Sie sind auf Seite 1von 19

Capacitor

From Wikipedia, the free encyclopedia

Jump to navigation
Jump to search

This article is missing information about there is no graph of discharge


characteristics showing time constant etc.. Please expand the article to include
this information. Further details may exist on the talk page. (February 2019)

This article is about the device. For the physical phenomenon, see capacitance. For
an overview of various kinds of capacitors, see types of capacitor.
"Capacitive" redirects here. For the term used when referring to touchscreens, see
capacitive sensing.
Capacitor

Type
Passive
Invented
Ewald Georg von Kleist
Electronic symbol

A capacitor is a passive two-terminal electronic component that stores electrical


energy in an electric field. The effect of a capacitor is known as capacitance.
While some capacitance exists between any two electrical conductors in proximity in
a circuit, a capacitor is a component designed to add capacitance to a circuit. The
capacitor was originally known as a condenser or condensator.[1] The original name
is still widely used in many languages, but not commonly in English.
The physical form and construction of practical capacitors vary widely and many
capacitor types are in common use. Most capacitors contain at least two electrical
conductors often in the form of metallic plates or surfaces separated by a
dielectric medium. A conductor may be a foil, thin film, sintered bead of metal, or
an electrolyte. The nonconducting dielectric acts to increase the capacitor's
charge capacity. Materials commonly used as dielectrics include glass, ceramic,
plastic film, paper, mica, air, and oxide layers. Capacitors are widely used as
parts of electrical circuits in many common electrical devices. Unlike a resistor,
an ideal capacitor does not dissipate energy.
When two conductors experience a potential difference, for example, when a
capacitor is attached across a battery, an electric field develops across the
dielectric, causing a net positive charge to collect on one plate and net negative
charge to collect on the other plate. No current actually flows through the
dielectric. However, there is a flow of charge through the source circuit. If the
condition is maintained sufficiently long, the current through the source circuit
ceases. If a time-varying voltage is applied across the leads of the capacitor, the
source experiences an ongoing current due to the charging and discharging cycles of
the capacitor.
Capacitance is defined as the ratio of the electric charge on each conductor to the
potential difference between them. The unit of capacitance in the International
System of Units (SI) is the farad (F), defined as one coulomb per volt (1 C/V).
Capacitance values of typical capacitors for use in general electronics range from
about 1 picofarad (pF) (10-12 F) to about 1 millifarad (mF) (10-3 F).
The capacitance of a capacitor is proportional to the surface area of the plates
(conductors) and inversely related to the gap between them. In practice, the
dielectric between the plates passes a small amount of leakage current. It has an
electric field strength limit, known as the breakdown voltage. The conductors and
leads introduce an undesired inductance and resistance.
Capacitors are widely used in electronic circuits for blocking direct current while
allowing alternating current to pass. In analog filter networks, they smooth the
output of power supplies. In resonant circuits they tune radios to particular
frequencies. In electric power transmission systems, they stabilize voltage and
power flow.[2] The property of energy storage in capacitors was exploited as
dynamic memory in early digital computers.[3]

Contents
1
History
2
Theory of operation
2.1
Overview
2.2
Hydraulic analogy
2.3
Parallel-plate capacitor
2.4
Interleaved Capacitor
2.5
Energy stored in a capacitor
2.6
Current�voltage relation
2.7
DC circuits
2.8
AC circuits
2.9
Laplace circuit analysis (s-domain)
2.10
Circuit analysis
3
Non-ideal behavior
3.1
Breakdown voltage
3.2
Equivalent circuit
3.3
Q factor
3.4
Ripple current
3.5
Capacitance instability
3.6
Current and voltage reversal
3.7
Dielectric absorption
3.8
Leakage
3.9
Electrolytic failure from disuse
4
Capacitor types
4.1
Dielectric materials
4.2
Voltage-dependent capacitors
4.3
Frequency-dependent capacitors
4.4
Styles
5
Capacitor markings
5.1
RKM code
5.2
Historical
6
Applications
6.1
Energy storage
6.2
Digital memory
6.3
Pulsed power and weapons
6.4
Power conditioning
6.4.1
Power factor correction
6.5
Suppression and coupling
6.5.1
Signal coupling
6.5.2
Decoupling
6.5.3
High-pass and low-pass filters
6.5.4
Noise suppression, spikes, and snubbers
6.6
Motor starters
6.7
Signal processing
6.7.1
Tuned circuits
6.8
Sensing
6.9
Oscillators
6.10
Producing light
7
Hazards and safety
8
See also
9
References
10
Bibliography
11
External links
History[edit]

Battery of four Leyden jars in Museum Boerhaave, Leiden, the Netherlands


In October 1745, Ewald Georg von Kleist of Pomerania, Germany, found that charge
could be stored by connecting a high-voltage electrostatic generator by a wire to a
volume of water in a hand-held glass jar.[4] Von Kleist's hand and the water acted
as conductors, and the jar as a dielectric (although details of the mechanism were
incorrectly identified at the time). Von Kleist found that touching the wire
resulted in a powerful spark, much more painful than that obtained from an
electrostatic machine. The following year, the Dutch physicist Pieter van
Musschenbroek invented a similar capacitor, which was named the Leyden jar, after
the University of Leiden where he worked.[5] He also was impressed by the power of
the shock he received, writing, "I would not take a second shock for the kingdom of
France."[6]
Daniel Gralath was the first to combine several jars in parallel to increase the
charge storage capacity.[7] Benjamin Franklin investigated the Leyden jar and came
to the conclusion that the charge was stored on the glass, not in the water as
others had assumed. He also adopted the term "battery",[8][9] (denoting the
increasing of power with a row of similar units as in a battery of cannon),
subsequently applied to clusters of electrochemical cells.[10] Leyden jars were
later made by coating the inside and outside of jars with metal foil, leaving a
space at the mouth to prevent arcing between the foils.[citation needed] The
earliest unit of capacitance was the jar, equivalent to about 1.11 nanofarads.[11]
Leyden jars or more powerful devices employing flat glass plates alternating with
foil conductors were used exclusively up until about 1900, when the invention of
wireless (radio) created a demand for standard capacitors, and the steady move to
higher frequencies required capacitors with lower inductance. More compact
construction methods began to be used, such as a flexible dielectric sheet (like
oiled paper) sandwiched between sheets of metal foil, rolled or folded into a small
package.
Early capacitors were known as condensers, a term that is still occasionally used
today, particularly in high power applications, such as automotive systems. The
term was first used for this purpose by Alessandro Volta in 1782, with reference to
the device's ability to store a higher density of electric charge than was possible
with an isolated conductor.[12][1] The term became deprecated because of the
ambiguous meaning of steam condenser, with capacitor becoming the recommended term
from 1926.[13]
Since the beginning of the study of electricity non conductive materials like
glass, porcelain, paper and mica have been used as insulators. These materials some
decades later were also well-suited for further use as the dielectric for the first
capacitors. Paper capacitors made by sandwiching a strip of impregnated paper
between strips of metal, and rolling the result into a cylinder were commonly used
in the late 19th century; their manufacture started in 1876,[14] and they were used
from the early 20th century as decoupling capacitors in telecommunications
(telephony).
Porcelain was used in the first ceramic capacitors. In the early years of Marconi's
wireless transmitting apparatus porcelain capacitors were used for high voltage and
high frequency application in the transmitters. On the receiver side smaller mica
capacitors were used for resonant circuits. Mica dielectric capacitors were
invented in 1909 by William Dubilier. Prior to World War II, mica was the most
common dielectric for capacitors in the United States.[14]
Charles Pollak (born Karol Pollak), the inventor of the first electrolytic
capacitors, found out that the oxide layer on an aluminum anode remained stable in
a neutral or alkaline electrolyte, even when the power was switched off. In 1896 he
was granted U.S. Patent No. 672,913 for an "Electric liquid capacitor with aluminum
electrodes". Solid electrolyte tantalum capacitors were invented by Bell
Laboratories in the early 1950s as a miniaturized and more reliable low-voltage
support capacitor to complement their newly invented transistor.
With the development of plastic materials by organic chemists during the Second
World War, the capacitor industry began to replace paper with thinner polymer
films. One very early development in film capacitors was described in British
Patent 587,953 in 1944.[14]
Last but not least the electric double-layer capacitor (now Supercapacitors) were
invented. In 1957 H. Becker developed a "Low voltage electrolytic capacitor with
porous carbon electrodes".[14][15][16] He believed that the energy was stored as a
charge in the carbon pores used in his capacitor as in the pores of the etched
foils of electrolytic capacitors. Because the double layer mechanism was not known
by him at the time, he wrote in the patent: "It is not known exactly what is taking
place in the component if it is used for energy storage, but it leads to an
extremely high capacity."
Theory of operation[edit]
Main article: Capacitance
Overview[edit]

Charge separation in a parallel-plate capacitor causes an internal electric field.


A dielectric (orange) reduces the field and increases the capacitance.

A simple demonstration capacitor made of two parallel metal plates, using an air
gap as the dielectric.
A capacitor consists of two conductors separated by a non-conductive region.[17]
The non-conductive region can either be a vacuum or an electrical insulator
material known as a dielectric. Examples of dielectric media are glass, air, paper,
plastic, ceramic, and even a semiconductor depletion region chemically identical to
the conductors. From Coulomb's law a charge on one conductor will exert a force on
the charge carriers within the other conductor, attracting opposite polarity charge
and repelling like polarity charges, thus an opposite polarity charge will be
induced on the surface of the other conductor. The conductors thus hold equal and
opposite charges on their facing surfaces,[18] and the dielectric develops an
electric field.
An ideal capacitor is characterized by a constant capacitance C, in farads in the
SI system of units, defined as the ratio of the positive or negative charge Q on
each conductor to the voltage V between them:[17]
C = Q V {\displaystyle C={\frac {Q}{V}}}

A capacitance of one farad (F) means that one coulomb of charge on each conductor
causes a voltage of one volt across the device.[19] Because the conductors (or
plates) are close together, the opposite charges on the conductors attract one
another due to their electric fields, allowing the capacitor to store more charge
for a given voltage than when the conductors are separated, yielding a larger
capacitance.
In practical devices, charge build-up sometimes affects the capacitor mechanically,
causing its capacitance to vary. In this case, capacitance is defined in terms of
incremental changes:
C = d Q d V {\displaystyle C={\frac {\mathrm {d} Q}{\mathrm {d} V}}}

Hydraulic analogy[edit]

In the hydraulic analogy, a capacitor is analogous to a rubber membrane sealed


inside a pipe� this animation illustrates a membrane being repeatedly stretched and
un-stretched by the flow of water, which is analogous to a capacitor being
repeatedly charged and discharged by the flow of charge
In the hydraulic analogy, charge carriers flowing through a wire are analogous to
water flowing through a pipe. A capacitor is like a rubber membrane sealed inside a
pipe. Water molecules cannot pass through the membrane, but some water can move by
stretching the membrane. The analogy clarifies a few aspects of capacitors:
The current alters the charge on a capacitor, just as the flow of water changes the
position of the membrane. More specifically, the effect of an electric current is
to increase the charge of one plate of the capacitor, and decrease the charge of
the other plate by an equal amount. This is just as when water flow moves the
rubber membrane, it increases the amount of water on one side of the membrane, and
decreases the amount of water on the other side.
The more a capacitor is charged, the larger its voltage drop; i.e., the more it
"pushes back" against the charging current. This is analogous to the more a
membrane is stretched, the more it pushes back on the water.
Charge can flow "through" a capacitor even though no individual electron can get
from one side to the other. This is analogous to water flowing through the pipe
even though no water molecule can pass through the rubber membrane. The flow cannot
continue in the same direction forever; the capacitor experiences dielectric
breakdown, and analogously the membrane will eventually break.
The capacitance describes how much charge can be stored on one plate of a capacitor
for a given "push" (voltage drop). A very stretchy, flexible membrane corresponds
to a higher capacitance than a stiff membrane.
A charged-up capacitor is storing potential energy, analogously to a stretched
membrane.
Parallel-plate capacitor[edit]

Parallel plate capacitor model consists of two conducting plates, each of area A,
separated by a gap of thickness d containing a dielectric.
The simplest model capacitor consists of two thin parallel conductive plates each
with an area of
A {\displaystyle A}
separated by a uniform gap of thickness
d {\displaystyle d}
filled with a dielectric with permittivity
e {\displaystyle \varepsilon }
. It is assumed the gap
d {\displaystyle d}
is much smaller than the dimensions of the plates. This model applies well to many
practical capacitors which are constructed of metal sheets separated by a thin
layer of insulating dielectric, since manufacturers try to keep the dielectric very
uniform in thickness to avoid thin spots which can cause failure of the capacitor.
Since the separation between the plates is uniform over the plate area, the
electric field between the plates
E {\displaystyle E}
is constant, and directed perpendicularly to the plate surface, except for an area
near the edges of the plates where the field decreases because the electric field
lines "bulge" out of the sides of the capacitor. This "fringing field" area is
approximately the same width as the plate separation,
d {\displaystyle d}
, and assuming
d {\displaystyle d}
is small compared to the plate dimensions, it is small enough to be ignored.
Therefore, if a charge of
+ Q {\displaystyle +Q}
is placed on one plate and
- Q {\displaystyle -Q}
on the other plate, the charge on each plate will be spread evenly in a surface
charge layer of constant charge density
s = � Q / A {\displaystyle \sigma =\pm Q/A}
coulombs per square meter, on the inside surface of each plate. From Gauss's law
the magnitude of the electric field between the plates is
E = s / e {\displaystyle E=\sigma /\varepsilon }
. The voltage
V {\displaystyle V}
between the plates is defined as the line integral of the electric field over a
line from one plate to another
V = ? 0 d E ( z ) d z = E d = s e d = Q d e A {\displaystyle V=\int
_{0}^{d}E(z)\,\mathrm {d} z=Ed={\sigma \over \varepsilon }d={Qd \over \varepsilon
A}}
The capacitance is defined as
C = Q / V {\displaystyle C=Q/V}
. Substituting
V {\displaystyle V}
above into this equation
C = e A d {\displaystyle C={\varepsilon A \over d}}

Therefore, in a capacitor the highest capacitance is achieved with a high


permittivity dielectric material, large plate area, and small separation between
the plates.
Since the area
A {\displaystyle A}
of the plates increases with the square of the linear dimensions and the
separation
d {\displaystyle d}
increases linearly, the capacitance scales with the linear dimension of a
capacitor (
C ? L {\displaystyle C\varpropto L}
), or as the cube root of the volume.
A parallel plate capacitor can only store a finite amount of energy before
dielectric breakdown occurs. The capacitor's dielectric material has a dielectric
strength Ud which sets the capacitor's breakdown voltage at V = Vbd = Udd. The
maximum energy that the capacitor can store is therefore
E = 1 2 C V 2 = 1 2 e A d ( U d d ) 2 = 1 2 e A d U d 2 {\displaystyle E={\frac {1}
{2}}CV^{2}={\frac {1}{2}}{\frac {\varepsilon A}{d}}(U_{d}d)^{2}={\frac {1}
{2}}\varepsilon AdU_{d}^{2}}

The maximum energy is a function of dielectric volume, permittivity, and dielectric


strength. Changing the plate area and the separation between the plates while
maintaining the same volume causes no change of the maximum amount of energy that
the capacitor can store, so long as the distance between plates remains much
smaller than both the length and width of the plates. In addition, these equations
assume that the electric field is entirely concentrated in the dielectric between
the plates. In reality there are fringing fields outside the dielectric, for
example between the sides of the capacitor plates, which increase the effective
capacitance of the capacitor. This is sometimes called parasitic capacitance. For
some simple capacitor geometries this additional capacitance term can be calculated
analytically.[20] It becomes negligibly small when the ratios of plate width to
separation and length to separation are large.
Interleaved Capacitor[edit]

The interleaved capacitor can be seen as combination of several parallel connected


capacitors.
For
n {\displaystyle n}
number of plates in a capacitor, the total capacitance would be
C = ? o A d ( n - 1 ) {\displaystyle C=\epsilon _{o}{\frac {A}{d}}(n-1)}

where
C = ? o A / d {\displaystyle C=\epsilon _{o}A/d}
is the capacitance for a single plate and
n {\displaystyle n}
is the number of interleaved plates.
As shown to the figure on the right, the interleaved plates can be seen as parallel
plates connected to each other. With the number of capacitor equal to the number of
the spaces in between the plates. Thus the
( n - 1 ) {\displaystyle (n-1)}
multiplier.
Energy stored in a capacitor[edit]
To increase the charge and voltage on a capacitor, work must be done by an external
power source to move charge from the negative to the positive plate against the
opposing force of the electric field.[21][22] If the voltage on the capacitor is
V {\displaystyle V}
, the work
d W {\displaystyle dW}
required to move a small increment of charge
d q {\displaystyle dq}
from the negative to the positive plate is
d W = V d q {\displaystyle dW=Vdq}
. The energy is stored in the increased electric field between the plates. The
total energy
W {\displaystyle W}
stored in a capacitor (expressed in Joule) is equal to the total work done in
establishing the electric field from an uncharged state.[23][22][21]
W = ? 0 Q V ( q ) d q = ? 0 Q q C d q = 1 2 Q 2 C = 1 2 V Q = 1 2 C V 2
{\displaystyle W=\int _{0}^{Q}V(q)\mathrm {d} q=\int _{0}^{Q}{\frac {q}{C}}\mathrm
{d} q={1 \over 2}{Q^{2} \over C}={1 \over 2}VQ={1 \over 2}CV^{2}}

where
Q {\displaystyle Q}
is the charge stored in the capacitor,
V {\displaystyle V}
is the voltage across the capacitor, and
C {\displaystyle C}
is the capacitance. This potential energy will remain in the capacitor until the
charge is removed. If charge is allowed to move back from the positive to the
negative plate, for example by connecting a circuit with resistance between the
plates, the charge moving under the influence of the electric field will do work on
the external circuit.
If the gap between the capacitor plates
d {\displaystyle d}
is constant, as in the parallel plate model above, the electric field between the
plates will be uniform (neglecting fringing fields) and will have a constant value
E = V / d {\displaystyle E=V/d}
. In this case the stored energy can be calculated from the electric field strength

W = 1 2 C V 2 = 1 2 ? A d ( E d ) 2 = 1 2 ? A d E 2 = 1 2 ? E 2 ( volume of
electric field ) {\displaystyle W={1 \over 2}CV^{2}={1 \over 2}{\epsilon A \over d}
(Ed)^{2}={1 \over 2}\epsilon AdE^{2}={1 \over 2}\epsilon E^{2}({\text{volume of
electric field}})}

The last formula above is equal to the energy density per unit volume in the
electric field multiplied by the volume of field between the plates, confirming
that the energy in the capacitor is stored in its electric field.
Current�voltage relation[edit]
The current I(t) through any component in an electric circuit is defined as the
rate of flow of a charge Q(t) passing through it, but actual
charges�electrons�cannot pass through the dielectric layer of a capacitor. Rather,
one electron accumulates on the negative plate for each one that leaves the
positive plate, resulting in an electron depletion and consequent positive charge
on one electrode that is equal and opposite to the accumulated negative charge on
the other. Thus the charge on the electrodes is equal to the integral of the
current as well as proportional to the voltage, as discussed above. As with any
antiderivative, a constant of integration is added to represent the initial voltage
V(t0). This is the integral form of the capacitor equation:[24]
V ( t ) = Q ( t ) C = 1 C ? t 0 t I ( t ) d t + V ( t 0 ) {\displaystyle
V(t)={\frac {Q(t)}{C}}={\frac {1}{C}}\int _{t_{0}}^{t}I(\tau )\mathrm {d} \tau
+V(t_{0})}

Taking the derivative of this and multiplying by C yields the derivative form:[25]
I ( t ) = d Q ( t ) d t = C d V ( t ) d t {\displaystyle I(t)={\frac {\mathrm {d}
Q(t)}{\mathrm {d} t}}=C{\frac {\mathrm {d} V(t)}{\mathrm {d} t}}}

The dual of the capacitor is the inductor, which stores energy in a magnetic field
rather than an electric field. Its current-voltage relation is obtained by
exchanging current and voltage in the capacitor equations and replacing C with the
inductance L.
DC circuits[edit]
See also: RC circuit

A simple resistor-capacitor circuit demonstrates charging of a capacitor.


A series circuit containing only a resistor, a capacitor, a switch and a constant
DC source of voltage V0 is known as a charging circuit.[26] If the capacitor is
initially uncharged while the switch is open, and the switch is closed at t0, it
follows from Kirchhoff's voltage law that
V 0 = v resistor ( t ) + v capacitor ( t ) = i ( t ) R + 1 C ? t 0 t i ( t ) d t
{\displaystyle V_{0}=v_{\text{resistor}}(t)+v_{\text{capacitor}}(t)=i(t)R+{\frac
{1}{C}}\int _{t_{0}}^{t}i(\tau )\mathrm {d} \tau }

Taking the derivative and multiplying by C, gives a first-order differential


equation:
R C d i ( t ) d t + i ( t ) = 0 {\displaystyle RC{\frac {\mathrm {d} i(t)}{\mathrm
{d} t}}+i(t)=0}

At t = 0, the voltage across the capacitor is zero and the voltage across the
resistor is V0. The initial current is then I(0) =V0/R. With this assumption,
solving the differential equation yields
I ( t ) = V 0 R � e - t t 0 V ( t ) = V 0 ( 1 - e - t t 0 ) Q ( t ) = C � V 0 ( 1 -
e - t t 0 ) {\displaystyle {\begin{aligned}I(t)&={\frac {V_{0}}{R}}\cdot e^{\frac
{-t}{\tau _{0}}}\\V(t)&=V_{0}\left(1-e^{\frac {-t}{\tau _{0}}}\right)\\Q(t)&=C\cdot
V_{0}\left(1-e^{\frac {-t}{\tau _{0}}}\right)\end{aligned}}}

where t0 = RC, the time constant of the system. As the capacitor reaches
equilibrium with the source voltage, the voltages across the resistor and the
current through the entire circuit decay exponentially. In the case of a
discharging capacitor, the capacitor's initial voltage (VCi) replaces V0. The
equations become
I ( t ) = V C i R � e - t t 0 V ( t ) = V C i � e - t t 0 Q ( t ) = C � V C i � e -
t t 0 {\displaystyle {\begin{aligned}I(t)&={\frac {V_{Ci}}{R}}\cdot e^{\frac {-t}
{\tau _{0}}}\\V(t)&=V_{Ci}\cdot e^{\frac {-t}{\tau _{0}}}\\Q(t)&=C\cdot V_{Ci}\cdot
e^{\frac {-t}{\tau _{0}}}\end{aligned}}}

AC circuits[edit]
See also: reactance (electronics) and electrical impedance � Deriving the device-
specific impedances
Impedance, the vector sum of reactance and resistance, describes the phase
difference and the ratio of amplitudes between sinusoidally varying voltage and
sinusoidally varying current at a given frequency. Fourier analysis allows any
signal to be constructed from a spectrum of frequencies, whence the circuit's
reaction to the various frequencies may be found. The reactance and impedance of a
capacitor are respectively
X = - 1 ? C = - 1 2 p f C Z = 1 j ? C = - j ? C = - j 2 p f C {\displaystyle
{\begin{aligned}X&=-{\frac {1}{\omega C}}=-{\frac {1}{2\pi fC}}\\Z&={\frac {1}
{j\omega C}}=-{\frac {j}{\omega C}}=-{\frac {j}{2\pi fC}}\end{aligned}}}
where j is the imaginary unit and ? is the angular frequency of the sinusoidal
signal. The -j phase indicates that the AC voltage V = ZI lags the AC current by
90�: the positive current phase corresponds to increasing voltage as the capacitor
charges; zero current corresponds to instantaneous constant voltage, etc.
Impedance decreases with increasing capacitance and increasing frequency [27]. This
implies that a higher-frequency signal or a larger capacitor results in a lower
voltage amplitude per current amplitude�an AC "short circuit" or AC coupling.
Conversely, for very low frequencies, the reactance is high, so that a capacitor is
nearly an open circuit in AC analysis�those frequencies have been "filtered out".
Capacitors are different from resistors and inductors in that the impedance is
inversely proportional to the defining characteristic; i.e., capacitance.
A capacitor connected to a sinusoidal voltage source causes a displacement current
to flow through it. In the case that the voltage source is V0cos(?t), the
displacement current can be expressed as:
I = C d V d t = - ? C V 0 sin ? ( ? t ) {\displaystyle I=C{\frac {dV}{dt}}=-\omega
{C}{V_{\text{0}}}\sin(\omega t)}

At sin(?t) = -1, the capacitor has a maximum (or peak) current whereby I0 = ?CV0.
The ratio of peak voltage to peak current is due to capacitive reactance (denoted
XC).
X C = V 0 I 0 = V 0 ? C V 0 = 1 ? C {\displaystyle X_{C}={\frac {V_{\text{0}}}
{I_{\text{0}}}}={\frac {V_{\text{0}}}{\omega CV_{\text{0}}}}={\frac {1}{\omega C}}}

XC approaches zero as ? approaches infinity. If XC approaches 0, the capacitor


resembles a short wire that strongly passes current at high frequencies. XC
approaches infinity as ? approaches zero. If XC approaches infinity, the capacitor
resembles an open circuit that poorly passes low frequencies.
The current of the capacitor may be expressed in the form of cosines to better
compare with the voltage of the source:
I = - I 0 sin ? ( ? t ) = I 0 cos ? ( ? t + 90 � ) {\displaystyle I=-{I_{\text{0}}}
{\sin({\omega t}})={I_{\text{0}}}{\cos({\omega t}+{90^{\circ }})}}

In this situation, the current is out of phase with the voltage by +p/2 radians or
+90 degrees, i.e. the current leads the voltage by 90�.
Laplace circuit analysis (s-domain)[edit]
When using the Laplace transform in circuit analysis, the impedance of an ideal
capacitor with no initial charge is represented in the s domain by:
Z ( s ) = 1 s C {\displaystyle Z(s)={\frac {1}{sC}}}

where
C is the capacitance, and
s is the complex frequency.
Circuit analysis[edit]
See also: Series and parallel circuits
For capacitors in parallel

Several capacitors in parallel

Illustration of the parallel connection of two capacitors


Capacitors in a parallel configuration each have the same applied voltage. Their
capacitances add up. Charge is apportioned among them by size. Using the schematic
diagram to visualize parallel plates, it is apparent that each capacitor
contributes to the total surface area.
C e q = ? i C i = C 1 + C 2 + ? + C n {\displaystyle C_{\mathrm {eq} }=\sum
_{i}C_{i}=C_{1}+C_{2}+\cdots +C_{n}}
For capacitors in series

Several capacitors in series

Illustration of the serial connection of two capacitors


Connected in series, the schematic diagram reveals that the separation distance,
not the plate area, adds up. The capacitors each store instantaneous charge build-
up equal to that of every other capacitor in the series. The total voltage
difference from end to end is apportioned to each capacitor according to the
inverse of its capacitance. The entire series acts as a capacitor smaller than any
of its components.
1 C e q = ? i 1 C i = 1 C 1 + 1 C 2 + ? + 1 C n {\displaystyle {\frac {1}
{C_{\mathrm {eq} }}}=\sum _{i}{\frac {1}{C_{i}}}={\frac {1}{C_{1}}}+{\frac {1}
{C_{2}}}+\cdots +{\frac {1}{C_{n}}}}

Capacitors are combined in series to achieve a higher working voltage, for example
for smoothing a high voltage power supply. The voltage ratings, which are based on
plate separation, add up, if capacitance and leakage currents for each capacitor
are identical. In such an application, on occasion, series strings are connected in
parallel, forming a matrix. The goal is to maximize the energy storage of the
network without overloading any capacitor. For high-energy storage with capacitors
in series, some safety considerations must be applied to ensure one capacitor
failing and leaking current does not apply too much voltage to the other series
capacitors.
Series connection is also sometimes used to adapt polarized electrolytic capacitors
for bipolar AC use. See electrolytic capacitor#Designing for reverse bias.
Voltage distribution in parallel-to-series networks.
To model the distribution of voltages from a single charged capacitor
( A ) {\displaystyle \left(A\right)}
connected in parallel to a chain of capacitors in series
( B n ) {\displaystyle \left(B_{\text{n}}\right)}
:
( v o l t s ) A e q = A ( 1 - 1 n + 1 ) ( v o l t s ) B 1..n = A n ( 1 - 1 n + 1 )
A - B = 0 {\displaystyle {\begin{aligned}(volts)A_{\mathrm {eq} }&=A\left(1-{\frac
{1}{n+1}}\right)\\(volts)B_{\text{1..n}}&={\frac {A}{n}}\left(1-{\frac {1}
{n+1}}\right)\\A-B&=0\end{aligned}}}

Note: This is only correct if all capacitance values are equal.


The power transferred in this arrangement is:
P = 1 R � 1 n + 1 A volts ( A farads + B farads ) {\displaystyle P={\frac {1}
{R}}\cdot {\frac {1}{n+1}}A_{\text{volts}}\left(A_{\text{farads}}
+B_{\text{farads}}\right)}

Non-ideal behavior[edit]
Capacitors deviate from the ideal capacitor equation in a number of ways. Some of
these, such as leakage current and parasitic effects are linear, or can be analyzed
as nearly linear, and can be dealt with by adding virtual components to the
equivalent circuit of an ideal capacitor. The usual methods of network analysis can
then be applied [28]. In other cases, such as with breakdown voltage, the effect is
non-linear and ordinary (normal, e.g., linear) network analysis cannot be used, the
effect must be dealt with separately. There is yet another group, which may be
linear but invalidate the assumption in the analysis that capacitance is a
constant. Such an example is temperature dependence. Finally, combined parasitic
effects such as inherent inductance, resistance, or dielectric losses can exhibit
non-uniform behavior at variable frequencies of operation.
Breakdown voltage[edit]
Main article: Breakdown voltage
Above a particular electric field, known as the dielectric strength Eds, the
dielectric in a capacitor becomes conductive. The voltage at which this occurs is
called the breakdown voltage of the device, and is given by the product of the
dielectric strength and the separation between the conductors,[29]
V bd = E ds d {\displaystyle V_{\text{bd}}=E_{\text{ds}}d}

The maximum energy that can be stored safely in a capacitor is limited by the
breakdown voltage. Due to the scaling of capacitance and breakdown voltage with
dielectric thickness, all capacitors made with a particular dielectric have
approximately equal maximum energy density, to the extent that the dielectric
dominates their volume.[30]
For air dielectric capacitors the breakdown field strength is of the order 2�5 MV/m
(or kV/mm); for mica the breakdown is 100�300 MV/m; for oil, 15�25 MV/m; it can be
much less when other materials are used for the dielectric.[31] The dielectric is
used in very thin layers and so absolute breakdown voltage of capacitors is
limited. Typical ratings for capacitors used for general electronics applications
range from a few volts to 1 kV. As the voltage increases, the dielectric must be
thicker, making high-voltage capacitors larger per capacitance than those rated for
lower voltages.
The breakdown voltage is critically affected by factors such as the geometry of the
capacitor conductive parts; sharp edges or points increase the electric field
strength at that point and can lead to a local breakdown. Once this starts to
happen, the breakdown quickly tracks through the dielectric until it reaches the
opposite plate, leaving carbon behind and causing a short (or relatively low
resistance) circuit. The results can be explosive as the short in the capacitor
draws current from the surrounding circuitry and dissipates the energy.[32]
However, in capacitors with particular dielectrics[33][34] and thin metal
electrodes shorts are not formed after breakdown. It happens because a metal melts
or evaporates in a breakdown vicinity, isolating it from the rest of the capacitor.
[35][36]
The usual breakdown route is that the field strength becomes large enough to pull
electrons in the dielectric from their atoms thus causing conduction. Other
scenarios are possible, such as impurities in the dielectric, and, if the
dielectric is of a crystalline nature, imperfections in the crystal structure can
result in an avalanche breakdown as seen in semi-conductor devices. Breakdown
voltage is also affected by pressure, humidity and temperature.[37]
Equivalent circuit[edit]

Two different circuit models of a real capacitor


An ideal capacitor only stores and releases electrical energy, without dissipating
any. In reality, all capacitors have imperfections within the capacitor's material
that create resistance. This is specified as the equivalent series resistance or
ESR of a component. This adds a real component to the impedance:
Z C = Z + R ESR = 1 j ? C + R ESR {\displaystyle
Z_{\text{C}}=Z+R_{\text{ESR}}={\frac {1}{j\omega C}}+R_{\text{ESR}}}

As frequency approaches infinity, the capacitive impedance (or reactance)


approaches zero and the ESR becomes significant. As the reactance becomes
negligible, power dissipation approaches PRMS = VRMS� /RESR.
Similarly to ESR, the capacitor's leads add equivalent series inductance or ESL to
the component. This is usually significant only at relatively high frequencies. As
inductive reactance is positive and increases with frequency, above a certain
frequency capacitance is canceled by inductance. High-frequency engineering
involves accounting for the inductance of all connections and components.
If the conductors are separated by a material with a small conductivity rather than
a perfect dielectric, then a small leakage current flows directly between them. The
capacitor therefore has a finite parallel resistance,[38] and slowly discharges
over time (time may vary greatly depending on the capacitor material and quality).
Q factor[edit]
The quality factor (or Q) of a capacitor is the ratio of its reactance to its
resistance at a given frequency, and is a measure of its efficiency. The higher the
Q factor of the capacitor, the closer it approaches the behavior of an ideal
capacitor.
The Q factor of a capacitor can be found through the following formula:
Q = X C R = 1 ? C R {\displaystyle Q={\frac {X_{C}}{R}}={\frac {1}{\omega CR}}}

where
? {\displaystyle \omega }
is angular frequency,
C {\displaystyle C}
is the capacitance,
X C {\displaystyle X_{C}}
is the capacitive reactance, and
R {\displaystyle R}
is the equivalent series resistance (ESR) of the capacitor.
Ripple current[edit]
Ripple current is the AC component of an applied source (often a switched-mode
power supply) whose frequency may be constant or varying. Ripple current causes
heat to be generated within the capacitor due to the dielectric losses caused by
the changing field strength together with the current flow across the slightly
resistive supply lines or the electrolyte in the capacitor. The equivalent series
resistance (ESR) is the amount of internal series resistance one would add to a
perfect capacitor to model this.
Some types of capacitors, primarily tantalum and aluminum electrolytic capacitors,
as well as some film capacitors have a specified rating value for maximum ripple
current.
Tantalum electrolytic capacitors with solid manganese dioxide electrolyte are
limited by ripple current and generally have the highest ESR ratings in the
capacitor family. Exceeding their ripple limits can lead to shorts and burning
parts.
Aluminum electrolytic capacitors, the most common type of electrolytic, suffer a
shortening of life expectancy at higher ripple currents. If ripple current exceeds
the rated value of the capacitor, it tends to result in explosive failure.
Ceramic capacitors generally have no ripple current limitation and have some of the
lowest ESR ratings.
Film capacitors have very low ESR ratings but exceeding rated ripple current may
cause degradation failures.
Capacitance instability[edit]
The capacitance of certain capacitors decreases as the component ages. In ceramic
capacitors, this is caused by degradation of the dielectric. The type of
dielectric, ambient operating and storage temperatures are the most significant
aging factors, while the operating voltage usually has a smaller effect, i.e.,
usual capacitor design is to minimize voltage coefficient. The aging process may be
reversed by heating the component above the Curie point. Aging is fastest near the
beginning of life of the component, and the device stabilizes over time.[39]
Electrolytic capacitors age as the electrolyte evaporates. In contrast with ceramic
capacitors, this occurs towards the end of life of the component.
Temperature dependence of capacitance is usually expressed in parts per million
(ppm) per �C. It can usually be taken as a broadly linear function but can be
noticeably non-linear at the temperature extremes. The temperature coefficient can
be either positive or negative, sometimes even amongst different samples of the
same type. In other words, the spread in the range of temperature coefficients can
encompass zero.
Capacitors, especially ceramic capacitors, and older designs such as paper
capacitors, can absorb sound waves resulting in a microphonic effect. Vibration
moves the plates, causing the capacitance to vary, in turn inducing AC current.
Some dielectrics also generate piezoelectricity. The resulting interference is
especially problematic in audio applications, potentially causing feedback or
unintended recording. In the reverse microphonic effect, the varying electric field
between the capacitor plates exerts a physical force, moving them as a speaker.
This can generate audible sound, but drains energy and stresses the dielectric and
the electrolyte, if any.
Current and voltage reversal[edit]
Current reversal occurs when the current changes direction. Voltage reversal is the
change of polarity in a circuit. Reversal is generally described as the percentage
of the maximum rated voltage that reverses polarity. In DC circuits, this is
usually less than 100%, often in the range of 0 to 90%, whereas AC circuits
experience 100% reversal.
In DC circuits and pulsed circuits, current and voltage reversal are affected by
the damping of the system. Voltage reversal is encountered in RLC circuits that are
underdamped. The current and voltage reverse direction, forming a harmonic
oscillator between the inductance and capacitance. The current and voltage tends to
oscillate and may reverse direction several times, with each peak being lower than
the previous, until the system reaches an equilibrium. This is often referred to as
ringing. In comparison, critically damped or overdamped systems usually do not
experience a voltage reversal. Reversal is also encountered in AC circuits, where
the peak current is equal in each direction.
For maximum life, capacitors usually need to be able to handle the maximum amount
of reversal that a system may experience. An AC circuit experiences 100% voltage
reversal, while underdamped DC circuits experience less than 100%. Reversal creates
excess electric fields in the dielectric, causes excess heating of both the
dielectric and the conductors, and can dramatically shorten the life expectancy of
the capacitor. Reversal ratings often affect the design considerations for the
capacitor, from the choice of dielectric materials and voltage ratings to the types
of internal connections used.[40]
Dielectric absorption[edit]
Capacitors made with any type of dielectric material show some level of "dielectric
absorption" or "soakage". On discharging a capacitor and disconnecting it, after a
short time it may develop a voltage due to hysteresis in the dielectric. This
effect is objectionable in applications such as precision sample and hold circuits
or timing circuits. The level of absorption depends on many factors, from design
considerations to charging time, since the absorption is a time-dependent process.
However, the primary factor is the type of dielectric material. Capacitors such as
tantalum electrolytic or polysulfone film exhibit relatively high absorption, while
polystyrene or Teflon allow very small levels of absorption.[41] In some capacitors
where dangerous voltages and energies exist, such as in flashtubes, television
sets, and defibrillators, the dielectric absorption can recharge the capacitor to
hazardous voltages after it has been shorted or discharged. Any capacitor
containing over 10 joules of energy is generally considered hazardous, while 50
joules or higher is potentially lethal. A capacitor may regain anywhere from 0.01
to 20% of its original charge over a period of several minutes, allowing a
seemingly safe capacitor to become surprisingly dangerous.[42][43][44][45][46]
Leakage[edit]
Leakage is equivalent to a resistor in parallel with the capacitor. Constant
exposure to heat can cause dielectric breakdown and excessive leakage, a problem
often seen in older vacuum tube circuits, particularly where oiled paper and foil
capacitors were used. In many vacuum tube circuits, interstage coupling capacitors
are used to conduct a varying signal from the plate of one tube to the grid circuit
of the next stage. A leaky capacitor can cause the grid circuit voltage to be
raised from its normal bias setting, causing excessive current or signal distortion
in the downstream tube. In power amplifiers this can cause the plates to glow red,
or current limiting resistors to overheat, even fail. Similar considerations apply
to component fabricated solid-state (transistor) amplifiers, but owing to lower
heat production and the use of modern polyester dielectric barriers this once-
common problem has become relatively rare.
Electrolytic failure from disuse[edit]
Aluminum electrolytic capacitors are conditioned when manufactured by applying a
voltage sufficient to initiate the proper internal chemical state. This state is
maintained by regular use of the equipment. If a system using electrolytic
capacitors is unused for a long period of time it can lose its conditioning.
Sometimes they fail with a short circuit when next operated.
Capacitor types[edit]
Main article: Types of capacitor
Practical capacitors are available commercially in many different forms. The type
of internal dielectric, the structure of the plates and the device packaging all
strongly affect the characteristics of the capacitor, and its applications.
Values available range from very low (picofarad range; while arbitrarily low values
are in principle possible, stray (parasitic) capacitance in any circuit is the
limiting factor) to about 5 kF supercapacitors.
Above approximately 1 microfarad electrolytic capacitors are usually used because
of their small size and low cost compared with other types, unless their relatively
poor stability, life and polarised nature make them unsuitable. Very high capacity
supercapacitors use a porous carbon-based electrode material.
Dielectric materials[edit]

Capacitor materials. From left: multilayer ceramic, ceramic disc, multilayer


polyester film, tubular ceramic, polystyrene, metalized polyester film, aluminum
electrolytic. Major scale divisions are in centimetres.
Most capacitors have a dielectric spacer, which increases their capacitance
compared to air or a vacuum. In order to maximise the charge that a capacitor can
hold, the dielectric material needs to have as high a permittivity as possible,
while also having as high a breakdown voltage as possible. The dielectric also
needs to have as low a loss with frequency as possible.
However, low value capacitors are available with a vacuum between their plates to
allow extremely high voltage operation and low losses. Variable capacitors with
their plates open to the atmosphere were commonly used in radio tuning circuits.
Later designs use polymer foil dielectric between the moving and stationary plates,
with no significant air space between the plates.
Several solid dielectrics are available, including paper, plastic, glass, mica and
ceramic.
Paper was used extensively in older capacitors and offers relatively high voltage
performance. However, paper absorbs moisture, and has been largely replaced by
plastic film capacitors.
Most of the plastic films now used offer better stability and ageing performance
than such older dielectrics such as oiled paper, which makes them useful in timer
circuits, although they may be limited to relatively low operating temperatures and
frequencies, because of the limitations of the plastic film being used. Large
plastic film capacitors are used extensively in suppression circuits, motor start
circuits, and power factor correction circuits.
Ceramic capacitors are generally small, cheap and useful for high frequency
applications, although their capacitance varies strongly with voltage and
temperature and they age poorly. They can also suffer from the piezoelectric
effect. Ceramic capacitors are broadly categorized as class 1 dielectrics, which
have predictable variation of capacitance with temperature or class 2 dielectrics,
which can operate at higher voltage. Modern multilayer ceramics are usually quite
small, but some types have inherently wide value tolerances, microphonic issues,
and are usually physically brittle.
Glass and mica capacitors are extremely reliable, stable and tolerant to high
temperatures and voltages, but are too expensive for most mainstream applications.
Electrolytic capacitors and supercapacitors are used to store small and larger
amounts of energy, respectively, ceramic capacitors are often used in resonators,
and parasitic capacitance occurs in circuits wherever the simple conductor-
insulator-conductor structure is formed unintentionally by the configuration of the
circuit layout.
Electrolytic capacitors use an aluminum or tantalum plate with an oxide dielectric
layer. The second electrode is a liquid electrolyte, connected to the circuit by
another foil plate. Electrolytic capacitors offer very high capacitance but suffer
from poor tolerances, high instability, gradual loss of capacitance especially when
subjected to heat, and high leakage current. Poor quality capacitors may leak
electrolyte, which is harmful to printed circuit boards. The conductivity of the
electrolyte drops at low temperatures, which increases equivalent series
resistance. While widely used for power-supply conditioning, poor high-frequency
characteristics make them unsuitable for many applications. Electrolytic capacitors
suffer from self-degradation if unused for a period (around a year), and when full
power is applied may short circuit, permanently damaging the capacitor and usually
blowing a fuse or causing failure of rectifier diodes. For example, in older
equipment, this may cause arcing in rectifier tubes. They can be restored before
use by gradually applying the operating voltage, often performed on antique vacuum
tube equipment over a period of thirty minutes by using a variable transformer to
supply AC power. The use of this technique may be less satisfactory for some solid
state equipment, which may be damaged by operation below its normal power range,
requiring that the power supply first be isolated from the consuming circuits. Such
remedies may not be applicable to modern high-frequency power supplies as these
produce full output voltage even with reduced input.[citation needed]
Tantalum capacitors offer better frequency and temperature characteristics than
aluminum, but higher dielectric absorption and leakage.[47]

Solid electrolyte, resin-dipped 10 �F 35 V tantalum capacitors. The + sign


indicates the positive lead.
Polymer capacitors (OS-CON, OC-CON, KO, AO) use solid conductive polymer (or
polymerized organic semiconductor) as electrolyte and offer longer life and lower
ESR at higher cost than standard electrolytic capacitors.
A feedthrough capacitor is a component that, while not serving as its main use, has
capacitance and is used to conduct signals through a conductive sheet.
Several other types of capacitor are available for specialist applications.
Supercapacitors store large amounts of energy. Supercapacitors made from carbon
aerogel, carbon nanotubes, or highly porous electrode materials, offer extremely
high capacitance (up to 5 kF as of 2010) and can be used in some applications
instead of rechargeable batteries. Alternating current capacitors are specifically
designed to work on line (mains) voltage AC power circuits. They are commonly used
in electric motor circuits and are often designed to handle large currents, so they
tend to be physically large. They are usually ruggedly packaged, often in metal
cases that can be easily grounded/earthed. They also are designed with direct
current breakdown voltages of at least five times the maximum AC voltage.
Voltage-dependent capacitors[edit]
The dielectric constant for a number of very useful dielectrics changes as a
function of the applied electrical field, for example ferroelectric materials, so
the capacitance for these devices is more complex. For example, in charging such a
capacitor the differential increase in voltage with charge is governed by:
d Q = C ( V ) d V {\displaystyle \ dQ=C(V)\,dV}

where the voltage dependence of capacitance, C(V), suggests that the capacitance is
a function of the electric field strength, which in a large area parallel plate
device is given by e = V/d. This field polarizes the dielectric, which
polarization, in the case of a ferroelectric, is a nonlinear S-shaped function of
the electric field, which, in the case of a large area parallel plate device,
translates into a capacitance that is a nonlinear function of the voltage.[48][49]
Corresponding to the voltage-dependent capacitance, to charge the capacitor to
voltage V an integral relation is found:
Q = ? 0 V C ( V ) d V {\displaystyle Q=\int _{0}^{V}C(V)\,dV\ }

which agrees with Q = CV only when C does not depend on voltage V.


By the same token, the energy stored in the capacitor now is given by
d W = Q d V = [ ? 0 V d V ' C ( V ' ) ] d V . {\displaystyle
dW=Q\,dV=\left[\int _{0}^{V}\ dV'\ C(V')\right]\ dV\ .}

Integrating:
W = ? 0 V d V ? 0 V d V ' C ( V ' ) = ? 0 V d V ' ? V ' V d V C ( V
' ) = ? 0 V d V ' ( V - V ' ) C ( V ' ) , {\displaystyle W=\int _{0}^{V}\
dV\ \int _{0}^{V}\ dV'\ C(V')=\int _{0}^{V}\ dV'\ \int _{V'}^{V}\ dV\ C(V')=\int
_{0}^{V}\ dV'\left(V-V'\right)C(V')\ ,}

where interchange of the order of integration is used.


The nonlinear capacitance of a microscope probe scanned along a ferroelectric
surface is used to study the domain structure of ferroelectric materials.[50]
Another example of voltage dependent capacitance occurs in semiconductor devices
such as semiconductor diodes, where the voltage dependence stems not from a change
in dielectric constant but in a voltage dependence of the spacing between the
charges on the two sides of the capacitor.Template:Sze This effect is intentionally
exploited in diode-like devices known as varicaps.
Frequency-dependent capacitors[edit]
If a capacitor is driven with a time-varying voltage that changes rapidly enough,
at some frequency the polarization of the dielectric cannot follow the voltage. As
an example of the origin of this mechanism, the internal microscopic dipoles
contributing to the dielectric constant cannot move instantly, and so as frequency
of an applied alternating voltage increases, the dipole response is limited and the
dielectric constant diminishes. A changing dielectric constant with frequency is
referred to as dielectric dispersion, and is governed by dielectric relaxation
processes, such as Debye relaxation. Under transient conditions, the displacement
field can be expressed as (see electric susceptibility):
D ( t ) = e 0 ? - 8 t e r ( t - t ' ) E ( t ' ) d t ' , {\displaystyle
{\boldsymbol {D(t)}}=\varepsilon _{0}\int _{-\infty }^{t}\ \varepsilon _{r}(t-t')
{\boldsymbol {E}}(t')\ dt',}

indicating the lag in response by the time dependence of er, calculated in


principle from an underlying microscopic analysis, for example, of the dipole
behavior in the dielectric. See, for example, linear response function.[51][52] The
integral extends over the entire past history up to the present time. A Fourier
transform in time then results in:
D ( ? ) = e 0 e r ( ? ) E ( ? ) , {\displaystyle {\boldsymbol {D}}(\omega )
=\varepsilon _{0}\varepsilon _{r}(\omega ){\boldsymbol {E}}(\omega )\ ,}

where er(?) is now a complex function, with an imaginary part related to absorption
of energy from the field by the medium. See permittivity. The capacitance, being
proportional to the dielectric constant, also exhibits this frequency behavior.
Fourier transforming Gauss's law with this form for displacement field:
I ( ? ) = j ? Q ( ? ) = j ? ? S ? D ( r , ? ) � d S {\displaystyle I(\omega )
=j\omega Q(\omega )=j\omega \oint _{\Sigma }{\boldsymbol {D}}({\boldsymbol
{r}},\ \omega )\cdot d{\boldsymbol {\Sigma }}\ }

= [ G ( ? ) + j ? C ( ? ) ] V ( ? ) = V ( ? ) Z ( ? ) , {\displaystyle
=\left[G(\omega )+j\omega C(\omega )\right]V(\omega )={\frac {V(\omega )}
{Z(\omega )}}\ ,}

where j is the imaginary unit, V(?) is the voltage component at angular


frequency ?, G(?) is the real part of the current, called the conductance, and C(?)
determines the imaginary part of the current and is the capacitance. Z(?) is the
complex impedance.
When a parallel-plate capacitor is filled with a dielectric, the measurement of
dielectric properties of the medium is based upon the relation:
e r ( ? ) = e r ' ( ? ) - j e r ? ( ? ) = 1 j ? Z ( ? ) C 0 = C cmplx ( ? ) C 0 ,
{\displaystyle \varepsilon _{r}(\omega )=\varepsilon '_{r}(\omega )-j\varepsilon
''_{r}(\omega )={\frac {1}{j\omega Z(\omega )C_{0}}}={\frac {C_{\text{cmplx}}
(\omega )}{C_{0}}}\ ,}

where a single prime denotes the real part and a double prime the imaginary part,
Z(?) is the complex impedance with the dielectric present, Ccmplx(?) is the so-
called complex capacitance with the dielectric present, and C0 is the capacitance
without the dielectric.[53][54] (Measurement "without the dielectric" in principle
means measurement in free space, an unattainable goal inasmuch as even the quantum
vacuum is predicted to exhibit nonideal behavior, such as dichroism. For practical
purposes, when measurement errors are taken into account, often a measurement in
terrestrial vacuum, or simply a calculation of C0, is sufficiently accurate.[55])
Using this measurement method, the dielectric constant may exhibit a resonance at
certain frequencies corresponding to characteristic response frequencies
(excitation energies) of contributors to the dielectric constant. These resonances
are the basis for a number of experimental techniques for detecting defects. The
conductance method measures absorption as a function of frequency.[56]
Alternatively, the time response of the capacitance can be used directly, as in
deep-level transient spectroscopy.[57]
Another example of frequency dependent capacitance occurs with MOS capacitors,
where the slow generation of minority carriers means that at high frequencies the
capacitance measures only the majority carrier response, while at low frequencies
both types of carrier respond.[58][59]
At optical frequencies, in semiconductors the dielectric constant exhibits
structure related to the band structure of the solid. Sophisticated modulation
spectroscopy measurement methods based upon modulating the crystal structure by
pressure or by other stresses and observing the related changes in absorption or
reflection of light have advanced our knowledge of these materials.[60]
Styles[edit]

Capacitor packages: SMD ceramic at top left; SMD tantalum at bottom left; through-
hole tantalum at top right; through-hole electrolytic at bottom right. Major scale
divisions are cm.
The arrangement of plates and dielectric has many variations in different styles
depending on the desired ratings of the capacitor. For small values of capacitance
(microfarads and less), ceramic disks use metallic coatings, with wire leads bonded
to the coating. Larger values can be made by multiple stacks of plates and disks.
Larger value capacitors usually use a metal foil or metal film layer deposited on
the surface of a dielectric film to make the plates, and a dielectric film of
impregnated paper or plastic � these are rolled up to save space. To reduce the
series resistance and inductance for long plates, the plates and dielectric are
staggered so that connection is made at the common edge of the rolled-up plates,
not at the ends of the foil or metalized film strips that comprise the plates.
The assembly is encased to prevent moisture entering the dielectric � early radio
equipment used a cardboard tube sealed with wax. Modern paper or film dielectric
capacitors are dipped in a hard thermoplastic. Large capacitors for high-voltage
use may have the roll form compressed to fit into a rectangular metal case, with
bolted terminals and bushings for connections. The dielectric in larger capacitors
is often impregnated with a liquid to improve its properties.

Several axial-lead electrolytic capacitors


Capacitors may have their connecting leads arranged in many configurations, for
example axially or radially. "Axial" means that the leads are on a common axis,
typically the axis of the capacitor's cylindrical body � the leads extend from
opposite ends. Radial leads are rarely aligned along radii of the body's circle, so
the term is conventional. The leads (until bent) are usually in planes parallel to
that of the flat body of the capacitor, and extend in the same direction; they are
often parallel as manufactured.
Small, cheap discoidal ceramic capacitors have existed from the 1930s onward, and
remain in widespread use. After the 1980s, surface mount packages for capacitors
have been widely used. These packages are extremely small and lack connecting
leads, allowing them to be soldered directly onto the surface of printed circuit
boards. Surface mount components avoid undesirable high-frequency effects due to
the leads and simplify automated assembly, although manual handling is made
difficult due to their small size.
Mechanically controlled variable capacitors allow the plate spacing to be adjusted,
for example by rotating or sliding a set of movable plates into alignment with a
set of stationary plates. Low cost variable capacitors squeeze together alternating
layers of aluminum and plastic with a screw. Electrical control of capacitance is
achievable with varactors (or varicaps), which are reverse-biased semiconductor
diodes whose depletion region width varies with applied voltage. They are used in
phase-locked loops, amongst other applications.

Das könnte Ihnen auch gefallen