Circuits
Circuits
R-L-C
Contents
1 Linear circuit 1
1.1 Linear and nonlinear components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Significance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Small signal approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2 LC circuit 3
2.1 Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3 Resonance effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.5 Time domain solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.5.1 Kirchhoff’s laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.5.2 Differential equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.5.3 Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.5.4 Initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.6 Series LC circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.6.1 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.6.2 Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.7 Parallel LC circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7.1 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7.2 Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.8 Laplace solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.9 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.10 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.11 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.12 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 RC circuit 15
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Natural response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3 Complex impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
i
ii CONTENTS
4 RL circuit 25
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 Complex impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2.1 Eigenfunctions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.2.2 Sinusoidal steady state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3 Series circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3.1 Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3.2 Transfer functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3.3 Gain and phase angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3.4 Phasor notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3.5 Impulse response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3.6 Zero-input response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3.7 Frequency domain considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3.8 Time domain considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.4 Parallel circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.5 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.7 Text and image sources, contributors, and licenses . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.7.1 Text . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.7.2 Images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.7.3 Content license . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Chapter 1
Linear circuit
A linear circuit is an electronic circuit in which, for a sinusoidal input voltage of frequency f, any steady-state output
of the circuit (the current through any component, or the voltage between any two points) is also sinusoidal with
frequency f. Note that the output need not be in phase with the input.[1]
An equivalent definition of a linear circuit is that it obeys the superposition principle. This means that the output of
the circuit F(x) when a linear combination of signals ax1 (t) + bx2 (t) is applied to it is equal to the linear combination
of the outputs due to the signals x1 (t) and x2 (t) applied separately:
It is called a linear circuit because the output of such a circuit is a linear function of its inputs. Informally, a linear
circuit is one in which the electronic components' values (such as resistance, capacitance, inductance, gain, etc.) do
not change with the level of voltage or current in the circuit. Linear circuits are important because they can amplify
and process electronic signals without distortion. An example of an electronic device that uses linear circuits is a
sound system.
1.2 Significance
Linear circuits are important because they can process analog signals without introducing intermodulation distor-
tion. This means that separate frequencies in the signal stay separate and do not mix, creating new frequencies
(heterodynes).
They are also easier to understand. Because they obey the superposition principle, linear circuits can be analyzed with
powerful mathematical frequency domain techniques, including Fourier analysis and the Laplace transform. These
also give an intuitive understanding of the qualitative behavior of the circuit, characterizing it using terms such as
gain, phase shift, resonant frequency, bandwidth, Q factor, poles, and zeros. The analysis of a linear circuit can often
be done by hand using a scientific calculator.
In contrast, nonlinear circuits usually do not have closed form solutions. They must be analyzed using approximate
numerical methods by electronic circuit simulation computer programs such as SPICE, if accurate results are desired.
The behavior of such linear circuit elements as resistors, capacitors, and inductors can be specified by a single number
1
2 CHAPTER 1. LINEAR CIRCUIT
(resistance, capacitance, inductance, respectively). In contrast, a nonlinear element's behavior is specified by its
detailed transfer function, which may be given as a graph. So specifying the characteristics of a nonlinear circuit
requires more information than is needed for a linear circuit.
“Linear” circuits and systems form a separate category within electronic manufacturing. Manufacturers of transistors
and integrated circuits often divide their product lines into 'linear' and 'digital' lines. “Linear” here means "analog";
the linear line includes integrated circuits designed to process signals linearly, such as op-amps, audio amplifiers,
and active filters, as well as a variety of signal processing circuits that implement nonlinear analog functions such as
logarithmic amplifiers, analog multipliers, and peak detectors.
1.5 References
[1] Zumbahlen, Hank (2008). Linear circuit design handbook. Newnes. ISBN 0-7506-8703-7.
Chapter 2
LC circuit
i
+
L v C
LC circuit diagram
An LC circuit, also called a resonant circuit, tank circuit, or tuned circuit, is an electric circuit consisting of an
inductor, represented by the letter L, and a capacitor, represented by the letter C, connected together. The circuit can
act as an electrical resonator, an electrical analogue of a tuning fork, storing energy oscillating at the circuit’s resonant
frequency.
LC circuits are used either for generating signals at a particular frequency, or picking out a signal at a particular
frequency from a more complex signal. They are key components in many electronic devices, particularly radio
equipment, used in circuits such as oscillators, filters, tuners and frequency mixers.
An LC circuit is an idealized model since it assumes there is no dissipation of energy due to resistance. Any practical
implementation of an LC circuit will always include loss resulting from small but non-zero resistance within the
components and connecting wires. The purpose of an LC circuit is usually to oscillate with minimal damping, so
the resistance is made as low as possible. While no practical circuit is without losses, it is nonetheless instructive to
study this ideal form of the circuit to gain understanding and physical intuition. For a circuit model incorporating
3
4 CHAPTER 2. LC CIRCUIT
LC circuit (left) consisting of ferrite coil and capacitor used as a tuned circuit in the receiver for a radio clock.
2.1 Terminology
The two-element LC circuit described above is the simplest type of inductor-capacitor network (or LC network).
It is also referred to as a second order LC circuit to distinguish it from more complicated (higher order) LC networks
with more inductors and capacitors. Such LC networks with more than two reactances may have more than one
resonant frequency.
The order of the network is the order of the rational function describing the network in the complex frequency variable
s. Generally, the order is equal to the number of L and C elements in the circuit and in any event cannot exceed this
number.
2.2 Operation
An LC circuit, oscillating at its natural resonant frequency, can store electrical energy. See the animation at right. A
capacitor stores energy in the electric field (E) between its plates, depending on the voltage across it, and an inductor
stores energy in its magnetic field (B), depending on the current through it.
If an inductor is connected across a charged capacitor, current will start to flow through the inductor, building up a
magnetic field around it and reducing the voltage on the capacitor. Eventually all the charge on the capacitor will be
gone and the voltage across it will reach zero. However, the current will continue, because inductors resist changes
in current. The current will begin to charge the capacitor with a voltage of opposite polarity to its original charge.
Due to Faraday’s law, the EMF which drives the current is caused by a decrease in the magnetic field, thus the energy
required to charge the capacitor is extracted from the magnetic field. When the magnetic field is completely dissipated
the current will stop and the charge will again be stored in the capacitor, with the opposite polarity as before. Then
the cycle will begin again, with the current flowing in the opposite direction through the inductor.
The charge flows back and forth between the plates of the capacitor, through the inductor. The energy oscillates
2.3. RESONANCE EFFECT 5
Animated diagram showing the operation of a tuned circuit (LC circuit). The capacitor C stores energy in its electric field E and the
inductor L stores energy in its magnetic field B (green). This jerky animation shows “snapshots” of the circuit at progressive points
in the oscillation. The oscillations are slowed down; in an actual tuned circuit the charge oscillates back and forth tens of thousands
to billions of times per second.
back and forth between the capacitor and the inductor until (if not replenished from an external circuit) internal
resistance makes the oscillations die out. In most applications the tuned circuit is part of a larger circuit which applies
alternating current to it, driving continuous oscillations. If these are at the natural oscillatory frequency (Natural
frequency), resonance will occur. The tuned circuit’s action, known mathematically as a harmonic oscillator, is
similar to a pendulum swinging back and forth, or water sloshing back and forth in a tank; for this reason the circuit
is also called a tank circuit.[1] The natural frequency (that is, the frequency at which it will oscillate when isolated
from any other system, as described above) is determined by the capacitance and inductance values. In typical tuned
circuits in electronic equipment the oscillations are very fast, thousands to billions of times per second.
1
ω0 = √
LC
where L is the inductance in henrys, and C is the capacitance in farads. The angular frequency ω0 has units of radians
per second.
6 CHAPTER 2. LC CIRCUIT
ω0 1
f0 = = √ .
2π 2π LC
LC circuits are often used as filters; the L/C ratio is one of the factors that determines their “Q” and so selectivity. For
a series resonant circuit with a given resistance, the higher the inductance and the lower the capacitance, the narrower
the filter bandwidth. For a parallel resonant circuit the opposite applies. Positive feedback around the tuned circuit
(“regeneration”) can also increase selectivity (see Q multiplier and Regenerative circuit).
Stagger tuning can provide an acceptably wide audio bandwidth, yet good selectivity.
2.4 Applications
The resonance effect of the LC circuit has many important applications in signal processing and communications
systems.
• The most common application of tank circuits is tuning radio transmitters and receivers. For example, when
we tune a radio to a particular station, the LC circuits are set at resonance for that particular carrier frequency.
• A parallel resonant circuit can be used as load impedance in output circuits of RF amplifiers. Due to high
impedance, the gain of amplifier is maximum at resonant frequency.
• Both parallel and series resonant circuits are used in induction heating.
LC circuits behave as electronic resonators, which are a key component in many applications:
• Amplifiers
• Oscillators
• Filters
• Tuners
• Mixers
• Foster-Seeley discriminator
• Contactless cards
• Graphics tablets
VC + VL = 0 .
2.5. TIME DOMAIN SOLUTION 7
Likewise, by Kirchhoff’s current law, the current through the capacitor equals the current through the inductor:
IC = IL .
From the constitutive relations for the circuit elements, we also know that
dIL
VL (t) = L ,
dt
dVC
IC (t) = C .
dt
d2 1
2
I(t) + I(t) = 0 .
dt LC
The parameter ω0 , the resonant angular frequency, is defined as:
1
ω0 = √ .
LC
Using this can simplify the differential equation
d2
I(t) + ω02 I(t) = 0 .
dt2
The associated polynomial is
s2 + ω02 = 0 ;
thus,
s = ±jω0 ,
2.5.3 Solution
Thus, the complete solution to the differential equation is
and can be solved for A and B by considering the initial conditions. Since the exponential is complex, the solution
represents a sinusoidal alternating current. Since the electric current I is a physical quantity, it must be real-valued.
As a result, it can be shown that the constants A and B must be complex conjugates:
A = B∗
8 CHAPTER 2. LC CIRCUIT
Now, let
I0 +jϕ
A= e
2
Therefore,
I0 −jϕ
B= e
2
Next, we can use Euler’s formula to obtain a real sinusoid with amplitude I 0 , angular frequency ω0 = 1/√LC, and
phase angle φ.
Thus, the resulting solution becomes:
and
dI
V (t) = L = −ω0 LI0 sin (ω0 t + ϕ) .
dt
I(0) = I0 cos ϕ .
and
dI
V (0) = L = −ω0 LI0 sin ϕ .
dt
t=0
V = VL + VC
I = IL = IC .
2.6.1 Resonance
Inductive reactance magnitude XL increases as frequency increases while capacitive reactance magnitude XC de-
creases with the increase in frequency. At one particular frequency, these two reactances are equal in magnitude but
opposite in sign; that frequency is called the resonant frequency f 0 for the given circuit.
Hence, at resonance:
2.6. SERIES LC CIRCUIT 9
+
+
v
L L
-
v
+
v
C C
- -
Series LC circuit
10 CHAPTER 2. LC CIRCUIT
XL = −XC
1
ωL = .
ωC
Solving for ω, we have
1
ω = ω0 = √ ,
LC
which is defined as the resonant angular frequency of the circuit. Converting angular frequency (in radians per second)
into frequency (in hertz), one has
ω0 1
f0 = = √ .
2π 2π LC
In a series configuration, XC and XL cancel each other out. In real, rather than idealised components, the current
is opposed, mostly by the resistance of the coil windings. Thus, the current supplied to a series resonant circuit is a
maximum at resonance.
• In the limit as f → f 0 current is maximum. Circuit impedance is minimum. In this state, a circuit is called an
acceptor circuit.
• For f < f 0 , XL ≪ −XC. Hence, the circuit is capacitive.
• For f > f 0 , XL ≫ −XC. Hence, the circuit is inductive.
2.6.2 Impedance
In the series configuration, resonance occurs when the complex electrical impedance of the circuit approaches zero.
First consider the impedance of the series LC circuit. The total impedance is given by the sum of the inductive and
capacitive impedances:
Z = ZL + ZC
Writing the inductive impedance as ZL = jωL and capacitive impedance as ZC = 1/jωC and substituting gives
1
Z(ω) = jωL + .
jωC
Writing this expression under a common denominator gives
( )
ω 2 LC − 1
Z(ω) = j .
ωC
Finally, defining the natural angular frequency as
1
ω0 = √ ,
LC
the impedance becomes
( )
ω 2 − ω02
Z(ω) = jL .
ω
The numerator implies that in the limit as ω → ±ω0 , the total impedance Z will be zero and otherwise non-zero.
Therefore the series LC circuit, when connected in series with a load, will act as a band-pass filter having zero
impedance at the resonant frequency of the LC circuit.
2.7. PARALLEL LC CIRCUIT 11
i(t)
v(t) L C
Parallel LC Circuit
In the parallel configuration, the inductor L and capacitor C are connected in parallel, as shown here. The voltage V
across the open terminals is equal to both the voltage across the inductor and the voltage across the capacitor. The
total current I flowing into the positive terminal of the circuit is equal to the sum of the current flowing through the
inductor and the current flowing through the capacitor:
V = VL = VC
I = IL + IC .
2.7.1 Resonance
When XL equals XC, the reactive branch currents are equal and opposite. Hence they cancel out each other to give
minimum current in the main line. Since total current is minimum, in this state the total impedance is maximum.
The resonant frequency is given by
12 CHAPTER 2. LC CIRCUIT
ω0 1
f0 = = √ .
2π 2π LC
Note that any reactive branch current is not minimum at resonance, but each is given separately by dividing source
voltage (V) by reactance (Z). Hence I = V/Z, as per Ohm’s law.
• At f 0 , the line current is minimum. The total impedance is at the maximum. In this state a circuit is called a
rejector circuit.
2.7.2 Impedance
The same analysis may be applied to the parallel LC circuit. The total impedance is then given by:
ZL ZC
Z= ,
ZL + ZC
and after substitution of ZL and ZC and simplification, gives
ωL
Z(ω) = −j ·
ω 2 LC −1
which further simplifies to
( )( )
1 ω
Z(ω) = −j ,
C ω − ω02
2
where
1
ω0 = √ .
LC
Note that
lim Z(ω) = ∞
ω→±ω0
but for all other values of ω the impedance is finite. The parallel LC circuit connected in series with a load will act
as band-stop filter having infinite impedance at the resonant frequency of the LC circuit. The parallel LC circuit
connected in parallel with a load will act as band-pass filter.
vC (t) = v(t)
2.8. LAPLACE SOLUTION 13
dvC
i(t) = C
dt
di
vL (t) = L
dt
Let the differential equation of LC series be:
di d2 v
vin (t) = vL (t) + vC (t) = L + v = LC 2 + v
dt dt
With initial condition:
{
v(0) = v0
i(0) = i0 = C ∗ v ′ (0) = C ∗ v0′
Let define:
1
ω0 = √
LC
f (t) = ω02 vin (t)
Gives:
d2 v
f (t) = + ω02 v
dt2
Transform with Laplace:
[ ]
d2 v
L [f (t)] = L + ω02 v
dt2
F (s) = s2 V (s) − sv0 − v0′ + ω02 V (s)
sv0 + v0′ + F (s)
V (s) =
s2 + ω02
Then antitransform:
[ ]
v0′ F (s)
v(t) = v0 cos(ω0 t) + sin(ω0 t) + L−1 2
ω0 s + ω02
In case input voltage is Heaviside step function:
U ωf
vin (t) = U sin(ωf t) ⇒ Vin (s) =
s2 + ωf2
[ ] [ ( )] ( )
−1 1 U ωf −1 ω02 U ωf 1 1 ω02 U ωf 1 1
L 2
ω0 2 =L − 2 = 2 sin(ω0 t) − sin(ωf t)
s + ω02 s2 + ωf2 ωf2 − ω02 s2 + ω02 s + ωf2 ωf − ω02 ω0 ωf
( )
v′ ω 2 U ωf 1 1
v(t) = v0 cos(ω0 t) + 0 sin(ω0 t) + 20 sin(ω0 t) − sin(ωf t)
ω0 ωf − ω02 ω0 ωf
14 CHAPTER 2. LC CIRCUIT
2.9 History
The first evidence that a capacitor and inductor could produce electrical oscillations was discovered in 1826 by French
scientist Felix Savary.[2][3] He found that when a Leyden jar was discharged through a wire wound around an iron
needle, sometimes the needle was left magnetized in one direction and sometimes in the opposite direction. He
correctly deduced that this was caused by a damped oscillating discharge current in the wire, which reversed the
magnetization of the needle back and forth until it was too small to have an effect, leaving the needle magnetized
in a random direction. American physicist Joseph Henry repeated Savary’s experiment in 1842 and came to the
same conclusion, apparently independently.[4][5] British scientist William Thomson (Lord Kelvin) in 1853 showed
mathematically that the discharge of a Leyden jar through an inductance should be oscillatory, and derived its resonant
frequency.[2][4][5] British radio researcher Oliver Lodge, by discharging a large battery of Leyden jars through a long
wire, created a tuned circuit with its resonant frequency in the audio range, which produced a musical tone from
the spark when it was discharged.[4] In 1857, German physicist Berend Wilhelm Feddersen photographed the spark
produced by a resonant Leyden jar circuit in a rotating mirror, providing visible evidence of the oscillations.[2][4][5] In
1868, Scottish physicist James Clerk Maxwell calculated the effect of applying an alternating current to a circuit with
inductance and capacitance, showing that the response is maximum at the resonant frequency.[2] The first example of
an electrical resonance curve was published in 1887 by German physicist Heinrich Hertz in his pioneering paper on
the discovery of radio waves, showing the length of spark obtainable from his spark-gap LC resonator detectors as a
function of frequency.[2]
One of the first demonstrations of resonance between tuned circuits was Lodge’s “syntonic jars” experiment around
1889.[2][4] He placed two resonant circuits next to each other, each consisting of a Leyden jar connected to an ad-
justable one-turn coil with a spark gap. When a high voltage from an induction coil was applied to one tuned circuit,
creating sparks and thus oscillating currents, sparks were excited in the other tuned circuit only when the circuits
were adjusted to resonance. Lodge and some English scientists preferred the term "syntony" for this effect, but the
term "resonance" eventually stuck.[2] The first practical use for LC circuits was in the 1890s in spark-gap radio trans-
mitters to allow the receiver and transmitter to be tuned to the same frequency. The first patent for a radio system
that allowed tuning was filed by Lodge in 1897, although the first practical systems were invented in 1900 by Italian
radio pioneer Guglielmo Marconi.[2]
2.11 References
[1] Rao, B. Visvesvara; et al. (2012). Electronic Circuit Analysis. India: Pearson Education India. p. 13.6. ISBN 9332511748.
[2] Blanchard, Julian (October 1941). “The History of Electrical Resonance”. Bell System Technical Journal. U.S.: American
Telephone & Telegraph Co. 20 (4): 415. doi:10.1002/j.1538-7305.1941.tb03608.x. Retrieved 2011-03-29.
[3] Savary, Felix (1827). “Memoirs sur l'Aimentation”. Annales de Chimie et de Physique. Paris: Masson. 34: 5–37.
[4] Kimball, Arthur Lalanne (1917). A College Text-book of Physics (2nd ed.). New York: Henry Hold. pp. 516–517.
[5] Huurdeman, Anton A. (2003). The Worldwide History of Telecommunications. U.S.: Wiley-IEEE. pp. 199–200. ISBN
0-471-20505-2.
RC circuit
A resistor–capacitor circuit (RC circuit), or RC filter or RC network, is an electric circuit composed of resistors
and capacitors driven by a voltage or current source. A first order RC circuit is composed of one resistor and one
capacitor and is the simplest type of RC circuit.
RC circuits can be used to filter a signal by blocking certain frequencies and passing others. The two most common
RC filters are the high-pass filters and low-pass filters; band-pass filters and band-stop filters usually require RLC
filters, though crude ones can be made with RC filters.
3.1 Introduction
There are three basic, linear passive lumped analog circuit components: the resistor (R), the capacitor (C), and the
inductor (L). These may be combined in the RC circuit, the RL circuit, the LC circuit, and the RLC circuit, with the
acronyms indicating which components are used. These circuits, among them, exhibit a large number of important
types of behaviour that are fundamental to much of analog electronics. In particular, they are able to act as passive
filters. This article considers the RC circuit, in both series and parallel forms, as shown in the diagrams below.
dV V
C + = 0,
dt R
where C is the capacitance of capacitor.
Solving this equation for V yields the formula for exponential decay:
V (t) = V0 e− RC ,
t
τ = RC .
15
16 CHAPTER 3. RC CIRCUIT
I---->
C R
RC circuit
The complex impedance, ZC (in ohms) of a capacitor with capacitance C (in farads) is
3.4. SERIES CIRCUIT 17
1
ZC =
sC
The complex frequency s is, in general, a complex number,
s = σ + jω ,
where
σ=0
s = jω .
VR
I
R
Vin C Vc
Series RC circuit
By viewing the circuit as a voltage divider, the voltage across the capacitor is:
18 CHAPTER 3. RC CIRCUIT
1
Cs 1
VC (s) = 1 Vin (s) = Vin (s)
R+ Cs
1 + RCs
and the voltage across the resistor is:
R RCs
VR (s) = 1 Vin (s) = 1 + RCs Vin (s) .
R + Cs
VC (s) 1
HC (s) = = .
Vin (s) 1 + RCs
Similarly, the transfer function from the input to the voltage across the resistor is
VR (s) RCs
HR (s) = = .
Vin (s) 1 + RCs
1
s=− .
RC
In addition, the transfer function for the resistor has a zero located at the origin.
VC (jω) 1
GC = HC (jω) = =√
Vin (jω) 2
1 + (ωRC)
and
VR (jω) ωRC
GR = HR (jω) = =√ ,
Vin (jω) 2
1 + (ωRC)
and
( )
−1 1
ϕR = ∠HR (jω) = tan .
ωRC
3.4. SERIES CIRCUIT 19
These expressions together may be substituted into the usual expression for the phasor representing the output:
VC = GC Vin ejϕC
VR = GR Vin ejϕR .
3.4.3 Current
The current in the circuit is the same everywhere since the circuit is in series:
Vin (s) Cs
I(s) = 1 = 1 + RCs Vin (s) .
R + Cs
1 − t 1 t
hC (t) = e RC u(t) = e− τ u(t) ,
RC τ
where u(t) is the Heaviside step function and τ = RC is the time constant.
Similarly, the impulse response for the resistor voltage is
1 − t 1 t
hR (t) = δ(t) − e RC u(t) = δ(t) − e− τ u(t) ,
RC τ
where δ(t) is the Dirac delta function
GC → 0 and GR → 1 .
As ω → 0:
GC → 1 and GR → 0 .
This shows that, if the output is taken across the capacitor, high frequencies are attenuated (shorted to ground) and
low frequencies are passed. Thus, the circuit behaves as a low-pass filter. If, though, the output is taken across the
resistor, high frequencies are passed and low frequencies are attenuated (since the capacitor blocks the signal as its
frequency approaches 0). In this configuration, the circuit behaves as a high-pass filter.
The range of frequencies that the filter passes is called its bandwidth. The point at which the filter attenuates the
signal to half its unfiltered power is termed its cutoff frequency. This requires that the gain of the circuit be reduced
to
20 CHAPTER 3. RC CIRCUIT
1
GC = GR = √
2
Solving the above equation yields
1 1
ωc = or fc =
RC 2πRC
which is the frequency that the filter will attenuate to half its original power.
Clearly, the phases also depend on frequency, although this effect is less interesting generally than the gain variations.
As ω → 0:
π
ϕC → 0 and ϕR → 90◦ = radians .
2
As ω → ∞:
π
ϕC → −90◦ = − radians and ϕR → 0 .
2
So at DC (0 Hz), the capacitor voltage is in phase with the signal voltage while the resistor voltage leads it by 90°.
As frequency increases, the capacitor voltage comes to have a 90° lag relative to the signal and the resistor voltage
comes to be in-phase with the signal.
The most straightforward way to derive the time domain behaviour is to use the Laplace transforms of the expressions
for VC and VR given above. This effectively transforms jω → s. Assuming a step input (i.e. Vᵢ = 0 before t = 0 and
then Vᵢ = V afterwards):
1
Vin (s) = V ·
s
1 1
VC (s) = V · ·
1 + sRC s
sRC 1
VR (s) = V · · .
1 + sRC s
Partial fractions expansions and the inverse Laplace transform yield:
( )
VC (t) = V 1 − e− RC
t
VR (t) = V e− RC .
t
These equations are for calculating the voltage across the capacitor and resistor respectively while the capacitor is
charging; for discharging, the equations are vice versa. These equations can be rewritten in terms of charge and
current using the relationships C = Q/V and V = IR (see Ohm’s law).
Thus, the voltage across the capacitor tends towards V as time passes, while the voltage across the resistor tends
towards 0, as shown in the figures. This is in keeping with the intuitive point that the capacitor will be charging from
the supply voltage as time passes, and will eventually be fully charged.
These equations show that a series RC circuit has a time constant, usually denoted τ = RC being the time it takes the
voltage across the component to either rise (across the capacitor) or fall (across the resistor) to within 1/e of its final
value. That is, τ is the time it takes VC to reach V(1 − 1/e) and VR to reach V(1/e).
3.4. SERIES CIRCUIT 21
99.3% 98.2%
95.0%
86.5%
63.2%
Vc
0 τ 2τ 3τ 4τ 5τ
Time
Capacitor voltage step-response.
100%
Vr
36.8%
13.5%
5%
1.8% 0.7%
0 τ 2τ 3τ 4τ 5τ
Time
Resistor voltage step-response.
The rate of change is a fractional 1 − 1/e per τ. Thus, in going from t = Nτ to t = (N + 1)τ, the voltage will have
moved about 63.2% of the way from its level at t = Nτ toward its final value. So the capacitor will be charged to
22 CHAPTER 3. RC CIRCUIT
about 63.2% after τ, and essentially fully charged (99.3%) after about 5τ. When the voltage source is replaced with
a short-circuit, with the capacitor fully charged, the voltage across the capacitor drops exponentially with t from V
towards 0. The capacitor will be discharged to about 36.8% after τ, and essentially fully discharged (0.7%) after
about 5τ. Note that the current, I, in the circuit behaves as the voltage across the resistor does, via Ohm’s Law.
These results may also be derived by solving the differential equations describing the circuit:
Vin − VC dVC
=C
R dt
VR = Vin − VC .
The first equation is solved by using an integrating factor and the second follows easily; the solutions are exactly the
same as those obtained via Laplace transforms.
Integrator
1
ω≫ .
RC
This means that the capacitor has insufficient time to charge up and so its voltage is very small. Thus the input voltage
approximately equals the voltage across the resistor. To see this, consider the expression for I given above:
Vin
I= 1 ,
R + jωC
1
ωC ≫ ,
R
so
Vin
I≈
R
which is just Ohm’s Law.
Now,
∫ t
1
VC = I dt ,
C 0
so
∫ t
1
VC ≈ Vin dt ,
RC 0
Differentiator
1
ω≪ .
RC
This means that the capacitor has time to charge up until its voltage is almost equal to the source’s voltage. Considering
the expression for I again, when
1
R≪ ,
ωC
so
Vin
I≈ 1
jωC
I
Vin ≈ = VC .
jωC
Now,
dVC
VR = IR = C R
dt
dVin
VR ≈ RC ,
dt
which is a differentiator across the resistor.
More accurate integration and differentiation can be achieved by placing resistors and capacitors as appropriate on
the input and feedback loop of operational amplifiers (see operational amplifier integrator and operational amplifier
differentiator).
IR IC
Vin Vout
R C
Parallel RC circuit
24 CHAPTER 3. RC CIRCUIT
The parallel RC circuit is generally of less interest than the series circuit. This is largely because the output voltage
Vₒᵤ is equal to the input voltage Vᵢ — as a result, this circuit does not act as a filter on the input signal unless fed by
a current source.
With complex impedances:
Vin
IR =
R
IC = jωCVin .
This shows that the capacitor current is 90° out of phase with the resistor (and source) current. Alternatively, the
governing differential equations may be used:
Vin
IR =
R
dVin
IC = C .
dt
When fed by a current source, the transfer function of a parallel RC circuit is:
Vout R
= .
Iin 1 + sRC
• LC circuit
• RLC circuit
• Electrical network
• List of electronics topics
• Step response
3.7 References
[1] Horowitz and Hill: The Art of Electronics
Chapter 4
RL circuit
A resistor–inductor circuit (RL circuit), or RL filter or RL network, is an electric circuit composed of resistors
and inductors driven by a voltage or current source. A first-order RL circuit is composed of one resistor and one
inductor and is the simplest type of RL circuit.
A first order RL circuit is one of the simplest analogue infinite impulse response electronic filters. It consists of a
resistor and an inductor, either in series driven by a voltage source or in parallel driven by a current source.
4.1 Introduction
The fundamental passive linear circuit elements are the resistor (R), capacitor (C) and inductor (L). These circuit
elements can be combined to form an electrical circuit in four distinct ways: the RC circuit, the RL circuit, the
LC circuit and the RLC circuit with the abbreviations indicating which components are used. These circuits exhibit
important types of behaviour that are fundamental to analogue electronics. In particular, they are able to act as passive
filters. This article considers the RL circuit in both series and parallel as shown in the diagrams.
In practice, however, capacitors (and RC circuits) are usually preferred to inductors since they can be more easily
manufactured and are generally physically smaller, particularly for higher values of components.
Both RC and RL circuits form a single-pole filter. Depending on whether the reactive element (C or L) is in series
with the load, or parallel with the load will dictate whether the filter is low-pass or high-pass.
Frequently RL circuits are used for DC power supplies to RF amplifiers, where the inductor is used to pass DC bias
current and block the RF getting back into the power supply.
This article relies on knowledge of the complex impedance representation of inductors and on knowledge
of the frequency domain representation of signals.
ZL = Ls .
s = σ + jω ,
where
25
26 CHAPTER 4. RL CIRCUIT
4.2.1 Eigenfunctions
The complex-valued eigenfunctions of any linear time-invariant (LTI) system are of the following forms:
From Euler’s formula, the real-part of these eigenfunctions are exponentially-decaying sinusoids:
Sinusoidal steady state is a special case in which the input voltage consists of a pure sinusoid (with no exponential
decay). As a result,
σ=0
s = jω .
Ls
VL (s) = Vin (s) ,
R + Ls
and the voltage across the resistor is:
R
VR (s) = Vin (s) .
R + Ls
4.3.1 Current
The current in the circuit is the same everywhere since the circuit is in series:
Vin (s)
I(s) = .
R + Ls
4.3. SERIES CIRCUIT 27
Series RL circuit
VL (s) Ls
HL (s) = = = GL ejϕL .
Vin (s) R + Ls
Similarly, the transfer function for the resistor is
VR (s) R
HR (s) = = = GR ejϕR .
Vin (s) R + Ls
R
s=− .
L
In addition, the transfer function for the inductor has a zero located at the origin.
VL (ω) ωL
GL = HL (ω) = =√
Vin (ω) 2
R2 + (ωL)
and
VR (ω) R
GR = HR (ω) = =√ ,
Vin (ω) 2
R2 + (ωL)
28 CHAPTER 4. RL CIRCUIT
( )
−1 R
ϕL = ∠HL (s) = tan
ωL
and
( )
−1 ωL
ϕR = ∠HR (s) = tan − .
R
These expressions together may be substituted into the usual expression for the phasor representing the output:
VL = GL Vin ejϕL
VR = GR Vin ejϕR
The impulse response for each voltage is the inverse Laplace transform of the corresponding transfer function. It
represents the response of the circuit to an input voltage consisting of an impulse or Dirac delta function.
The impulse response for the inductor voltage is
R −t R 1 t
hL (t) = δ(t) − e L u(t) = δ(t) − e− τ u(t) ,
L τ
where u(t) is the Heaviside step function and τ = L/R is the time constant.
Similarly, the impulse response for the resistor voltage is
R −t R 1 t
hR (t) = e L u(t) = e− τ u(t) .
L τ
The zero-input response (ZIR), also called the natural response, of an RL circuit describes the behavior of the
circuit after it has reached constant voltages and currents and is disconnected from any power source. It is called the
zero-input response because it requires no input.
The ZIR of an RL circuit is:
These are frequency domain expressions. Analysis of them will show which frequencies the circuits (or filters) pass
and reject. This analysis rests on a consideration of what happens to these gains as the frequency becomes very large
and very small.
As ω → ∞:
4.3. SERIES CIRCUIT 29
GL → 1 and GR → 0 .
As ω → 0:
GL → 0 and GR → 1 .
This shows that, if the output is taken across the inductor, high frequencies are passed and low frequencies are
attenuated (rejected). Thus, the circuit behaves as a high-pass filter. If, though, the output is taken across the resistor,
high frequencies are rejected and low frequencies are passed. In this configuration, the circuit behaves as a low-pass
filter. Compare this with the behaviour of the resistor output in an RC circuit, where the reverse is the case.
The range of frequencies that the filter passes is called its bandwidth. The point at which the filter attenuates the
signal to half its unfiltered power is termed its cutoff frequency. This requires that the gain of the circuit be reduced
to
1
GL = GR = √ .
2
Solving the above equation yields
R R
ωc = rad/s or fc = Hz ,
L 2πL
which is the frequency that the filter will attenuate to half its original power.
Clearly, the phases also depend on frequency, although this effect is less interesting generally than the gain variations.
As ω → 0:
π
ϕL → 90◦ = radians and ϕR → 0 .
2
As ω → ∞:
π
ϕL → 0 and ϕR → −90◦ = − radians .
2
So at DC (0 Hz), the resistor voltage is in phase with the signal voltage while the inductor voltage leads it by 90°. As
frequency increases, the resistor voltage comes to have a 90° lag relative to the signal and the inductor voltage comes
to be in-phase with the signal.
The most straightforward way to derive the time domain behaviour is to use the Laplace transforms of the expressions
for VL and VR given above. This effectively transforms jω → s. Assuming a step input (i.e., Vᵢ = 0 before t = 0 and
then Vᵢ = V afterwards):
1
Vin (s) = V ·
s
sL 1
VL (s) = V · ·
R + sL s
R 1
VR (s) = V · · .
R + sL s
30 CHAPTER 4. RL CIRCUIT
100%
Vr
36.8%
13.5%
5%
1.8% 0.7%
0 τ 2τ 3τ 4τ 5τ
Time
Inductor voltage step-response.
99.3% 98.2%
95.0%
86.5%
63.2%
Vc
0 τ 2τ 3τ 4τ 5τ
Time
Resistor voltage step-response.
VL (t) = V e−t L
R
( )
VR (t) = V 1 − e−t L .
R
Thus, the voltage across the inductor tends towards 0 as time passes, while the voltage across the resistor tends towards
V, as shown in the figures. This is in keeping with the intuitive point that the inductor will only have a voltage across
as long as the current in the circuit is changing — as the circuit reaches its steady-state, there is no further current
change and ultimately no inductor voltage.
These equations show that a series RL circuit has a time constant, usually denoted τ = L/R being the time it takes the
voltage across the component to either fall (across the inductor) or rise (across the resistor) to within 1/e of its final
value. That is, τ is the time it takes VL to reach V(1/e) and VR to reach V(1 − 1/e).
The rate of change is a fractional 1 − 1/e per τ. Thus, in going from t = Nτ to t = (N + 1)τ, the voltage will have
moved about 63% of the way from its level at t = Nτ toward its final value. So the voltage across the inductor will
have dropped to about 37% after τ, and essentially to zero (0.7%) after about 5τ. Kirchhoff’s voltage law implies that
the voltage across the resistor will rise at the same rate. When the voltage source is then replaced with a short-circuit,
the voltage across the resistor drops exponentially with t from V towards 0. The resistor will be discharged to about
37% after τ, and essentially fully discharged (0.7%) after about 5τ. Note that the current, I, in the circuit behaves as
the voltage across the resistor does, via Ohm’s Law.
The delay in the rise or fall time of the circuit is in this case caused by the back-EMF from the inductor which, as
the current flowing through it tries to change, prevents the current (and hence the voltage across the resistor) from
rising or falling much faster than the time-constant of the circuit. Since all wires have some self-inductance and
resistance, all circuits have a time constant. As a result, when the power supply is switched on, the current does not
instantaneously reach its steady-state value, V/R. The rise instead takes several time-constants to complete. If this
were not the case, and the current were to reach steady-state immediately, extremely strong inductive electric fields
would be generated by the sharp change in the magnetic field — this would lead to breakdown of the air in the circuit
and electric arcing, probably damaging components (and users).
These results may also be derived by solving the differential equation describing the circuit:
dI
Vin = IR + L
dt
VR = Vin − VL .
The first equation is solved by using an integrating factor and yields the current which must be differentiated to
give VL; the second equation is straightforward. The solutions are exactly the same as those obtained via Laplace
transforms.
Vin
IR =
R
Vin jVin
IL = =− .
jωL ωL
This shows that the inductor lags the resistor (and source) current by 90°.
The parallel circuit is seen on the output of many amplifier circuits, and is used to isolate the amplifier from capacitive
loading effects at high frequencies. Because of the phase shift introduced by capacitance, some amplifiers become
unstable at very high frequencies, and tend to oscillate. This affects sound quality and component life (especially the
transistors), and is to be avoided.
32 CHAPTER 4. RL CIRCUIT
IR IL
Vin Vout
R L
Parallel RL circuit
• RC circuit
• RLC circuit
• Electrical network
4.6 References
4.7. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 33
4.7.2 Images
• File:Discharging_capacitor.svg Source: https://upload.wikimedia.org/wikipedia/commons/a/a4/Discharging_capacitor.svg License: CC
BY-SA 3.0 Contributors: Own work Original artist: Ismaelteodoro
• File:LC_parallel_simple.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b2/LC_parallel_simple.svg License: CC BY-
SA 3.0 Contributors: Own work Original artist: First Harmonic
• File:Low_cost_DCF77_receiver.jpg Source: https://upload.wikimedia.org/wikipedia/commons/9/9f/Low_cost_DCF77_receiver.jpg
License: Public domain Contributors: Own work Original artist: User:Jahoe
• File:Parallel_LC_Circuit.svg Source: https://upload.wikimedia.org/wikipedia/commons/5/53/Parallel_LC_Circuit.svg License: CC
BY-SA 3.0 Contributors: Own work Original artist: First Harmonic
• File:Question_book-new.svg Source: https://upload.wikimedia.org/wikipedia/en/9/99/Question_book-new.svg License: Cc-by-sa-3.0
Contributors:
Created from scratch in Adobe Illustrator. Based on Image:Question book.png created by User:Equazcion Original artist:
Tkgd2007
• File:RC_Parallel_Filter_(with_I_Labels).svg Source: https://upload.wikimedia.org/wikipedia/commons/2/27/RC_Parallel_Filter_%28with_
I_Labels%29.svg License: Public domain Contributors: Own work Original artist: Inductiveload (talk)
• File:RC_Series_Filter_(with_V&I_Labels).svg Source: https://upload.wikimedia.org/wikipedia/commons/e/e0/RC_Series_Filter_%28with_
V%26I_Labels%29.svg License: Public domain Contributors: Own work Original artist: Inductiveload
• File:RL_Parallel_Filter_(with_I_Labels).svg Source: https://upload.wikimedia.org/wikipedia/commons/b/bf/RL_Parallel_Filter_%28with_
I_Labels%29.svg License: Public domain Contributors: Own work Original artist: Inductiveload
• File:Series-RL.png Source: https://upload.wikimedia.org/wikipedia/commons/8/8b/Series-RL.png License: CC-BY-SA-3.0 Contribu-
tors: ? Original artist: ?
• File:Series_LC_Circuit.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b9/Series_LC_Circuit.svg License: CC BY-
SA 3.0 Contributors: Own work Original artist: First Harmonic
34 CHAPTER 4. RL CIRCUIT