Fact-checked by Grok 2 weeks ago

Capacitance

Capacitance is the measure of an object's or device's ability to store in relation to an applied difference across it. It is quantitatively defined as the ratio of the stored charge Q to the potential difference V, expressed by the formula C = \frac{Q}{V}, where the SI unit of capacitance is the (F), equivalent to one per volt (C/V). A , the device embodying this property, consists of two conductive plates separated by an insulating material known as a , which enables the storage of energy in an between the plates. In practical terms, the capacitance of a simple parallel-plate capacitor is given by C = \epsilon_0 \frac{A}{d}, where \epsilon_0 is the permittivity of free space ($8.85 \times 10^{-12} F/m), A is the surface area of each plate, and d is the separation distance between them; inserting a dielectric material with relative permittivity \kappa > 1 increases the effective capacitance to C = \kappa \epsilon_0 \frac{A}{d}. The energy stored in a charged capacitor is U = \frac{1}{2} C V^2 or equivalently U = \frac{1}{2} \frac{Q^2}{C}, highlighting its role in temporarily holding electrical energy. Commercial capacitors exhibit capacitances ranging from picofarads (pF) to thousands of farads (F) for supercapacitors, with conventional types typically from pF to microfarads (μF) or millifarads (mF) due to size constraints, and are constructed from materials like ceramics, tantalum, or electrolytes depending on the application. Capacitors play a crucial role in electronic circuits, where they function for signal filtering to remove , energy storage in devices like defibrillators, voltage smoothing in power supplies, and timing in oscillators and memory elements. In series and parallel configurations, equivalent capacitances follow rules analogous to resistors but inverted: for series, \frac{1}{C_{eq}} = \sum \frac{1}{C_i}, and for parallel, C_{eq} = \sum C_i, allowing complex designs. Beyond electronics, capacitance principles underpin phenomena like the behavior of biological membranes and the design of sensors for measuring or .

History

Early Concepts

The early concepts of capacitance emerged in the mid-18th century through experiments demonstrating the storage of , predating formal mathematical definitions. In 1745, Ewald Georg von Kleist in accidentally discovered a method to store by connecting a generator to a nail inserted into a bottle filled with alcohol, which retained the charge until discharged. Independently in the same year, at the University of Leiden in the developed a similar device using a glass jar filled with water and coated externally with metal foil, connected via a brass rod through a stopper; this apparatus, known as the , could store significant charges produced by friction machines and deliver powerful shocks upon discharge. The revolutionized electrical experimentation by providing a reliable means to accumulate and control , shifting observations from fleeting sparks to sustained phenomena. Researchers found that the jar's glass walls acted as an between inner and outer conductors (initially water and foil), with charge separation occurring across the material; multiple jars could be connected in to increase storage or in series to heighten voltage. This setup enabled studies of electrical conduction, , and , influencing theories of as a fluid-like substance. By the late , the device spread across , facilitating demonstrations and quantitative comparisons of electrical effects. Benjamin Franklin advanced these concepts in the 1750s through systematic investigations using Leyden jars, integrating them into his broader theory of electricity. In his 1751 publication Experiments and Observations on Electricity, Franklin described dissectible capacitors—jars that could be assembled and disassembled to reveal charge distribution—and used them to explore conservation of charge, proposing that electricity involved the redistribution of a single fluid rather than creation or destruction. His 1752 kite experiment employed a Leyden jar to capture atmospheric electricity from lightning, confirming its identity with laboratory-generated charge and underscoring the jar's role in safe, portable storage. Franklin's work emphasized practical applications, such as lightning rods, while highlighting the jar's capacity to hold "electrical fire" proportional to its size and materials. Further theoretical progress came from 's private experiments between 1771 and 1781, where he quantified the ability of conductors to store charge relative to their potential difference, effectively inventing the concept of capacitance. Cavendish measured the "quantity of electricity" (charge) and "degree of " (potential) using torsion balances and self-induced shocks from charged objects like plates and spheres, determining that capacitance depended on , such as the area and separation of parallel plates. He also explored dielectric effects, noting how insulators like increased storage capacity compared to air. Although unpublished during his lifetime, these findings were edited and released by in 1879 as The Electrical Researches of the Honourable Henry Cavendish, providing an early mathematical for what Maxwell later termed capacitance: C = \frac{Q}{V}, where Q is charge and V is potential.

Development of Capacitors

The development of capacitors originated in with the independent invention of the by Ewald Georg von Kleist in and in the . This device, consisting of a glass jar coated internally and externally with metal foil and containing an like , was the first to store substantial electrical charge, enabling sustained electrical experiments beyond the limitations of electrostatic generators. The Leyden jar's design demonstrated the principles of capacitance through separated conductive plates and an insulating , laying the foundation for all subsequent technologies. In the late 18th and early 19th centuries, advancements focused on improving form factors and materials for practical use. constructed flat-plate capacitors using sheets separated by metal foil in the 1740s, offering a more compact alternative to jars for demonstrations and early applications. Michael Faraday's work in the 1830s introduced the dielectric constant, quantifying how materials like or enhance charge storage, which facilitated the creation of fixed and variable capacitors for emerging electrical instruments. The early 20th century saw specialization in capacitor types to meet growing electrical demands. In 1897, Pollak patented the , exploiting the anodization of valve metals to form a thin , though initial designs suffered from and were limited to power applications. Samuel Ruben's 1925 patent refined this into the modern aluminum electrolytic structure, enabling compact, high-capacitance devices for radio and circuits, with commercialization following in 1936 by Cornell-Dubilier. capacitors, invented by William Dubilier in 1909, offered for high-frequency radio due to mica's low and . Innovations from the 1920s onward diversified capacitor performance for electronics and power systems. The 1926 introduction of titanium dioxide (rutile) capacitors provided higher permittivity than mica, while the 1941 discovery of barium titanate revolutionized ceramics, leading to multilayer ceramic capacitors (MLCCs) via tape-casting and cofiring processes in the 1970s–1980s, achieving annual production exceeding 10¹² units for consumer devices. Film capacitors advanced with metalized polymer dielectrics, starting from polyethylene terephthalate (PET, patented 1941) and polystyrene (1949), enabling self-healing metallization by Bell Labs in 1954 for reliable, high-voltage applications in audio and power filtering. Electrochemical advancements culminated in double-layer capacitors, patented by in 1957 using porous carbon electrodes to exploit electric double-layer capacitance, though commercialization waited until 1978 by Nippon Electric with capacities up to 9 kF as of 2025 for in hybrids and renewables. These developments transformed capacitors from experimental curiosities into indispensable components, scaling capacitance from microfarads in Leyden jars to kilofarads in modern supercapacitors while prioritizing stability, miniaturization, and application-specific dielectrics.

Fundamentals

Definition

Capacitance is the property of a system of conductors that determines the amount of it can store for a given difference between the conductors. In , it arises from the separation of charges on conductors, leading to an that stores energy. This property is intrinsic to the and materials of the system, independent of the specific charges placed on it./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/08%3A_Capacitance/8.02%3A_Capacitors_and_Capacitance) For a simple consisting of two with charges +Q and -Q, the capacitance C is defined as the ratio of the charge magnitude to the potential difference V across the conductors: C = \frac{Q}{V} This relation holds when the conductors are isolated and the potential is measured relative to each other. The definition extends to a single isolated conductor by considering its potential relative to , where V is the potential of the conductor with respect to a distant reference point./02%3A_Charges_and_Conductors/2.05%3A_Capacitance) The SI unit of capacitance is the farad (F), such that a capacitance of 1 F stores 1 coulomb of charge per volt of potential difference. This unit is named after Michael Faraday and is equivalent to $1 \, \mathrm{F} = 1 \, \mathrm{C/V}. In practice, most capacitors have capacitances in the picofarad to microfarad range due to the large scale of the farad. The capacitance value depends solely on the physical configuration of the conductors and the dielectric medium between them, not on the applied voltage or stored charge./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/08%3A_Capacitance/8.02%3A_Capacitors_and_Capacitance) In broader terms, capacitance characterizes the electrostatic energy storage in the between charged bodies, with the stored energy given by \frac{1}{2} C V^2. This energy perspective underscores capacitance as a measure of the system's response to charge separation, fundamental to applications in and ./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/08%3A_Capacitance/8.02%3A_Capacitors_and_Capacitance)

Units

The unit of capacitance is the , symbolized as F. It is a derived unit in the (), named after the English physicist and chemist (1791–1867), who contributed significantly to the understanding of . The is defined as the capacitance of a between the plates of which there appears a potential difference of 1 volt when it is charged by a of equal to 1 . This corresponds to the relation C = \frac{Q}{V}, where C is capacitance in s, Q is in coulombs, and V is difference in volts. In terms of base units, the has the dimensions \mathrm{kg}^{-1} \mathrm{m}^{-2} \mathrm{s}^{4} \mathrm{A}^{2}, reflecting its derivation from the base units of mass, length, time, and . One represents a substantial capacitance value impractical for most electronic applications, as it implies storing 1 of charge at only 1 volt—a requiring enormous plate areas or specialized materials. Consequently, decimal submultiples defined by prefixes are : the microfarad (μF, $10^{-6} F), nanofarad (, $10^{-9} F), and picofarad (, $10^{-12} F) are commonly employed in circuits, while even smaller values like the femtofarad (fF, $10^{-15} F) appear in high-frequency or nanoscale contexts.

Types of Capacitance

Self-Capacitance

Self-capacitance, also referred to as isolated capacitance, describes the ability of a single to store when isolated from other conductors, with the reference potential taken at or . It is defined as the of the charge Q placed on the conductor to the resulting V relative to , expressed as C = \frac{Q}{V}. This quantity measures how much charge can be accumulated on the conductor before it reaches a specified potential, and it arises naturally in for any isolated conducting body. For simple geometries, self-capacitance can be derived analytically using and the of . Consider an isolated conducting sphere of radius R in . The outside the sphere is equivalent to that of a point charge Q at the center, leading to the potential V = \frac{Q}{4\pi\epsilon_0 R}. Thus, the self-capacitance is C = 4\pi\epsilon_0 R, independent of the charge magnitude but proportional to the sphere's size. This result highlights that larger conductors exhibit greater self-capacitance, as they distribute charge over a larger surface area, reducing the potential for a given charge. In contrast to mutual capacitance, which involves the interaction between two or more conductors, self-capacitance pertains solely to one conductor's interaction with the surrounding space, treating the "second plate" as an imaginary surface at . For non-spherical shapes, such as cylinders or arbitrary objects, exact formulas are more complex and often require solving numerically, but the spherical case serves as a fundamental benchmark. For instance, the self-capacitance of a conducting disk or can be approximated using series expansions or variational methods, typically yielding values on the order of the linear dimensions times \epsilon_0. Self-capacitance plays a key role in analyzing parasitic effects in high-voltage systems and isolated components, where unintended charge accumulation can influence performance.

Mutual Capacitance

Mutual capacitance refers to the electrostatic between two distinct s in a , quantifying the amount of charge induced on one by a potential difference applied to the other. For a simple consisting of two isolated s carrying equal and opposite charges Q and -Q, the mutual capacitance C_m is defined as C_m = \frac{|Q|}{|\Delta V|}, where \Delta V is the potential difference between the conductors. This parameter depends solely on the geometry of the conductors and the properties of the medium between them, such as the permittivity \epsilon./02%3A_Charges_and_Conductors/2.05%3A_Capacitance) In contrast to self-capacitance, which measures a single conductor's ability to hold charge relative to or , mutual capacitance emphasizes the between pairs of conductors. For example, in a parallel-plate configuration, the mutual capacitance approximates C_m = \epsilon_0 \frac{A}{d}, where A is the plate area and d is the separation distance, illustrating how closer proximity increases . This concept is fundamental to devices like capacitors, where mutual capacitance determines the device's charge storage capacity. For systems with more than two conductors, mutual capacitance is incorporated into the capacitance \mathbf{C}, where the charge on the i-th is given by Q_i = \sum_j C_{ij} V_j. Here, the off-diagonal elements C_{ij} (for i \neq j) represent mutual capacitances, which are typically negative, indicating induced charges of opposite sign. The is symmetric (C_{ij} = C_{ji}), a consequence of the reciprocity theorem in , ensuring that the effect of potential on one is symmetric with respect to charge induction on another. This formalism allows for accurate modeling of complex systems, such as transmission lines or integrated circuits, where mutual effects influence overall performance. Mutual capacitance plays a critical role in applications involving parasitic effects, such as in electronic circuits, where unintended coupling between adjacent wires can degrade . In capacitive sensors, like those in touchscreens, mutual capacitance sensing detects changes in the between transmit and receive electrodes caused by a nearby object, enabling precise position detection. These phenomena underscore the importance of geometric design in minimizing or exploiting mutual interactions.

Capacitance Matrix

In , the capacitance provides a complete description of the electrostatic interactions among a of multiple isolated . For a consisting of N , the charge Q_i on the i-th conductor is related to the electric potentials V_j of all conductors by the linear relation \mathbf{Q} = \mathbf{C} \mathbf{V}, where \mathbf{Q} and \mathbf{V} are N-dimensional vectors, and \mathbf{C} is the N \times N capacitance . This encapsulates both self-capacitance effects, where a conductor's own potential induces charge on itself, and mutual capacitance effects, where the potential on one conductor induces charge on others. The diagonal elements C_{ii} of the matrix represent the self-capacitances of each , defined as the charge induced on the i-th when it is held at unit potential while all other are grounded (V_i = 1, V_{j \neq i} = 0). The off-diagonal elements C_{ij} (for i \neq j) are the mutual capacitances, which quantify the charge induced on i due to unit potential on j with i grounded; these elements are typically negative, reflecting the opposite sign of induced charges on neighboring . The capacitance is symmetric, satisfying C_{ij} = C_{ji} for all i, j, a derived from the reciprocity in , which ensures that the response of the system to interchanged charge and potential excitations is identical. This symmetry can also be demonstrated through the invariance of the system's electrostatic under potential swaps. Furthermore, the capacitance matrix is positive semi-definite, guaranteeing that the electrostatic energy stored in the system, W = \frac{1}{2} \mathbf{V}^T \mathbf{C} \mathbf{V} = \frac{1}{2} \mathbf{Q}^T \mathbf{C}^{-1} \mathbf{Q}, is always non-negative for physically admissible charge or potential distributions. The inverse matrix \mathbf{C}^{-1}, known as the elastance matrix or matrix of potential coefficients, relates potentials to charges via \mathbf{V} = \mathbf{C}^{-1} \mathbf{Q}, with its elements P_{ij} representing the potential at i due to unit charge on j while others are grounded. In practice, computing the capacitance matrix for complex geometries often requires numerical methods, such as the method of moments or finite element analysis, as analytical solutions are limited to simple configurations like parallel plates or spherical s. For a two-conductor system, the matrix takes the form \mathbf{C} = \begin{pmatrix} C_{11} & C_{12} \\ C_{12} & C_{22} \end{pmatrix}, where the mutual capacitance between the conductors is -C_{12}, and the effective capacitance when the conductors are connected in parallel or series can be derived from the eigenvalues or submatrices of \mathbf{C}. This formalism extends naturally to larger systems, such as in integrated circuits where parasitic capacitances form a full matrix describing crosstalk and signal integrity.

Components and Phenomena

Capacitors

A is a passive designed to store in an , consisting of at least two conductive electrodes or plates separated by an insulating material. When a voltage is applied across the plates, positive charge accumulates on one plate and negative charge on the other, creating a potential and enabling the device to hold charge Q according to the relation C = \frac{Q}{V}, where C is the capacitance and V is the voltage. In its simplest form, a exploits the principle of capacitance between conductors, blocking while allowing to pass through by repeatedly charging and discharging. The basic construction of a capacitor typically involves two metal plates (such as aluminum or foil) with a interposed between them to prevent conduction while permitting electrostatic influence. For a parallel-plate , the capacitance is calculated as C = \frac{\epsilon_0 A}{d}, where \epsilon_0 = 8.85 \times 10^{-12} F/m is the of free space, A is the plate area, and d is the separation ; this value increases if a with relative \kappa > 1 is used, yielding C = \kappa \frac{\epsilon_0 A}{d}. , such as air (\kappa = 1), paper (\kappa \approx 3.7), or glass (\kappa = 4-6), reduce the strength by a factor of \kappa through , allowing greater charge storage without increasing plate size or reducing separation. Common construction techniques include stacking or rolling plates with dielectric sheets to achieve compact, high-capacitance designs suitable for integration into circuits. Capacitors are classified primarily by their dielectric material and construction method, each offering trade-offs in capacitance value, voltage rating, stability, and size. Ceramic capacitors, the most common type, use a ceramic material like as the , often in multilayer configurations for capacitances from picofarads to microfarads, prized for their small size, low cost, and high across temperatures but limited by lower voltage ratings. Electrolytic capacitors employ an layer formed on a metal ( or ) via , with a or as the , enabling high capacitance (up to tens of thousands of microfarads) in compact volumes for filtering, though they are polarized and degrade over time. Film capacitors feature thin plastic films (e.g., , ) as wound or stacked with metal foil electrodes, providing excellent , low losses, and self-healing properties for audio and timing applications, albeit at higher cost and larger size. Other variants include capacitors, using sheets for precision and high-voltage use in resonant circuits, and air or capacitors, which rely on gas or for tunable, low-loss performance in tuning.

Stray Capacitance

Stray capacitance, often referred to as , is the unintended and unwanted capacitance that exists between adjacent conductors or components in an due to their physical proximity and the presence of a medium, such as air or insulating materials. This capacitance arises from the coupling between elements that are not intentionally designed to form a , including traces on printed circuit boards (PCBs), wire leads, and even between windings in inductors or transformers. In practical terms, it can be modeled as an equivalent in parallel with the intended circuit path, with values typically ranging from picofarads to nanofarads depending on geometry and materials. The primary sources of stray capacitance include the overlap area between parallel conductors and the fringing fields at their edges, which follow the fundamental capacitance formula C = \epsilon \frac{A}{d}, where \epsilon is the , A is the effective area, and d is the separation distance. In high-frequency circuits, such as those operating above 1 MHz, stray capacitance becomes particularly problematic because it introduces low-impedance paths that bypass intended signals, leading to effects like , (), and signal attenuation. For instance, in and RF amplifiers, it can cause voltage spikes, ringing, and reduced by coupling between adjacent traces or layers. Experimental studies on high-frequency transformers have shown that unaccounted stray capacitances can alter dynamic models. To mitigate stray capacitance, designers employ techniques such as increasing the physical separation between conductors to reduce A and increase d, which directly lowers capacitance values. Other strategies include using grounded guard traces or shielding to divert , minimizing trace widths and lengths, and incorporating multilayer layouts with dedicated planes to confine fringing fields. In designs for high-frequency applications, optimizing winding geometry—such as employing single-layer air-core structures—has been shown to predict and reduce stray capacitance compared to multilayer coils, improving self-resonant . These methods are essential in precision applications like interfaces, where stray capacitance can introduce measurement errors if not addressed.

Theory and Calculations

Capacitance of Simple Shapes

The capacitance of simple geometric shapes is derived from fundamental principles of electrostatics, particularly Gauss's law and the relationship between electric field, potential difference, and charge. These calculations assume vacuum or air as the dielectric (permittivity \epsilon_0 = 8.85 \times 10^{-12} \, \mathrm{F/m}) and ideal conductors, providing baseline values that can be adjusted for other materials by replacing \epsilon_0 with \epsilon = \kappa \epsilon_0, where \kappa is the dielectric constant.

Parallel-Plate Capacitor

The most straightforward geometry is the parallel-plate capacitor, consisting of two large, flat conducting plates of area A separated by a small d (with d \ll \sqrt{A}). The E between the plates is uniform and given by E = \sigma / \epsilon_0, where \sigma = Q/A is the surface , assuming the fringing fields at the edges are negligible. The potential difference V is then V = E d = (Q d) / (\epsilon_0 A). Thus, the capacitance C = Q/V simplifies to: C = \frac{\epsilon_0 A}{d}. This formula highlights the direct proportionality to plate area and inverse proportionality to separation, making parallel-plate designs common in practical capacitors. For example, increasing A or decreasing d boosts capacitance, though practical limits arise from dielectric breakdown.

Cylindrical (Coaxial) Capacitor

A cylindrical capacitor features two concentric conducting cylinders: an inner one of radius a and length L, surrounded by an outer one of radius b > a. To find the capacitance, apply Gauss's law to a Gaussian surface (a cylinder of radius r, a < r < b), yielding a radial electric field E(r) = (\lambda)/(2\pi \epsilon_0 r), where \lambda = Q/L is the charge per unit length. The potential difference V is obtained by integrating E from a to b: V = \int_a^b E(r) \, dr = \frac{\lambda}{2\pi \epsilon_0} \ln\left(\frac{b}{a}\right) = \frac{Q \ln(b/a)}{2\pi \epsilon_0 L}. Therefore, the capacitance per unit length (or total for length L) is: C = \frac{2\pi \epsilon_0 L}{\ln(b/a)}. This configuration is useful for high-voltage applications, such as cables, where the logarithmic dependence allows capacitance to remain finite even as b approaches a.

Spherical (Concentric) Capacitor

For a spherical capacitor, two concentric conducting spheres have inner radius a and outer radius b > a. Using on a spherical of radius r (a < r < b), the is E(r) = Q / (4\pi \epsilon_0 r^2). Integrating from a to b gives the potential difference: V = \int_a^b E(r) \, dr = \frac{Q}{4\pi \epsilon_0} \left( \frac{1}{a} - \frac{1}{b} \right) = \frac{Q (b - a)}{4\pi \epsilon_0 a b}. The capacitance is thus: C = \frac{4\pi \epsilon_0 a b}{b - a}. In the limit b \to \infty, this reduces to the self-capacitance of an isolated , C = 4\pi \epsilon_0 a, representing the ability of a single to store charge relative to . Spherical geometries are less common in devices but illustrate radial effects in theoretical models. These derivations rely on to simplify the calculation, enabling exact solutions for ideal cases. Real-world deviations, such as or non-vacuum dielectrics, require numerical methods or approximations, but these formulas provide essential benchmarks for design and analysis.

Energy Storage

Capacitors store by accumulating opposite charges on two conductors separated by an , creating an that holds the . This energy arises from the work done to separate the charges against the attractive electrostatic force. To derive the stored energy, consider charging a capacitor from an initial uncharged state. The work dW required to add an infinitesimal charge dq when the existing charge is q is dW = V \, dq, where V = q / C is the potential difference and C is the capacitance. Integrating from 0 to the final charge Q, W = \int_0^Q \frac{q}{C} \, dq = \frac{1}{2} \frac{Q^2}{C}. This work equals the stored energy U, yielding U = \frac{1}{2} \frac{Q^2}{C}. Substituting Q = C V, where V is the final voltage, gives equivalent forms: U = \frac{1}{2} Q V = \frac{1}{2} C V^2. These expressions hold for any capacitor geometry, as the derivation relies only on the definition of capacitance. The energy is fundamentally stored in the electric field within the capacitor. The energy density u (energy per unit volume) in vacuum is u = \frac{1}{2} \epsilon_0 E^2, where \epsilon_0 is the permittivity of free space and E is the electric field strength. For a parallel-plate capacitor, E = V / d (with plate separation d) is uniform between the plates, so the total energy is U = u \times (volume) = \frac{1}{2} \epsilon_0 E^2 \cdot (A d) = \frac{1}{2} C V^2, matching the earlier result since C = \epsilon_0 A / d. In dielectrics, the energy density becomes \frac{1}{2} \epsilon E^2, where \epsilon = \kappa \epsilon_0 and \kappa is the dielectric constant, accounting for reduced field strength due to polarization. This field-based perspective generalizes to arbitrary charge distributions, where the total electrostatic is the volume of \frac{1}{2} \epsilon_0 E^2 over all space. In practice, capacitors enable rapid energy discharge in applications like power supplies, but energy loss occurs during charging due to in the , though ideal models assume reversible storage.

Applications in Devices

Electronic Devices

Capacitors provide the primary means of implementing capacitance in electronic devices, enabling the storage of and the of signal flow in circuits. They are fundamental components in a wide range of consumer and industrial , including cell phones, computers, and pacemakers, where they store charge for brief power needs and filter signals to ensure reliable operation. In these devices, capacitance allows for rapid charge and discharge, distinguishing it from slower battery-based storage. One key application is in power supply conditioning, where capacitors smooth voltage ripples from rectified sources, converting them to stable for device operation. capacitors, often ceramic types with values around 0.1 μF, are placed close to integrated circuits to absorb high-frequency from switching transients, providing a low-impedance path to and preventing with sensitive . This is critical in devices like microprocessors, where voltage stability directly impacts performance and reliability. Electrolytic capacitors, with higher capacitance up to thousands of μF, handle bulk for low-frequency filtering in power rails. In , capacitors facilitate and functions essential for audio, (RF), and communication circuits. As elements, they block between stages while passing signals, maintaining signal integrity in multi-stage designs like audio preamplifiers. In networks, capacitors combine with resistors or inductors to form high-pass, low-pass, or band-pass configurations; for instance, in RF tuners, variable capacitors adjust resonance to select , enabling radio reception. The capacitive reactance, given by X_C = \frac{1}{2\pi f C}, decreases with frequency, making capacitors ideal for attenuating low frequencies in high-pass filters used in crossovers. Capacitance also underpins timing and in electronic devices, forming RC networks that determine pulse widths or frequencies. In relaxation oscillators, such as those in simple timers or flashing LEDs, the \tau = RC governs charge-discharge cycles, producing periodic outputs for applications like metronomes or warning lights. Supercapacitors, with capacitances exceeding 1 F, extend this to energy backup in portable gadgets, bridging short power interruptions during swaps without in devices like digital cameras. These roles highlight capacitance's versatility in enhancing device efficiency and functionality across .

Semiconductor Devices

In semiconductor devices, capacitance arises primarily from the charge separation in depletion regions of p-n junctions and from the insulating oxide layers in structures. These capacitances influence device performance, including switching speeds, , and power consumption, and are characterized through capacitance-voltage (C-V) measurements that reveal doping profiles, oxide thicknesses, and interface properties. The MOS capacitor, a foundational structure in integrated circuits, consists of a , a thin , and a substrate, typically p-type . Its capacitance varies with applied gate voltage, exhibiting three operational regimes: accumulation, depletion, and inversion. In accumulation (negative gate voltage for p-type substrate), majority carriers (holes) accumulate at the - interface, yielding the maximum capacitance C_{ox} = \frac{\epsilon_{ox} A}{t_{ox}}, where \epsilon_{ox} is the , A is the area, and t_{ox} is the thickness. In depletion, a space-charge layer forms, reducing capacitance to the series combination C = \left( \frac{1}{C_{ox}} + \frac{1}{C_{dep}} \right)^{-1}, with depletion capacitance C_{dep} = \frac{\epsilon_s A}{W} and depletion width W = \sqrt{\frac{2\epsilon_s \phi_s}{q N_A}}, where \epsilon_s is , \phi_s is surface potential, q is electron charge, and N_A is acceptor doping. In strong inversion (positive gate voltage), minority carriers (electrons) form an inversion layer, restoring capacitance near C_{ox} at low frequencies but showing frequency dependence at high frequencies due to minority carrier response time. The C-V curve thus provides critical data for threshold voltage and interface trap density. In p-n junction devices, such as diodes and transistors, capacitance includes depletion (junction) and diffusion components. The depletion capacitance, dominant under reverse bias, stems from the varying width of the charge-depleted region, modeled as C_j = \frac{\epsilon_s A}{W}, where W = \sqrt{\frac{2\epsilon_s (V_{bi} - V)}{q} \left( \frac{1}{N_A} + \frac{1}{N_D} \right)} for an abrupt , with built-in voltage V_{bi}, applied voltage V, and donor doping N_D; thus, C_j \propto (V_{bi} - V)^{-m} where m = 1/2 for abrupt junctions. Diffusion capacitance, prominent in forward bias, arises from stored minority carriers and is given by C_d = \tau g_m, where \tau is and g_m is , scaling with forward current. These capacitances limit high-frequency operation in bipolar junction transistors (BJTs), where base-emitter diffusion capacitance and collector-base capacitance contribute to the f_T = \frac{g_m}{2\pi (C_{\pi} + C_{\mu})}, with C_{\pi} and C_{\mu} as base-emitter and base-collector capacitances, respectively. In MOSFETs, gate capacitance dominates, comprising oxide, overlap, and fringing components, with gate-drain capacitance introducing the Miller effect that amplifies effective input capacitance during switching. Variable capacitance is exploited in varactor (varicap) diodes, specialized p-n junctions designed for tunable capacitance in RF applications like voltage-controlled oscillators. Reverse bias modulates the depletion width, yielding C \propto V^{-1/2} for abrupt junctions or steeper tuning (e.g., V^{-1/3} for hyperabrupt profiles), with typical ranges from 1 pF to 100 pF depending on doping gradients. C-V profiling techniques, using quasi-static or high-frequency measurements, further enable non-destructive characterization of doping profiles via N_d(w) = -\frac{C^3}{q \epsilon_s A^2} \frac{d(1/C^2)}{dV}, essential for in fabrication.

Nanoscale and Advanced Topics

Nanoscale Systems

At the nanoscale, capacitance deviates from classical descriptions due to quantum mechanical effects, particularly in low-dimensional systems where the (DOS) is quantized. , defined as C_q = e^2 \frac{dn}{d\mu} where e is the charge, n is the carrier density, and \mu is the , arises from the finite DOS and acts in series with the geometric capacitance. This effect becomes prominent in structures like or carbon nanotubes, where the low DOS limits charge accumulation, reducing the total effective capacitance. In nanoscale transistors, quantum capacitance influences device performance by modulating the gate-channel coupling, especially when the oxide thickness approaches atomic scales. In electronic devices such as structures and field-effect transistors, size quantization and lead to shifts in and altered carrier injection. For instance, in ultrathin silicon-on-insulator (SOI) transistors gated by single-walled carbon nanotubes, room-temperature quantization features manifest in the transfer characteristics due to Van Hove singularities in the nanotube's one-dimensional , limiting gate charge and impacting drain current. These effects are critical for sub-10 nm devices, where they can reduce the total by up to 20-30% in low- materials, necessitating engineering for optimized performance. Nanoscale capacitor designs exploit two-dimensional materials; for example, dielectric nanocapacitors using metallic borophene electrodes separated by hexagonal (h-BN) achieve high capacitance densities through atomic-scale control of the thickness. In energy storage applications, nanoscale systems enhance capacitance via increased surface area and quantum-enhanced charge screening at interfaces. Nanoporous electrodes, such as with surface areas up to 3100 m²/g, enable specific capacitances exceeding 100 F/g in supercapacitors by maximizing ion adsorption sites. Super-capacitance phenomena originate from mesoscale (0.3-1.8 nm) electrochemical processes, where quantum mechanical Hamiltonians describe charge accumulation, unifying non-faradaic double-layer and faradaic pseudocapacitive mechanisms; examples include and nanostructured , which exhibit enhanced screening due to quantized energy levels. Composites like MnO₂ nanowires on carbon nanotubes deliver 279 F/g at 1 A/g with excellent cycling stability, highlighting the role of nanoscale architecture in boosting energy and power densities. Measuring capacitance at the nanoscale poses challenges from stray fields and quantum fluctuations, addressed by advanced techniques like (SMM). SMM, using modified short-open-load on reference SiO₂ samples, achieves traceable measurements from 0.2 to 10 with 3% uncertainty, mitigating errors from dimensional variations and environmental factors like . For spatially resolved , multifrequency electrostatic force (MFH-EFM) enables high-resolution capacitance mapping at arbitrary frequencies, distinguishing local properties in nanostructures with sub-10 nm precision. These methods are essential for validating models in devices like nanoscale capacitors used in .

Single-Electron Devices

Single-electron devices operate by exploiting the quantization of , enabling the precise and detection of individual s through nanoscale structures where capacitance plays a central role in the phenomenon. In these devices, a small conducting island, such as a metallic or , is isolated by tunnel junctions with capacitances typically in the femtofarad () range, ensuring that the required to add or remove a single —known as the charging —dominates . The charging is given by E_c = \frac{e^2}{2C}, where e is the and C is the total capacitance of the island, including contributions from the tunnel junctions and gate. When E_c \gg k_B T (with k_B Boltzmann's and T ), tunneling is suppressed unless the applied bias or gate voltage compensates the electrostatic , leading to quantized charge states observable at low temperatures below 1 K. This orthodox theory of single- tunneling, developed in the late , provides the foundational framework for device operation, emphasizing as the for charge . The (SET), a cornerstone of this field, consists of two tunnel junctions in series forming , island, and electrodes, with a third electrode capacitively coupled to the island via a C_g. The total island capacitance C_\Sigma = C_1 + C_2 + C_g (where C_1 and C_2 are junction capacitances) determines the voltage periodicity of the device's Coulomb staircase characteristics, with the voltage step \Delta V_g = e / C_g setting the scale for charge quantization. In operation, varying the voltage tunes the island's , allowing sequential single-electron tunneling at specific points where conductance exhibits sharp peaks, enabling amplification of tiny charge signals with charge sensitivity approaching $10^{-5} e / \sqrt{\mathrm{Hz}}. Stray capacitances, often comparable to junction capacitances in fabricated devices, can shift operating thresholds and reduce gain, necessitating careful design to minimize them for high-fidelity single-electron control. Beyond basic transistors, single-electron devices enable metrological applications by leveraging precise capacitive control for standards of capacitance and . For instance, single-electron pumps use cyclic gate voltage modulation to transfer exactly one per cycle, generating a quantized I = n e f (with n cycles and f ), accurate to parts in $10^{10} at cryogenic temperatures, directly linking to capacitance through the pump's and gate capacitances. In capacitance , arrays of SETs or single-electron boxes measure absolute capacitance values by monitoring charge-induced shifts in blockade voltage, achieving uncertainties below 0.1 fF. Fabrication advances, such as silicon-based quantum dots in foundry processes, have demonstrated stable single-electron occupancy in multi-dot arrays, paving the way for scalable elements where inter-dot capacitances dictate coupling strengths. Challenges persist in room-temperature operation due to the need for sub-atttofarad capacitances, though hybrid designs integrating SETs with circuits mitigate this by using the SET's effective input capacitance for charge sensing.

Few-Electron Devices

Few-electron devices encompass nanoscale structures, such as s and double quantum dots fabricated in semiconductor heterostructures like GaAs/AlGaAs or Si/SiGe, where the number of confined electrons is deliberately limited to 1–10 to exploit quantum mechanical effects. In these systems, the capacitance of the confining island (the ) is typically on the order of 1–10 , resulting in a charging energy E_C = \frac{e^2}{2C} of approximately 8–80 meV, which dominates over k_B T at cryogenic temperatures below 1 K, enabling discrete charge quantization and phenomena essential for device functionality. This small capacitance arises from the tiny effective area of the dot (∼10–100 nm²) and thin barriers, making gate-induced charge control highly sensitive, with lever arms (gate-to-dot coupling) often exceeding 0.1e/V. A critical aspect of capacitance in few-electron devices is the distinction between geometric (electrostatic) capacitance C_g, which stores charge classically between the dot and surrounding electrodes, and quantum capacitance C_q = e^2 \frac{dN}{d\mu}, which reflects the compressibility of the electron gas due to its discrete density of states. In the few-electron regime, C_q becomes prominent near charge addition lines, where the chemical potential \mu jumps between discrete levels, leading to negative or divergent quantum capacitance that modulates the total effective capacitance through the series combination \frac{1}{C_\text{tot}} = \frac{1}{C_g} + \frac{1}{C_q}. This effect is pivotal for dispersive charge sensing, where a nearby rf sensor (e.g., a gate electrode) detects shifts in its resonance frequency proportional to changes in C_q during interdot charge transitions, such as from (0,2) to (1,1) states in double dots, with sensitivities down to single-electron changes. For instance, in InAs nanowire quantum dots, quantum capacitance variations have enabled millimeter-wave detection of few-electron states with sub-μs readout times. Capacitance techniques, such as rf reflectometry or direct measurements, are employed to map the landscape of few-electron devices by scanning voltages and observing capacitance peaks at degeneracy points between charge states. These peaks correspond to the addition E_A = E_C + E_\text{orb}, where E_\text{orb} includes orbital level spacing, allowing extraction of parameters like mutual capacitance C_m between coupled dots (typically 0.01–0.1 ) and tunnel coupling t via peak splitting. In silicon-based few-electron double dots, such measurements have quantified interdot capacitances to precision better than 1%, facilitating coherent control for processing. In applications to , few-electron quantum dots serve as spin qubits, with capacitive gates tuning the detuning energy \epsilon = \alpha (V_L - V_R), where \alpha is the lever arm ratio influenced by dot-lead and interdot capacitances. Precise capacitance engineering suppresses charge noise while enabling fast (ns) swap operations via electric dipole , with charging energies ensuring stability against thermal decoherence. Multidimensional quantum , generalizing C_q to and valley , further allows discrimination of qubit states in arrays, as demonstrated in Hubbard-model simulations of two-site dots where C_q matrices reveal singlet-triplet splittings. These devices highlight capacitance as a foundational parameter bridging classical and quantum transport in nanoscale .

Negative Capacitance

Negative capacitance refers to a in certain materials, particularly ferroelectrics, where the incremental change in voltage across the material is negative for a positive change in charge, resulting in dV/dQ < 0. This counterintuitive behavior arises from the thermodynamic properties of ferroelectric materials and has been predicted by Landau's of transitions since the 1940s. In the Landau-Devonshire framework, the density F of a ferroelectric as a function of P is expanded as a power series: F(P) = F_0 + \frac{1}{2} \alpha P^2 + \frac{1}{4} \beta P^4 + \frac{1}{6} \gamma P^6 - E P, where \alpha, \beta, \gamma are temperature-dependent coefficients, E is the electric field, and higher-order terms account for anharmonic effects. Near the phase transition, \alpha < 0, leading to a double-well potential. The differential capacitance C = \frac{dQ}{dV} = \left( \frac{d^2 F}{dP^2} \right)^{-1} becomes negative in regions where \frac{d^2 F}{dP^2} < 0, corresponding to the unstable branches of the polarization-electric field (P-E) hysteresis loop. However, a standalone ferroelectric cannot sustain a stable negative capacitance state due to thermodynamic instability; it requires coupling with a positive capacitor, such as a dielectric, to stabilize the effect through internal voltage amplification. The concept gained renewed interest for nanoscale electronics in 2008, when it was proposed to exploit negative capacitance in ferroelectric-insulator stacks to overcome the fundamental 60 mV/decade limit on the subthreshold swing (S) in field-effect transistors (FETs) at . In such a configuration, the ferroelectric layer provides a negative capacitance that amplifies the gate voltage, effectively boosting the transistor's gate control and enabling steeper switching for lower power dissipation. Experimental evidence of transient negative capacitance was first reported in 2014 using epitaxial ferroelectric thin films, where voltage pulses showed a temporary decrease in voltage across the , consistent with inductance-like during polarization switching. This was followed by direct observation of stabilized negative capacitance in 2015 in a thin-film ferroelectric , demonstrating voltage and capacitance enhancement beyond conventional limits. In nanoscale devices, negative capacitance has been integrated into negative capacitance FETs (NCFETs), where ferroelectric materials like doped HfO₂ are used due to their compatibility with processes and room-temperature operation. These devices achieve sub-60 mV/decade swings, with reported values as low as 10-30 mV/decade in prototypes, potentially reducing energy consumption in logic and applications by factors of 10 or more compared to standard transistors. Challenges include mitigation and scalability, addressed through strain engineering and heterostructure designs that stabilize the negative capacitance region without pinning. Ongoing focuses on integrating negative capacitance into beyond- architectures, such as tunnel FETs and steep-slope switches, to enable energy-efficient at the 1-5 nm scale. In July 2025, Terra Quantum reported the successful fabrication and validation of the first foundry-grade NC-FET, advancing toward integration in and applications.

Measurement

Basic Methods

Basic methods for measuring capacitance rely on fundamental principles of charge storage and impedance, typically categorized into () and () techniques. These approaches are suitable for laboratory settings and educational purposes, providing accurate results for capacitors in the picofarad to microfarad range without requiring specialized high-frequency equipment. DC methods determine capacitance by quantifying the charge accumulated for a known voltage, while AC methods exploit the reactive impedance of the capacitor at a specific . In DC measurement, one common technique involves charging the capacitor through a known resistor and observing the voltage-time response using an oscilloscope. The capacitor is charged from a voltage source V_0 in series with a resistor R, and the time constant \tau is measured as the time for the voltage across the capacitor to reach approximately 63% of V_0 during charging. The capacitance is then calculated as C = \tau / R. This method assumes an ideal RC circuit and is effective for values above 0.1 \muF, with errors minimized by using low-leakage components. For smaller capacitances, the ballistic galvanometer method discharges the charged capacitor through a galvanometer, where the charge Q is found from the area under the current-time curve, yielding C = Q / V with V being the charging voltage. AC bridge methods, such as the De Sauty bridge, offer precise measurements for ideal capacitors by balancing a Wheatstone-like at a low (typically 1 kHz). The bridge consists of two resistive arms (R_1 and R_2) and two capacitive arms (known standard C_s and unknown C_x). Balance occurs when the voltage drop is null, satisfying R_1 / R_2 = C_x / C_s, so C_x = C_s \cdot (R_1 / R_2). A detector like a null or confirms balance by adjusting the variable resistor ratio. This technique achieves accuracies of 0.1-1% for capacitances from 10 pF to 10 \muF but assumes negligible losses; for lossy capacitors, the Schering bridge variant incorporates additional components to account for . Another basic AC approach measures the current through the capacitor driven by a known sinusoidal voltage V = V_0 \sin(\omega t), where the capacitive reactance X_C = 1 / (\omega C) determines the current magnitude I = \omega C V_0. Using an AC voltmeter and ammeter, C is computed as C = I / (\omega V_0), with frequency \omega = 2\pi f set by a signal generator. This method is straightforward for moderate capacitances but requires phase correction for non-ideal behavior and is limited to frequencies below 10 kHz to avoid parasitic effects.

Advanced Techniques

Advanced techniques for capacitance measurement extend the capabilities of basic methods to handle challenges such as ultra-small values (down to femtofarads), high frequencies, significant dielectric losses, high-voltage biases, and complex impedance behaviors in materials and devices. These methods often incorporate , frequency-domain analysis, and specialized circuitry to achieve resolutions better than 0.1% and sensitivities in the attofarad range, enabling applications in , high-precision sensors, and electrochemical systems. Impedance spectroscopy is a widely adopted frequency-domain that applies a small sinusoidal voltage across a range of frequencies and analyzes the resulting current response to determine capacitance from the imaginary component of the complex impedance Z(\omega) = R + jX, where capacitance C is extracted as C = \frac{\text{Im}(1/Z)}{\omega} and \omega = 2\pi f is the . This method excels in distinguishing capacitive contributions from resistive and inductive effects, particularly in lossy dielectrics or electrochemical interfaces, with commercial analyzers achieving measurements from millihertz to megahertz. A dual-frequency variant enhances accuracy by simultaneously exciting at two frequencies and using phase-sensitive detection to isolate components, reducing noise in measurements by up to 50% compared to single-frequency approaches. For small capacitances in the 1 pF to 1 nF range, capacitance-to-digital converters (CDCs) represent a modern integrated approach, where the unknown capacitance C_x is charged from a reference voltage and the accumulated charge is converted to a digital count via sigma-delta modulation or successive approximation. These devices offer high resolution (e.g., 20-bit, corresponding to ~1 aF) and immunity to parasitic capacitances through differential configurations, making them ideal for sensors and applications. Resonance-based methods complement CDCs by embedding C_x in an LC oscillator and measuring the shift in resonant f_r = \frac{1}{2\pi \sqrt{LC_x}}, providing sub-picofarad precision with phase-locked loops for frequency tracking. Capacitance-to-relaxation oscillators further advance this by timing the discharge of C_x through a , yielding linear digital outputs with minimal analog components. In high-voltage or lossy environments, specialized techniques address limitations of standard bridges. A correlation method correlates the test signal with a reference under DC biases up to 3 kV, measuring capacitance with 0.1% accuracy by suppressing distortions. For capacitors exhibiting large losses (high ), externally modulated bridges amplify the signal, improving sensitivity by factors of 10–100 through lock-in amplification and achieving measurements down to 0.01 pF. phase-sensitive detection (PSD) directly demodulates the charge signal into orthogonal components, enabling absolute capacitance readout with signal-to-noise ratios exceeding 100 dB for sensor capacitances below 1 pF. Time-encoded ballistic techniques, meanwhile, quantify C_x by the timing of discrete charge packets in a switched-capacitor chain, suitable for ultra-low-power scenarios with resolutions approaching 0.01 . These techniques often integrate with automated systems for monitoring, such as in pulsed where fast-sampling ADCs capture transient responses to derive capacitance from voltage decay curves, achieving 0.5% accuracy under extreme conditions. Overall, advancements prioritize low amplitudes to minimize perturbation, with cryogenic-compatible variants extending to millikelvin temperatures for quantum characterization.

References

  1. [1]
    8.1 Capacitors and Capacitance – University Physics Volume 2
    The capacitance C of a capacitor is defined as the ratio of the maximum charge Q that can be stored in a capacitor to the applied voltage V across its plates.
  2. [2]
    [PDF] Chapter 5 Capacitance and Dielectrics - MIT
    A capacitor is a device which stores electric charge. Capacitors vary in shape and size, but the basic configuration is two conductors carrying equal but ...
  3. [3]
    Capacitance - Ximera - The Ohio State University
    Capacitance is a constant that relates the amount of charge on a conductor with a potential difference between conductors.Missing: physics | Show results with:physics
  4. [4]
    Capacitors - Engineering and Technology History Wiki
    A capacitor is a device for temporarily storing electric charge. What is considered to be the very first capacitor was called the Leyden jar.
  5. [5]
    Electrical Years: Part 2 | National Museum of American History
    Early explanations of the Leyden experiment suggested that electricity accumulated in water held in the jar. A person holding the jar who touched a wire that ...Missing: capacitance | Show results with:capacitance
  6. [6]
    Lecture 2 - The electrical experiments of Benjamin Franklin
    Leyden Jar capacitors were used to store electrical charges produced by the friction machines. Franklin described a demonstration involving a dissectible leyden ...
  7. [7]
    [PDF] Editing Cavendish: Maxwell and The Electrical Researches of Henry ...
    Apr 27, 2015 · the effect of coatings on the capacitance of plates (anticipating Faraday's discovery of specific inductive capacity), and on the resistance ...<|control11|><|separator|>
  8. [8]
    [PDF] A Brief History of The Development of Classical Electrodynamics
    ca. 1775: Henry Cavendish invents the idea of capacitance and resistance (the latter without any way of measuring current other than the level of personal ...
  9. [9]
    Leyden Jars – 1745 - Magnet Academy
    ### Summary of Leyden Jars (1745)
  10. [10]
    [PDF] Historical Introduction to Capacitor Technology - Zenodo
    The development of multilayer ceramic capacitor (MLCC) fabrication using the tape casting and ceramic-electrode cofiring processes during the 1970s and 1980s ...
  11. [11]
    Electrolytic Capacitors, 1890 to 1925: Early History and Basic Principle
    Aug 7, 2025 · This first of three articles on electrolytic-capacitor technology places the technology in a broad historical context and traces the technology from the 1880s ...
  12. [12]
    The Capacitance Unit Farad - PTB.de
    The capacitance C of a capacitor is the ratio of the charge Q stored in the capacitor to the applied dc voltage U: C = Q/U. In the case of alternating current ( ...
  13. [13]
    [PDF] A Capacitance Standard Based on Counting Electrons
    The capacitance standard uses an electron pump to place electrons on a vacuum-gap capacitor. Measuring the resulting voltage yields the capacitance value.
  14. [14]
    NIST Guide to the SI, Chapter 4: The Two Classes of SI Units and ...
    Jan 28, 2016 · The SI now consists of only two classes of units: base units and derived units. The radian and steradian, which were the two supplementary units, are now ...
  15. [15]
    [PDF] The International System of Units (SI), 2019 Edition
    The farad is the capacitance of a capacitor between the plates of which there appears a potential difference of 1 volt when ...
  16. [16]
    [PDF] 5. Electrostatics III: Capacitances and the Reciprocation Theorem - -
    (Sorry we also use V for volume, but this is standard.) The self-capacitance (or just capacitance) is defined as. / . C Q V. =.Missing: formula | Show results with:formula
  17. [17]
    [PDF] Electromagnetism II: Electricity - Kevin Zhou
    For a single, isolated conductor with charge Q, the self-capacitance is defined as. Q = Cϕ where ϕ is the potential difference between the conductor and ...
  18. [18]
    2.5: Capacitance - Physics LibreTexts
    Mar 5, 2022 · Our goal here is to calculate the distributions of the electric field \(\ \mathbf{E}\) and potential \(\ \phi\) in space, and the distribution ...
  19. [19]
    10. Capacitances - Galileo and Einstein
    It can be understood as a measure of how much charge can be loaded on the conductor before it reaches a given voltage (potential). Experimentally, the ...
  20. [20]
    [PDF] CAPACITANCE EXTRACTION - ResearchGate
    For example, the capacitance of an isolated conductor sphere with radius of R can be calculated as C ¼ 4πε0R. Many conductor interconnect wires are involved ...
  21. [21]
    [PDF] Electromagnetism II, Lecture Notes 4 - MIT OpenCourseWare
    Sep 21, 2012 · Cij is called the capacitance matrix, while Pij is called either the elastance matrix, or simply the reciprocal capacitance matrix. In my ...
  22. [22]
    How to Calculate a Capacitance Matrix in COMSOL Multiphysics
    Jun 7, 2017 · Mutual capacitance, also referred to as parasitic or stray capacitance, is desired or undesired capacitance (a buildup of charge) that ...
  23. [23]
    [PDF] 2 Electrostatics
    iii) Capacitance and the capacitance matrix and energy of system of conductors. For a single metal surface, the charge induced on the surface is proportional ...
  24. [24]
    [PDF] Unit 4-3: Capacitance and Inductance
    Note, the mutual inductance matrix Mij is determined solely by the geometry of the current loops, and not by the values of the currents flowing in them. The ...
  25. [25]
    [PDF] The Maxwell Capacitance Matrix - Fast Field Solvers
    Jun 2, 2020 · The relation between the elements of C (Maxwell capacitance matrix) and the physical capacitances between a set of conductors will now be ...Missing: electrostatics seminal
  26. [26]
    [PDF] EECS 117 - University of California, Berkeley
    But the energy of the system would surely be the same, or W = W. 0 which implies that c12 = c21. Thus the capacitance matrix is symmetric. A symmetric matrix ...
  27. [27]
    [PDF] Chapter 2. Charges and Conductors
    ... electrostatic coupling of the two conductors is evidently very strong. As an ... mutual capacitance between two selected conductors may depend on the ...<|control11|><|separator|>
  28. [28]
    [2007.10251] Capacitance matrix revisited - arXiv
    Jul 20, 2020 · The capacitance matrix relates potentials and charges on a system of conductors. We review and rigorously generalize its properties, block-diagonal structure ...Missing: seminal | Show results with:seminal
  29. [29]
    [PDF] CHAPTER 10: PASSIVE COMPONENTS - Analog Devices
    A capacitor is a passive electronic component that stores energy in the form of an electrostatic field. In its simplest form, a capacitor consists of two ...
  30. [30]
    Introduction to Capacitors, Capacitance and Charge
    Capacitance is the electrical property of a capacitor and is the measure of a capacitors ability to store an electrical charge onto its two plates with the unit ...Missing: authoritative | Show results with:authoritative
  31. [31]
    Types of Capacitor and their Construction - Electronics Tutorials
    Commercial types of capacitors are made from metallic foil interlaced with thin sheets of either paraffin-impregnated paper or Mylar as the dielectric material.
  32. [32]
    Explaining Capacitors and the Different Types | DigiKey
    Jun 2, 2023 · Capacitors are distinguished by the materials used in their construction, and to some extent by their operating mechanism. “Ceramic” capacitors ...
  33. [33]
    Stray Capacitance - an overview | ScienceDirect Topics
    Stray capacitance is defined as the finite capacitance that any wire possesses with respect to ground in a real-world environment, which can introduce ...
  34. [34]
    How to Reduce Stray Capacitance in Electronic Circuits
    Jan 25, 2021 · At high-frequency, elements with stray capacitance tend to act like a short circuit or a virtual resistor. The effects of stray capacitance are ...
  35. [35]
  36. [36]
    Reducing Parasitic Capacitance in PCB Layout - Sierra Circuits
    Feb 23, 2021 · Parasitic capacitance effect in PCBs results in EMI and crosstalk. It can be reduced by following good layout practices.
  37. [37]
    Experimental determination of stray capacitances in high frequency ...
    This paper presents practical techniques for determining stray capacitances in a two-winding high frequency transformer for circuit simulation and computer- ...
  38. [38]
    Addressing Parasitic Inductance and Capacitance in High ...
    Nov 5, 2024 · For example, in power electronics, parasitic effects can lead to voltage spikes, electromagnetic interference, and decreased efficiency.<|separator|>
  39. [39]
  40. [40]
    [PDF] Section 5 High Speed PCB Layout Techniques - Texas Instruments
    To minimize stray capacitance, it is as easy as separating the ground plane away from the signal trace. This can involve increasing th distance on the top layer ...
  41. [41]
    Stray capacitances of single-layer air-core inductors for high ...
    A method for predicting stray capacitances of HF inductors dependent on their geometry is presented. The analysis is performed for inductors made of one layer ...
  42. [42]
    Stray Capacitance: Mitigation Techniques - VSE
    Stray capacitance is a byproduct formed by a dielectric material between two conductive surfaces; it can significantly affect high-speed performance.
  43. [43]
    [PDF] 12. Capacitance of and energy stored in capacitors. Parallel and ...
    Oct 7, 2020 · • Capacitance for parallel-plate geometry: C ≡. Q. V. = c0A d. E. +Q. −Q ... • Capacitance for cylindrical geometry: C ≡. Q. V. = 27c0L ln(b ...
  44. [44]
    [PDF] Chapter 5 Capacitance and Dielectrics
    The capacitance of a (“air–filled”) spherical capacitor is. C = 4π 0 ab. (b − a) . where a and b are the radii of the concentric spherical plates. Here we ...
  45. [45]
    8.3 Energy Stored in a Capacitor – University Physics Volume 2
    W = ∫ 0 W ( Q ) d W = ∫ 0 Q q C d q = 1 2 Q 2 C . Since the geometry of the capacitor has not been specified, this equation holds for any type of capacitor. ...
  46. [46]
    Energy Stored on a Capacitor - HyperPhysics
    Proceeding with the integral, which takes a quadratic form in q, gives a summed energy on the capacitor Q2/2C = CVb2/2 = QVb/2 where the Vb here is the battery ...
  47. [47]
    Energy Stored by Capacitors - Richard Fitzpatrick
    In order to fully charge the capacitor, we must do work against this field, and this work becomes energy stored in the capacitor. Let us calculate this energy.
  48. [48]
    Energy in a Capacitor - Physics
    If ΔV is the final potential difference on the capacitor, and Q is the magnitude of the charge on each plate, the energy stored in the capacitor is: U = 1/2 QΔ ...
  49. [49]
    Ch. 8 Introduction - University Physics Volume 2 | OpenStax
    ### Summary of Applications of Capacitors in Electronic Devices and Circuits
  50. [50]
    [PDF] Capacitor and inductors
    We continue with our analysis of linear circuits by introducing two new passive and linear elements: the capacitor and the inductor.
  51. [51]
    [PDF] MT-101: Decoupling Techniques - Analog Devices
    This is generally done with a combination of electrolytic capacitors (for low frequency decoupling), ceramic capacitors (for high frequency decoupling), and ...
  52. [52]
    Lesson 7. Inductors and Capacitors and Their Uses
    Oct 25, 2021 · Capacitors are used in circuits to smooth voltages, reduce hum, dampen spikes, and to link circuits. For example, a capacitor might be used ...Missing: applications | Show results with:applications
  53. [53]
    Electricity and Magnetism - Capacitance - Molecular Expressions
    Nov 13, 2015 · Capacitors are used in many ways in electronic circuits, such as barriers to direct currents, storing memory in a computer chip, storing a ...
  54. [54]
    Capacitive Reactance is the Complex Impedance of a Capacitor
    The property of a capacitor to change its Capacitive Reactance with frequency, makes the capacitor ideal for use in AC filter circuits, or in DC power supply ...
  55. [55]
    What Is An Oscillator? Everything You Need to Know | Blog | Altium
    Oct 29, 2020 · An RC Oscillator (resistor-capacitor) is a type of feedback oscillator which is built using resistors and capacitors, along with an amplifying ...
  56. [56]
    Energy Storage | Applications | Capacitor Guide - EEPower
    Capacitors are devices which store electrical energy in the form of electrical charge accumulated on their plates.
  57. [57]
    None
    ### Summary of Fundamentals of C-V Measurements in Semiconductors
  58. [58]
    [PDF] MOS Capacitor
    What is the doping type of the semiconductor? (b) If the maximum capacitance of the structure C0 (which is equal to Cox × Area) is. 82 pF and the gate ...
  59. [59]
    [PDF] ECE 340 Lecture 27 : Junction Capacitance
    Oct 29, 2012 · Basically, there are two types of capacitance associated with a p-n junction… The first is junction capacitance: • due to the dipole in the ...
  60. [60]
    [PDF] Bipolar Transistor
    It also reduces the base–collector junction capacitance, which is a good thing. Therefore, the collector doping is typically ten times lighter than base doping.
  61. [61]
    Performance of MOSFETs: Characteristic of Capacitance
    The capacitances (Ciss, Crss, and Coss) of a MOSFET are important parameters that affect its switching characteristics.
  62. [62]
    Variable-capacitance Diodes (Varicap Diodes)
    A variable-capacitance diode is a type of diode that uses the capacitance characteristics of the depletion layer. The depletion layer appears in the p-n ...
  63. [63]
    Quantum capacitance in nanoscale device modeling - AIP Publishing
    Expressions for the “quantum capacitance” are derived, and regimes are discussed in which this concept may be useful in modeling electronic devices.
  64. [64]
    [PDF] Quantum Capacitance in Nanoscale Device Modeling
    Aug 16, 2004 · The concept of “quantum capacitance” was used by Luryi1 in order to develop an equiv- alent circuit model for devices that incorporate a highly ...
  65. [65]
  66. [66]
    Effect of size quantization and quantum capacitance on the ...
    Also, the role of quantum capacitance on the threshold voltages of these nanoscale devices is investigated as a function of channel dimensions and electron ...
  67. [67]
  68. [68]
  69. [69]
  70. [70]
    Nanoscale capacitance spectroscopy based on multifrequency ...
    We have presented a novel method for high-resolution nanoscale capacitance characterization based on multifrequency electrostatic force microscopy, ...
  71. [71]
    The single-electron transistor | Rev. Mod. Phys.
    Jul 1, 1992 · Submicronsize transistors, isolated from their leads by tunnel junctions, turn on and off again every time an electron is added.
  72. [72]
    Single Electron Devices - an overview | ScienceDirect Topics
    Single-electron devices are based on the discrete nature of electric charge that can be transferred from a conducting island (quantum dot or metallic cluster)Missing: seminal | Show results with:seminal
  73. [73]
    Single Electron Transistor: Applications & Problems - ResearchGate
    Aug 7, 2025 · The goal of this paper is to review in brief the basic physics of nanoelectronic device single-electron transistor (SET) as well as to prospect ...
  74. [74]
    Dynamic input capacitance of single-electron transistors and the ...
    We examine the “input capacitance,” CSETT, of a single-electron tunneling (SET) transistor. We note that this quantity is crucial in quantifying the ...
  75. [75]
    Single-Electron Transistors as Ultrasensitive Electrometers
    Likharev, IBM J. Res. Develop. 32, 144 (1988). Article Google Scholar. D.V. Averin and K.K. Likharev, in: Mesoscopic Phenomena in Solids, ed. by B. Altshuler ...Missing: seminal | Show results with:seminal
  76. [76]
    [PDF] Standards of Current and Capacitance Based on Single-Electron ...
    3.1 SET capacitance standard​​ In defining capacitance, we consider a transfer of charge between two conductors. The transferred charge Q causes a potential ...
  77. [77]
    Single-electron operations in a foundry-fabricated array of quantum ...
    Dec 16, 2020 · We demonstrate single-electron occupations in all four quantum dots of a 2 x 2 split-gate silicon device fabricated entirely by 300-mm-wafer foundry processes.
  78. [78]
    Dynamic Single-Electron Transistor Modeling for High-Frequency ...
    A new single-electron transistor (SET) model is proposed, which can represent intrinsic terminal capacitances and transcapacitances.Missing: seminal | Show results with:seminal<|control11|><|separator|>
  79. [79]
    [PDF] arXiv:cond-mat/9711276v1 [cond-mat.mes-hall] 26 Nov 1997
    based few-electron devices. While capacitance and tunneling spectroscopy experiments are becoming available for different dot structures,2,3,4,5,6 it is ...
  80. [80]
    [PDF] arXiv:1403.7561v1 [cond-mat.mes-hall] 28 Mar 2014
    Mar 28, 2014 · and one/few-electron devices, or target simple and few geometries ... In quantum dots, dot-to-gate capacitances are impor- tant ...
  81. [81]
    Dispersive Readout of a Few-Electron Double Quantum Dot with ...
    Jan 25, 2013 · We report the dispersive charge-state readout of a double quantum dot in the few-electron regime using the in situ gate electrodes as sensitive detectors.
  82. [82]
    Quantum-Dot Single-Electron Transistors as Thermoelectric ...
    Oct 7, 2021 · We exploit the rich few-electron physics to develop millimeter-wave nanodetectors employing as a sensing element an InAs/InAs 0.3 P 0.7 quantum-dot nanowire.
  83. [83]
    Single-electron capacitance spectroscopy of a few electron box
    Single-electron capacitance spectroscopy of a few electron box. Author ... The ground state energy in a quantum dot can be measured for an arbitrary ...
  84. [84]
    Theory of multidimensional quantum capacitance and its application ...
    Being the QC dependent on the charge and spin of few-electron systems in tunnel-coupled quantum dots, its measurement can be exploited to discriminate between ...Article Text · INTRODUCTION · MULTIDIMENSIONAL... · APPLICATION TO A...
  85. [85]
    Multi-Dimensional Quantum Capacitance of the Two-Site Hubbard ...
    Observation of Quantum Capacitance in the Cooper-Pair Transistor. Phys ... Few-Electron Double Quantum Dot with Fast rf Gate Sensors. Phys. Rev. Lett ...
  86. [86]
  87. [87]
    Why Do Ferroelectrics Exhibit Negative Capacitance? - PMC - NIH
    Recently, the use of negative capacitance in ferroelectrics was proposed to overcome the fundamental limits of power dissipation in integrated circuits [1].
  88. [88]
    Experimental Observation of Negative Capacitance in Ferroelectrics ...
    Jun 10, 2014 · Effective negative capacitance has been postulated in ferroelectrics because there is a hysteresis in plots of polarization-electric field.
  89. [89]
    Negative capacitance in a ferroelectric capacitor - Nature Materials
    ### Summary of "Negative Capacitance in a Ferroelectric Capacitor" (Nature Materials)
  90. [90]
    Rethinking negative capacitance research | Nature Electronics
    Sep 18, 2020 · Negative capacitance field-effect transistors have been proposed as a route to low-power electronics, but a lack of fundamental understanding limits progress.
  91. [91]
    The ferroelectric field-effect transistor with negative capacitance
    Mar 28, 2022 · Here we put forth an ingenious design for the ferroelectric domain-based field-effect transistor with the stable reversible static negative capacitance.
  92. [92]
    [PDF] LAB 2: Measuring Capacitance - Cloudfront.net
    We reference an analytical solution below. We can measure the capacitance of an element by applying an AC voltage across it while we measure the AC current ...
  93. [93]
    [PDF] Capacitance Measurements
    Capacitance is measured by charging a capacitor, then discharging it through a resistor, and measuring voltage over time to find the time constant.
  94. [94]
    Capacitance measurement: Understand and use the right technique ...
    Aug 22, 2006 · Many low-cost bench and handheld multi-meters make capacitance measurements under the assumption that the current source and capacitor are both ...
  95. [95]
    [PDF] Physics 202 Make-up Lab: Measuring Capacitance With the Bridge ...
    The bridge method uses a circuit with resistors and capacitors. By adjusting the ratio, the unknown capacitance (CX) can be calculated using known values.
  96. [96]
    Capacitance measurement method | IEEE Conference Publication
    This paper presents an automatic measurement method based on the classic Schering bridge applying analog-to-digital (ADC) and digital-to-analog (DAC) ...
  97. [97]
    Measurement Methods for Capacitances in the Range of 1 pF–1 nF
    May 31, 2022 · In this paper, we provide an actual overview of capacitance measurement circuits considering well-known and modern measurement methods.
  98. [98]
    High precision, low excitation capacitance measurement methods ...
    May 20, 2022 · Here, we report one approach that can measure capacitance from mK to room temperature with excellent accuracy.
  99. [99]
    Capacitive Impedance Measurement: Dual-frequency Approach - PMC
    Jun 4, 2019 · This paper presents a sensor to measure complex electrical permittivity based on an alternative approach to amplitude and phase measurement.2. Theory · 3. Methodology · 3.1. Sensor And Measurement...
  100. [100]
    Absolute Capacitance Measurement by Direct Digital Fitting of ...
    The AC excitation method applies an alternating voltage to the capacitor under test using an AC power source and measures the capacitor's response to these ...Absolute Capacitance... · 2.2. 1. Capacitance... · 3. Measurement Program...
  101. [101]
    A high voltage capacitance measurement method based on ...
    Apr 18, 2022 · This paper proposes a capacitance measurement method that can accurately measure the capacitance under a DC bias of 3 kV.
  102. [102]
    Measuring the capacitance of capacitors with large losses
    The utilization of circuits with an external modulation serves to raise the sensitivity of capacitance measurements. The automation of these circuits makes them ...
  103. [103]
    New Approaches to the Direct Measurement of Capacitance
    Aug 6, 2025 · This article reviews two new approaches which are presently being employed – the quadrature P.S.D. technique and time encoded ballistic ...
  104. [104]
    Development of a Capacitance Measurement for Pulsed Magnetic ...
    Mar 28, 2025 · In this study, we develop and implement a capacitance measurement technique optimized for pulsed magnetic fields.Ii Capacitance Measurements... · Ii. 3 Voltage Detection · Iii. 3 Capacitance Versus...