RC Circuit: Natural Response Complex Impedance Series Circuit

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

RC circuit

A resistor–capacitor circuit (RC circuit), or RC filter or RC network, is an electric circuit composed of


resistors and capacitors. It may be driven by a voltage or current source and these will produce different
responses. 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.

Contents
Introduction
Natural response
Complex impedance
Sinusoidal steady state
Series circuit
Transfer functions
Poles and zeros
Gain and phase
Current
Impulse response
Frequency-domain considerations
Time-domain considerations
Integrator
Differentiator
PWM Averaging Responses
Parallel circuit
Synthesis
See also
References
Bibliography

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.
Natural response
The simplest RC circuit consists of a resistor and a charged capacitor
connected to one another in a single loop, without an external voltage
source. Once the circuit is closed, the capacitor begins to discharge its
stored energy through the resistor. The voltage across the capacitor,
which is time-dependent, can be found by using Kirchhoff's current law.
The current through the resistor must be equal in magnitude (but opposite
in sign) to the time derivative of the accumulated charge on the capacitor.
This results in the linear differential equation

where C is the capacitance of the capacitor.

Solving this equation for V yields the formula for exponential decay:
RC circuit

where V0 is the capacitor voltage at time t = 0 .

V
The time required for the voltage to fall to 0 is called the RC time constant and is given by,[1]
e

In this formula, τ is measured in seconds, R in ohms and C in farads.

Complex impedance
The complex impedance, Z C (in ohms) of a capacitor with capacitance C (in farads) is

The complex frequency s is, in general, a complex number,

where

j represents the imaginary unit: j2 = −1,


σ is the exponential decay constant (in nepers per second), and
ω is the sinusoidal angular frequency (in radians per second).

Sinusoidal steady state

Sinusoidal steady state is a special case in which the input voltage consists of a pure sinusoid (with no
exponential decay). As a result, and the impedance becomes
Series circuit
By viewing the circuit as a voltage divider, the voltage across the
capacitor is:

and the voltage across the resistor is: Series RC circuit

Transfer functions

The transfer function from the input voltage to the voltage across the capacitor is

Similarly, the transfer function from the input to the voltage across the resistor is

Poles and zeros

Both transfer functions have a single pole located at

In addition, the transfer function for the voltage across the resistor has a zero located at the origin.

Gain and phase

The magnitude of the gains across the two components are

and
Amplitude and phase transfer functions for a series RC
circuit

and the phase angles are

and

These expressions together may be substituted into the usual expression for the phasor representing the output:

Current

The current in the circuit is the same everywhere since the circuit is in series:
Impulse response

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 capacitor voltage is

The impulse response of a series RC circuit


where u(t) is the Heaviside step function and
τ = RC is the time constant.
Similarly, the impulse response for the resistor voltage is

where δ(t) is the Dirac delta function

Frequency-domain considerations

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 → ∞ :

As ω → 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

Solving the above equation yields


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 :

As ω → ∞ :

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.

Time-domain considerations

This section relies on knowledge of e, the natural logarithmic constant.

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.
Vin = 0 before t = 0 and then Vin = V afterwards):

Partial fractions expansions and the inverse Laplace transform yield:

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
Capacitor voltage step-response.
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
1
resistor) to within of its final value. That is, τ is the time it takes
e
VC to reach V(1 − 1e ) and VR to reach V( 1e ).

1
The rate of change is a fractional 1 − e per τ. Thus, in going from
t = Nτ to t = (N + 1)τ, the voltage will have moved about 63.2% Resistor voltage step-response.
of the way from its level at t = Nτ toward its final value. So the
capacitor will be charged to 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:

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

Consider the output across the capacitor at high frequency, i.e.

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 given
above:

but note that the frequency condition described means that

so

which is just Ohm's Law.


Now,

so

which is an integrator across the capacitor.

Differentiator

Consider the output across the resistor at low frequency i.e.,

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

so

Now,

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).

PWM Averaging Responses

We begin with the analysis using the capacitor definition:


i=C*dv/dt

In this circuit, the current i is noted to be (E-v)/R if v is the average


capacitor voltage and E is a constant DC voltage.. This gives us:

C*dv/dt=(E-v)/R

Since E takes on two values here, both E (a particular DC voltage)


and 0 (zero voltage) we need two equations, one when it is E and one
when it is zero. The second equation is the same with E set to zero
and the polarity of v made positive since when the cap is discharging
v is some positive value. This gives us:
PWM RC Series Circuit
C*dv/dt=v/R

Now simply multiply both sides of this set of equations by C (and RC=R*C) we get:

dv/dt=(E-v)/RC (when the PWM input E is high)

and

dv/dt=v/RC (when the PWM input E is zero)

The time increment dt is only the same for a 50 percent PWM duty cycle, so we need a more general
expression. For this we simply set each time increment dt to a unique value:

dv/dt1=(E-v)/RC

dv/dt2=v/RC

Now simply solve for dv in each equation:

dv=dt1*(E-v)/RC

dv=dt2*v/RC

Now applying the theory of Continuity of States, we can say that these two values for dv must be the same
when the voltage across the capacitor is at its average value. That is because when the voltage goes up starting
at a certain lower value it must later come down to the same lower voltage or else the voltage is not yet at its
average value. So noting that, we can equate the two:

dt1*(E-v)/RC=dt2*v/RC

and now solving for the average voltage v we get:

v=(dt1*E)/(dt2+dt1)

Now we could stop here and note that v is the average voltage across the cap and dt1 is the 'on' time and dt2 is
the 'off' time, but in most cases we want to relate this to the duty cycle D. It is quite simple to note that if we
know the total time period tp we can equate these two:

dt1=D*tp

dt2=(1-D)*tp
where D is the fractional duty cycle (0.30 is 30 percent for example) and so substituting those two into the
previous solution for v we end up with:

v=(tp*D*E)/(tp*D+tp*(1-D))

and when we simplify this expression we get:

v=D*E

A very simple result! So the average voltage is D*E and we might immediately note that the actual values of R
and C did not matter. That is actually the case for any values as long as the capacitor voltage does not get close
to zero or E for either the 'on' or 'off' periods. Later when we calculate the peaks, we will find that the R and C
values do not matter if RC>>tp but if the RC time constant is comparable to the total time period tp then we
will find a difference in the upper and lower peaks of the time domain calculations and the upper and lower
peaks of the averaged calculations, although the differences may be small.

A simple example is when E=10v and D=0.25, the average voltage v is 2.5 volts.

Another simple example is when E=20v and D=0.50, the average voltage is 10 volts.

Next we will calculate the two peak values, the upper peak and the lower peak. There are at least two ways to
do this one using an averaging technique and another using a straight up time domain solution. The averaging
technique assumes a short total time period tp while the time domain solution assumes nothing except ideal
components (as is typical in theoretical solutions).

Using the averaging technique, we note that previously we got the result for the low to high going capacitor
voltage deviation as:

dv=dt1*(E-v)/RC

and we substituted dt1=D*tp and got:

dv=D*tp*(E-v)/RC

and we assumed that 'v' was the average voltage. Since 'v' was the average voltage and we later calculated that
as:

v(avg)=D*E

we insert that into the above expression and end up with:

dv=D*tp*(E-D*E)/RC

or:

dv=D*tp*E*(1-D)/RC

and that is the entire deviation from lowest point to highest point across the capacitor, often referred to as the
voltage peak to peak.

Since that is the entire deviation and in the straight line approximation it is triangular and a triangular wave has
average that is 1/2 of its entire amplitude peak to peak, to get the excursion above the average we just divide
that in half. For the same reason the amplitude below the average will also be one half of that.
We might note that the RC values do in fact make a difference for this calculation even though the did not
matter for the average calculation.

Example: R=1000 Ohms, C=100uf, tp=0.001 seconds, E=10 volts, D=0.50 (50 percent duty cycle)

Result: dv=0.025 volts peak to peak.

Positive excursion: 0.025/2=0.0125 volts peak

Negative excursion: 0.025/2=0.0125 volts peak

Also noteworthy is that if we calculate the maximum of dv with respect to a variation in D the duty cycle, we
will find that the value of D that causes the highest voltage peak to peak is D=0.50 which is a 50 percent duty
cycle.

Parallel circuit
The parallel RC circuit is generally of less interest than the series
circuit. This is largely because the output voltage Vout is equal to
the input voltage Vin — as a result, this circuit does not act as a
filter on the input signal unless fed by a current source.

With complex impedances:

Parallel RC circuit

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:

When fed by a current source, the transfer function of a parallel RC circuit is:

Synthesis
It is sometimes required to synthesise an RC circuit from a given rational function in s. For synthesis to be
possible in passive elements, the function must be a positive-real function. To synthesise as an RC circuit, all
the critical frequencies (poles and zeroes) must be on the negative real axis and alternate between poles and
zeroes with an equal number of each. Further, the critical frequency nearest the origin must be a pole,
assuming the rational function represents an impedance rather than an admittance.

The synthesis can be achieved with a modification of the Foster synthesis or Cauer synthesis used to
synthesise LC circuits. In the case of Cauer synthesis, a ladder network of resistors and capacitors will
result.[2]
See also
RL circuit
LC circuit
RLC circuit
Electrical network
List of electronics topics
Step response

References
1. Horowitz & Hill, p. 1.13
2. Bakshi & Bakshi, pp. 3-30–3-37

Bibliography
Bakshi, U.A.; Bakshi, A.V., Circuit Analysis - II, Technical Publications, 2009
ISBN 9788184315974.
Horowitz, Paul; Hill, Winfield, The Art of Electronics (3rd edition), Cambridge University Press,
2015 ISBN 0521809266.

Retrieved from "https://en.wikipedia.org/w/index.php?title=RC_circuit&oldid=1017523708"

This page was last edited on 13 April 2021, at 07:58 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like