, resorting to graphene field-effect transistors (GFETs).
  92. [92]
    Bound and Free Water Determination by Dielectric Spectros | PDF
    Dielectric spectroscopy can detect two relaxation peaks, one around 100MHz related to bound water and one around 20GHz related to free water. The document ...<|control11|><|separator|>
  93. [93]
    Broadband Dielectric Spectroscopy (BDS) investigation of molecular ...
    May 1, 2022 · It is capable to distinguish the carbohydrates contribution and the different interactions between water and dough components. Moreover, this ...