You are on page 1of 59

Time Response Of Electrical Circuits

By this time you have probably studied DC circuits and looked at AC


circuits. You may have calculated how a circuit responds to a sinusoidal
signal. However, there are many situations in which the time behavior of a
circuit is important, and the signal is not a sinusoidal signal, or the circuit is not
a DC circuit. Here is an example.
In a 555 timer circuit, a DC voltage is applied to a series combination of
two resistors and a capacitor. The capacitor charges and eventually triggers
an electronic switch to change state causing the capacitor to discharge
through one of the resistors. In that way, the charging and discharging of the
capacitor controls the timing of an ever continuing sequence of square pulses
at the output of the 555 chip.
In this lesson, we will look at the time response of
electrical circuits. We will work from simple responses to
more complicated response. We will find that circuits can
have several causes for the responses we observe. We
start with the simple circuit shown at the right. We will
assume that we have a capacitor that has a charge on it so
that the initial voltage across the capacitor is 10 volts. At
some arbitrary time - we will call it t = 0 - the switch is closed and current
begins to flow from the capacitor through the resistor, resulting in a discharge
of the capacitor. We want to determine - mathematically - how the voltage
across the capacitor behaves in time.
Start by writing KCL at the node between the capacitor and the resistor.
C (dVC/dt) + (VC /R) = 0
This equation can be rearranged to work toward a solution.
dVC/dt = - VC /(RC)
This expression for the rate of change of the capacitor voltage is interesting by
itself. It says that as time goes on the capacitor voltage decreases - it tends
to zero. As the capacitor voltage gets smaller the rate of change also gets
smaller.
Ad Info

System Dynamics - Time Constants


Where Are Time Constants Found?
Measuring Time Constants

Time Constants - Where are they found?


Time Constants are ubiquitous. They are found in many different
kinds of systems, including the following.

Electrical Systems.
o Resistor-Capacitor circuits have time constants.
o Resistor-Inductor circuits have time constants.
Mechanical Systems
o Many motions exhibit time-constant behavior. For example
when a motor speeds up, the measurements of speed would
reveal time-constant behavior.
Thermal Systems
o Heating up and cooling down in simple thermal systems shows
time-constant behavior. Lakes cooling off after the summer
slow down the approach of winter near the lake, and later
lakes slowly heating up in the spring retard the approach of
summer. That behavior can be explained using time constants.
Physiological Systems
o The inner ear has dynamics that can be explained using two
time constants
Psychological Systems
o Psychologists have measured the time constant that
determines how much learned material you retain as a function
of time after learning.

How Do Time Constants Come About?

Time constants are parameters of systems that obey first order,


linear differential equations. That would be a differential equation like
this one.
(dx(t)/dt) + x(t) = Gu(t)
In this situation, the variables are:

x(t) = response
u(t) = input or driving function
Examples:
o u(t) = heat in, x(t) = temperature
o u(t) = voltage into a circuit, x(t) = output voltage
o u(t) = voltage driving a motor, x(t) = motor speed.
The constants in the differential equation are:

= the Time Constant


o The time constant is a measure of how quickly the system
responds.
G = the System Gain
o At steady state, the response, x(t), can be calculated by
multiplying the input by the gain, G.
The responses you need to know about for this sytem include:

Response when the input is zero (There is no input), but the output
has some initial value.
-t/
o x(t) = x(0)e
, which looks like:

Examples:
Motor slowing down when the power is turned off.
Temperature rise above ambient decaying to zero when
the heat source is turned off.
Response when the input is a constant (Commonly called the Step
Response), which looks like:
o

o
o

x(t) = G*u*(1 - e-t/ )


Examples:
Motor speeding up when the power is turned on.
Temperature rising when the heat is turned on.

Measuring Time Constants

In many experimental situations you need to measure the time


constant of a system. In this section we will examine properties of time
constant response that permit you to get a measurement of a time
constant.
We start by examining some typical time-constant behavior. Here is
a response of a first order system that exhibits time response behavior.
We need to think about the features of
this response that will permit us to get
measurements that look like this and
determine the time constant of the
system from these kinds of
measurements.
If this is the impulse response of a
system, it would have this for.
x(t) = (Gdc/)e-t/
The general form is given by:
x(t) = x(0)e-t/
We can look for important points in this response. The most obvious is
when t = . That's when the elapsed time is one time constant. When one
time constant has elapsed the response is:
x(t) = x(0)e-1 = x(0)(0.36788)
The 37% Method
This gives us the first method for finding a time constant. Here is the
method.

Determine the initial value of the variable, x(0) in the above.

Determine the time when the variable has decayed to 37% of the
initial value.
That time is the measured time constant.

That is a simple method that can be applied to numerous situations. You


need to note the following.

You need a response that dies out to zero. If you have a response
that has the right shape, but which does not die out to zero, then
you may have to redefine a component of the response in order to
meet this requirement.
The method depends upon the value of the response at only one
point. In essence, you throw away all of the data that you have
except for the starting value and the 37% value.

The first note above has some implications. Here is a response that
exhibits time constant behavior.

This response does not decay to zero. However, it does apparently decay
to a value of 17, so if you subtracted 17 from the response you would get
a response function that did decay to zero. That might make some sense
physically in many situations. For example, the plot above might be the
temperature above ambient (in Celsius degrees) and the temperature rise

above ambient would fall to zero as the temperature fell to ambient in a


thermal system. Look for situations like that.
The same kind of reasoning would also apply to finding time
constants from step responses. Here is a step response.

Here the steady state is 50, and you could take the difference between
the steady state (50) and the response. That difference decays to zero
as the response approaches the steady state, and that gives you the kind
of function that you want.
Here is the method.

Determine the initial slope of the variable, x(0) in the above.


Determine the time when the variable would have decayed to zero if
the initial rate of decay (slope) had continued.
That time is the measured time constant.

You need to note the following.

You do not need a response that dies out to zero. If you have a
response that decays to some non-zero final value, just extend the
initial slope line until the extended line hits the final value.

The method depends upon the slope of the response at only one
point. In essence, you throw away all of the data that you have
except for the starting value and the initial slope.

The Initial Slope Method


There are other methods that you can use. One other method usese
the initial slope of the response. Consider the equation for the response
curve.
x(t) = x(0)e-t/
The slope is given by:
dx(t)/dt = [-x(0)/t]e-t/
Interestingly, if the response continued to decay at this rate - and we
know that it doesn't - the response would decay to zero in seconds.
Since the rate of change is -x(0)/t, if it continued t for seconds the total
change would be -x(0), which would bring the response to zero.
The implications of the argument above are that you can draw a line
from the initial point at the initial slope and check where that line
intersects the final value, and the time when that happens is the time
constant. That is better when you have a visual representation. Here is a
plot.

In this plot, the initial slope has been extended (blue line) and intersects
the final value (0) at about 25 seconds, so the time constant would be 25
seconds.
The Logarithmic Method
There are other methods that you can use. Both of the previous
methods really depend upon values at a single time. There is another
method that uses all of the data. Consider the equation for the response
curve.
x(t) = x(0)e-t/
Take the logarithm of the response curve and we have the following.
ln[x(t)] = ln[x(0)] - t/
This is the equation of a straight line when plotted against t. That leads
us to the following strategy.

Take a set of data at different times. That data set would be x(tk)
and would comprise all of the values and all of the times.
Plot the values of the data, x(tk), against the times, tk. That should
give you a straight line.
Determine the equation of the line.
o The slope of the line is the time constant.
o The intercept is the natural log of the initial value.

Note the following about this method.

This method uses all of the data.


If the data is noisy, you can use regression to get the straight line
parameters (slope and intercept).
You must (repeat MUST) have data that decays to zero. If you do
not have data that decays to zero, see the note above about

creating a set of data that decays to zero by extracting the


essential part of the data from the data set.
What happens if you don't have a decaying exponential? In that
case, you need to isolate the decaying exponential that is buried in the
data. Here is an example.

In this response, the data does not decay to zero. However, if you take
the following steps, you can get a data set that you can use with this
method.

First, determine the steady state. In real data sets you will have to
make the best estimate you can.
Secondly, subtract the data from the steady state value you
determined in the first step.
o If the data can be put into a spreadsheet that's an easy step.
o You subtract the data from the steady state so you can get
numbers that are all positive so that you can take a logarithm
of your result in this step.
Finally, you have some exponentially decaying data. Take the
logarithm and you should see a straight line. You might have to
adjust the steady state value a little in either direction to see what
gives you the best straight line.

Capacitors
Why Are Capacitors Important?
What Is A Capacitor?
Voltage Current Relationships In Capacitors
Energy In Capacitors
What Is Impedance?
Using Impedance
Some Impedance Laws/Combinations
Why Are Capacitors Important?
The capacitor is a widely used electrical component. It has several
features that make it useful and important:
A capacitor can store energy, so capacitors
are often found in power supplies.
A capacitor has a voltage that is proportional
to the charge (the integral of the current) that
is stored in the capacitor, so a capacitor can
be used to perform interesting computations
in op-amp circuits, for example.
Circuits with capacitors exhibit frequencydependent behavior so that circuits that
amplify certain frequencies selectively can
be built.
What Is A Capacitor?
Capacitors are two-terminal electrical
elements. Capacitors are essentially two
conductors, usually conduction plates - but
any two conductors - separated by an
insulator - a dielectric - with conection wires
connected to the two conducting plates.
Capacitors occur naturally. On printed circuit boards two wires running
parallel to each other on opposite sides of the board form a capacitor. That's a
capacitor that comes about inadvertently, and we would normally prefer that it

not be there. But, it's there. It has electrical effects, and it will affect your
circuit. You need to understand what it does.
At other times, you specifically want to use capacitors because of their
frequency dependent behavior. There are lots of situations where we want to
design for some specific frequency dependent behavior. Maybe you want to
filter out some high frequency noise from a lower frequency signal. Maybe you
want to filter out power supply frequencies in a signal running near a 60 Hz
line. You're almost certainly going to use a circuit with a capacitor.
Sometimes you can use a capacitor to store energy. In a subway car, an
insulator at a track switch may cut off power from the car for a few feet along
the line. You might use a large capacitor to store energy to drive the subway
car through the insulator in the power feed.
Capacitors are used for all these purposes, and more. In this chapter
you're going to start learning about this important electrical component.
Remember capacitors do the following and more.

Store energy
Change their behavior with frequency
Come about naturally in circuits and can change a circuit's
behavior

GOALS
You need to know what you should get from this lesson on capacitors.
Here's the story.
Given a capacitor,
Be able to write and use the voltage-current
relationship for the capacitor,
Be able to compute the current through a
capacitor when you know the voltage across
a capacitor.
Given a capacitor that is charged,
Be able to compute the amount of energy that
is stored in the capacitor.

Capacitors and inductors are both elements that can store energy in
purely electrical forms. These two elements were both invented early in
electrical history. The capacitor appeared first as the legendary Leyden jar, a
device that consisted of a glass jar with metal foil on the inside and outside of
the jar, kind of like the picture below. This schematic/picture shows a battery
attached to leads on the Leyden jar capacitor.
Although this device first appeared in
Leyden, a city in the Netherlands sometime
before 1750. It was discovered by E. G. von
Kleist and Pieter van Musschenbroek.
Although it has been around for about 250
years, it has all of the elements of a modern
capacitor, including:

Two conducting plates. That's the metallic foil in the Leyden jar.
An insulator that separates the plates so that they make no
electrical contact. That's the glass jar - the Leyden jar.

The way the Leyden jar operated was that


charge could be put onto both foil elements. If
positive charge was put onto the inside foil, and
negative charge on the outside foil, then the two
charges would tend to hold each other in place.
Modern capacitors are no different and usually
consist of two metallic or conducting plates that
are arranged in a way that permits charge to be
bound to the two plates of the capacitor. A simple physical situation is the one
shown at the right.
If the top plate contains positive charge, and the bottom plate contains
negative charge, then there is a tendency for the charge to be bound on the
capacitor plates since the positive charge attracts the negative charge (and
thereby keeps the negative charge from flowing out of the capacitor) and in
turn, the negative charge tends to hold the positive charge in place. Once
charge gets on the plates of a capacitor, it will tend to stay there, never
moving unless there is a conductive path that it can take to flow from one
plate to the other.
There is also a standard circuit symbol for a capacitor. The figure below
shows a sketch of a physical capacitor, the corresponding circuit symbol, and
the relationship between Q and V. Notice how the symbol for a capacitor

captures the essence of the two plates and the insulating dielectric between
the plates.

Now, consider a capacitor that starts out with no charge on either plate. If
the capacitor is connected to a circuit, then the same charge will flow into one
plate as flows out from the other. The net result will be that the same amount
of charge, but of opposite sign, will be on each plate of the capacitor. That is
the usual situation, and we usually assume that if an amount of charge, Q, is
on the positive plate then -Q is the amount of charge on the negative plate.
The essence of a capacitor is that it stores charge. Because they
store charge they have the properties mentioned earlier - they store energy
and they have frequency dependent behavior. When we examine charge
storage in a capacitor we can understand other aspects of the behavior of
capacitors.
In a capacitor charge can accumulate on the two plates. Normally charge
of opposite polarity accumulates on the two plates, positive on one plate and
negative on the other. It is possible for that charge to stay there. The positive
charge on one plate attracts and holds the negative charge on the other plate.
In that situation the charge can stay there for a long time.
That's it for this section. You now know pretty much what a capacitor is.
What you need to learn yet is how the capacitor is used in a circuit - what it
does when you use it. That's the topic of the next section. If you can learn that
then you can begin to learn some of the things that you can do with a
capacitor. Capacitors are a very interesting kind of component. Capacitors are
one large reason why electrical engineers have to learn calculus, especially
about derivatives. In the next section you'll learn how capacitors influence
voltage and current.
Voltage-Current Relationships In Capacitors

There is a relationship between the charge on a capacitor and the


voltage across the capacitor. The relationship is simple. For most
dielectric/insulating materials, charge and voltage are linearly
related.
Q=CV
where:

V is the voltage across the plates.

You will need to define a polarity for that voltage. We've defined the voltage
above. You could reverse the "+" and "-".

Q is the charge on the plate with the "+" on the voltage polarity
definition.
C is a constant - the capacitance of the capacitor.

The relationship between the charge on a capacitor and the voltage


across the capacitor is linear with a constant, C, called thecapacitance.
Q=CV
When V is measured in volts, and Q is measured in couloumbs, then C
has the units of farads. Farads are reallycoulombs/volt.
The relationship, Q = C V, is the most important thing you can know
about capacitance. There are other details you may need to know at times,
like how the capacitance is constructed, but the way a capacitor behaves
electrically is determined from this one basic relationship.
Shown to the right is a circuit
that has a voltage source, Vs, a
resistor, R, and a capacitor, C. If
you want to know how this circuit
works, you'll need to apply KCL
and KVL to the circuit, and you'll
need to know how voltage and
current are related in the capacitor.
We have a relationship between
voltage and charge, and we need
to work with it to get a voltage
current relationship. We'll look at that in some detail in the next section.

The basic relationship in a capacitor is that the voltage is proportional to


the charge on the "+" plate. However, we need to know how current and
voltage are related. To derive that relationship you need to realize that
the current flowing into the capacitor is the rate of charge flow into the
capacitor. Here's the situation. We'll start with a capacitor with a time-varying
voltage, v(t), defined across the capacitor, and a time-varying current, i(t),
flowing into the capacitor. The current, i(t), flows into the "+" terminal taking
the "+" terminal using the voltage polarity definition. Using this definition we
have:
ic(t) = C dvc(t)/dt
This relationship is the fundamental relationship between current and
voltage in a capacitor. It is not a simple proportional relationship like we found
for a resistor. The derivative of voltage that appears in the expression for
current means that we have to deal with calculus and differential equations
here - whether we want to or not.
Q1. If the voltage across a capacitor is descreasing (and voltage and current
are defined as above) is the current positive of negative?

This derivative kind of relationship also has some implications for what
happens in a capacitor, and we are going to spend some time exploring that
relationship. Clearly, we need to understand what this relationship implies,
and then we need to learn how it affects things when we write circuit
equations using KVL and KCL.
We'll start by considering a time varying
voltage across a capacitor. To have something
specific, let's say that we have a 4.7mf capacitor,
and that the voltage across the capacitor is the
voltage time function shown below. That voltage
rises from zero to ten volts in one millisecond,
then stays constant at ten volts. Before you go
on try to determine what the current through the
capacitor looks like, then answer these
questions.

Q2. Is the current constant in the time interval from t = 0 to t = 10 msec?


Q3. Is the current constant in the time interval from t = 10 msec to the last
time shown?

If current is proportional to the time derivative of voltage, there is


only one time segment, from t = 0 to 1 millisecond, where the
voltage derivative is non-zero, so that's the only time there is any
currrent that is non-zero.
After one millisecond has elapsed, the voltage derivative goes to
zero, so there isn't any current then. If there isn't any current, then
the voltage stays constant because no charge is flowing in or out.
Remember, current is charge flow!
The voltage derivative is constant from t = 0 to 1 millisecond. If
that's true, then the current is constant in that period.
Now, you should be able to compute the current.

Energy In Capacitors
Storing energy is very important. You count on the energy stored in your
gas tank if you drove a car to school or work today. That's an obvious case of
energy storage. There are lots of other places where energy is stored. Many
of them are not as obvious as the gas tank in a car. Here are a few.

You're reading this on a computer, and the computer keeps track


of the date and time. It does that by keeping a small part of the
computer running when you think that the computer is turned off.
There's a small battery that stores the energy to keep the clock
running when everything else is turned off.
If you own a stereo or television that you have to plug into the wall
plug, then you should realize that the wall plug voltage becomes
zero 120 times a second. When that happens, the system keeps
running because there are capacitors inside the system that store
energy to carry you through those periods when the line voltage
isn't large enough to keep things going!

Capacitors can't really be used to store a lot of energy, but there are
many situations in which a capacitor's ability to store energy becomes
important. In this lesson we will discuss how much energy a capacitor can
store.
Capacitors are often used to store energy.

When relatively small amounts of energy are needed.


Where batteries are not desired because they might deteriorate.
For larger power/short duration applications - as in power supply
filters, or to keep power up long enough for a computer to shut
down gracefully when the line power fails.

To calculate how much energy is stored in a capacitor, we start by


looking at the basic relationship between voltage and current in a capacitor.
i(t) = C dv(t)/dt
Once we have this relationship, we can calculate the power the rate of flow of energy into the capacitor - by multiplying the
current flowing through the capacitor by the voltage across the
capacitor.
P(t) = i(t)v(t)

Given the expression for the power:


P(t) = i(t)v(t)

And given the expression for the current:


i(t) = C dv(t)/dt

We can use the expression for current in the power expression:


P(t) = (C dv(t)/dt) v(t)

We can recognize that power is simply rate of energy input.


P(t) = dE/dt = (C dv(t)/dt) v(t)

Now, the derivative of energy can be integrated to find the total


energy input.

P(t) = dE/dt = (C dv(t)/dt) v(t)

gives

Now, assuming that the initial voltage is zero (there is no energy


stored in the capacitor initially, we find that the energy stored in a
capacitor is proportional to the capacitance and to the square of
the voltage across the capacitor.
Ec = (1/2)CV2

The expression for the energy stored in a capacitor resembles


other energy storage formulae.

For kinetic energy, with a mass, M, and a velocity, v.


EM = (1/2)MV2

For potential energy, with a spring constant, K, and an elongation,


x.
ESpring = (1/2)Kx2

Since the square of the voltage appears in the energy formula, the
energy stored is always positive. You can't have a negative amount
of energy in the capacitor. That means you can put energy into the
capacitor, and you can take it out, but you can't take out more than
you put in.

Power in to the capacitor can be negative. Voltage can be positive


while current is negative. Imagine a capacitor that is charged. You
could charge a capacitor by putting a battery across the capacitor,
for example. Then, if you placed a resistor across the capacitor,
charge would leave the capacitor - current would flow out of the
capacitor - and the energy in the capacitor would leave the
capacitor only to become heat energy in the resistor.
When energy leaves the capacitor, power is negative.

When you use capacitors in a circuit and you analyze the


circuit you need to be careful about sign conventions.
Here are the conventions we used, and these
conventions were assumed in any results we got in this
lesson.

Frequency Dependent Behavior For A Capacitor


We start with a capacitor with a sinusoidal voltage across it.
where:

vC(t) = Voltage across the capacitor


iC(t) = Current through the capacitor
C = Capacitance (in farads)

We will assume that the voltage across the capacitor is sinusoidal:


vC(t) = Vmax sin(t)
Knowing the voltage across the capacitor allows us to calculate the
current:
iC(t) = C dvC(t)/dt = C Vmax cos(t) = Imax cos(t)
where Imax = C Vmax
Comparing the expressions for the voltage and current we note the following.

The voltage and the current are both sinusoidal signals (a sine
function or a cosine function) at the same frequency.
The current leads the voltage. In other words, the peak of the
current occurs earlier in time than the peak of the voltage signal.

The current leads the voltage by exactly 90o. It will always be


exactly 90o in a capacitor.
The magnitude of the current and the magnitude of the voltage are
related:
Vmax/Imax = 1/ C

Now, with these observations in hand, it is possible to see that there may
be an algebraic way to express all of these facts and relationships. The
method reduces to the following.

If we have a circuit with sinusoidally varying voltages and currents


(as we would have in a circuit with resistors, capacitors and
inductors and sinusoidal voltage and current sources) we
associate a complex variable with every voltage and current in the
circuit.
The complex variable for a voltage or current encodes the
amplitude and phase for that voltage or current.
The voltage and current variables can be used (using complex
algebra) to predict circuit behavior just as though the circuit were
a resistive circuit.

We need to do two things here. First, we can illustrate what we mean


with an example. Secondly, we need to justify the claim above. We will look
at an example first, and we will do two examples. The first example is jsut the
capacitor - all by itself. The second example will be one that you have
considered earlier, a simple RC low-pass filter.
Example 1 - The Capacitor
In a capacitor with sinusoidal voltage and currents, we have:
where:

vC(t) = Voltage across the inductor

vC(t) = Vmax sin(t)

iC(t) = Current through the inductor

iC(t) = C Vmax cos(t) = Imax cos(t)

C = Capacitance (in farads)

We represent the voltage with a complex variable, V. Considering this as


a complex variable, it has a magnitude of Vmaxand and angle of 0o. We would
write:
V = Vmax/0o
Similarly, we can get a representation for the current. However, first
note:
iC(t) = C Vmax cos(t) = Imax cos(t) = Imax sin(t + 90o)
(Here you must excuse the mixing of radians and degrees in the argument of
the sine. The only excuse is that everyone does it!) Anyhow, we have:
I = Imax/90o = j Imax = jC Vmax
Where j is the square root of -1.
Then we would write:
V/I = Vmax/jC Vmax = 1/jC
and the quantity 1/jC is called the impedance of the capacitor. In the next
section we will apply that concept to a small circuit - one you should have
seen before.
Before moving to the next section, a little reflection is in order. Here are
some points to think about.

A phasor summarizes information about a sinusoidal signal.


Magnitude and phase information are encoded into the phasor.
Frequency information is not encoded, and there is a tacit
assumption that all signals are of the same frequency, which
would be the case in a linear circuit with sinusoidal voltage and
current sources.
We looked at a case where we encoded a signal Vmax sin(t) into a
phasor of Vmaxo. That was completely arbitrary, and many others
would have encoded Vmax cos(t) into a phasor of Vmaxo.
Phasors are intended only to show relative phase information, and
it doesn't matter which way you go.

Using Impedance
In the last section we began to talk about the concept of impedance. Let
us do that a little more formally. We begin by defining terms.

A sinusoidally varying signal (vC(t) = Vmax sin(t) for example) will be


represented by a phasor, V, that incorporates the magnitude and phase angle
of the signal as a magnitude and angle in a complex number. Examples
include these taken from the last section. (Note that these phasors have
nothing to do with any TV program about outer space.)
vC(t) = Vmax sin(t)
is represented by a phasor V = Vmaxo
iC(t) = Imax sin(t + 90o)
is represented by a phasor I = Imaxo
va(t) = VA sin(t + )
is represented by a phasor Va = VA
Next, we can use the relationships for voltage and current phasors to
analyze a circuit. Here is the circuit.
Now, this circuit is really a frequency
dependent voltage divider, and it is
analyzed differently in another lesson.
However, here we will use phasors. At
the end of this analysis, you should
compare how difficult it is using phasors
to the method in the other lesson.
We start by noting that the current in
the circuit - and there is only one current
- has a phasor representation:
I = Imaxo
We will use the current phase as a reference, and measure all other phases
from the current's phase. That's an arbitrary decision, but that's the way we
will start.
Next we note that we can compute the voltage across the capacitor.
VC = I/jC

This expression relates the current phasor to the phasor that represents the
voltage across the capacitor. The quantity 1/jC is the impedance of the
capacitor. In the last section we justified this relationship.
We can also compute the phasor for the voltage across the resistor.
VR = IR
This looks amazingly like Ohm's law, and it is, in fact, Ohm's law, but it is in
phasor form. For that matter, the relationship between voltage and current
phasors in a capacitor - just above - may be considered a generalized form of
Ohm's law!
Now, we can also apply Kirchhoff's Voltage Law (KVL) to compute the
phasor for the input voltage.
VIN = VR + VC = IR + I/jC = I(R + 1/jC)
You should note the similarities in what happens here and what happens
when you have two resistors in series.
If you have a resistor, R, and a capacitor, C, in series, the current
phasor can be computed by dividing the input voltage phasor by the
sum of R and 1/jC.
If you have two resistors in series (call them R1and R2), the current can
be computed by dividing the input voltage by the sum of R1and R2.
Example
Consider a series circuit of a resistor and capacitor. The series
impedance is:
Z = R + 1/jC
That's the same as we showed just above. The impedance can be used to
predict relationships between voltage and current. Assume that the voltage
across the series connection is given by:
vSeries(t) = Vmax cos(t)
That corresponds to having a voltage phasor of:

V = Vmaxo
We also know that the impedance establishes a relationship between the
voltage and current phasors in the series circuit. In particular, the voltage
phasor is the product of the current phasor and the impedance.
V=IZ
For our particular impedance, we have:
V = I*(R + 1/jC)
So, we can solve for the current phasor:
I = V / (R + 1/jC)
Now, we know the voltage phasor and we know the impedance so we can
compute the current phasor. Let us look at some particular values.
Assume:
R = 1.0 k
C = .1f = 10-7 f
f = 1 kHz, so = 2 103
Vmax = 20 v
Then:
ZR = 1.0 k
ZC = 1/(jC) = 1/(j2 103 10-7 ) = j 1.59 k
And, the total impedance is:
Z = ZR + ZC = (1.0 + j 1.59) k
This impedance value can also be expressed in polar notation:
Z = 1.878 o
Now, compute the current phasor:
I = V / (R + 1/jC)
Substituting values, we find:
I = V / Z = Vmaxo / 1.878 o =20o / 1.878 o
I = V / Z = (20 / 1.878) o = 10.65 oamps
And, we need to examine exactly what this means for the current as a function
of time. But that isn't very difficult. We can write out the expression for the
current from what we have above.

iC(t) = 10.65 cos(t - 62o) amps


Ad Info

Inductors - UNDER CONSTRUCTION - Barely Started


The inductor is a peculiar electrical element. Ideally, it only exhibits its'
behavior when things are changing. Keep the status quo and it is well
behaved. Change things and you may see some effect you may consider to
be truly wierd.
Consider this. The inductor has a relationship between voltage and
current:
INSERT V = L di/dt
where:
v(t) = Voltage across the inductor
i(t) = Current through the inductor
L = Inductance (in henries)
So, if the current through the inductor is not changing, there is no voltage
across the inductor. However, if the current changes then things start to
happen. When you try to force the current through an inductor to change,
then you can have problems.
Consider this simple circuit:
INSERT V-R-L CIRCUIT
At some point in time the switch is opened. Prior to opening the switch, an
equilibrium is achieved. The current is constant and is simply Vs / RL. That's
because there is no voltage across the inductor. Remember the current is
constant so di/dt = 0 and there is no voltage across the inductor. It's almost
as though the circuit doesn't know that the inductor is there.
Now, consider what happens when the switch is opened. You probably
want to believe that opening the switch causes the current to go to zero.
Here's a little exercise for you then.
At the moment the switch is opened, write KVL and be sure to include
the voltage across the switch. Assume that the current drops from Vs / RL at

the instant that the switch opens and becomes zero instantaneously. Now,
answer the following questions:
What is the voltage across the inductor?
What is the voltage across the source?
What is the voltage across the resistor?
What is the voltage across the switch?
In each case, as you answer the question, be sure to supply reasons for your
conclusions.

Frequency Dependent Circuits


Why is Frequency Response Important?
An Example Circuit
Using Frequency Response Concepts
A Frequency Response Lab Problem

Why Worry About Frequency Response?


Did you ever buy audio equipment and look carefully at how the
manufacturer specified how well the equipment would work? (And audio
equipment is one of the few consumer items where people actually try to
sell things on the basis of how well they work!) If you looked at the
specifications for audio equipment you would probably find the following.

A frequency response for the unit.


If the unit is a speaker set, you'll find separate frequency
responses for the different speakers like the mid-range, or the
woofer and the tweeter.
If the unit is a microphone, you'll find a frequency response that
tells you how the unit responds to different frequencies.

Frequency response is an important concept in many areas - within


electrical engineering and outside of electrical engineering. Having a good
grasp of frequency response is important in many areas, so our objectives
in this lesson include the following.

Given a linear system or circuit described mathematically,


o Be able to compute a frequency response for the system.
o Be able to predict an output signal from a given input
sinusoidal signal.

An Example Circuit
We are going to examine a simple circuit that has frequency
dependent behavior, a resistor-capacitor (RC) circuit. It is shown below.
To illustrate how this circuit responds to a sinusoidal signal input we can
do any of the following.

We can write the differential equation relating the input and output
voltages and solve for the output assuming a sinusoidal signal input.
We can assume a sinusoidal input and use LaPlace transform
methods to compute the output voltage.
Since the input is a sinusoid, we know that the output contains a
sinusoid and terms that decay to zero. We can work from there.

We will use the third approach - and we will assume a steady state
output and work backwards from the output to compute the input.
Since the first thing we want to do is just to look at how a circuit
can affect sinsusoidal signals, we're going to assume a sinusoidal output
and work backwards to calculate the input voltage that produces that

output. That's not a very general approach, but it will get us what we want
now, and prepare us for other things to come. We will be able to do that
without too much algebraic pain, and we can learn some things from the
result.
So, we will assume that the output voltage is given by:
vout(t) = B sin(t)
Be sure that you understand that B is the magnitudeof the output
signal
Now, what does that form for the
output voltage imply?

If the output voltage is given by


vout(t) = B sin(t)

Then, since the output voltage is


across a capacitor, we can
compute the current flowing
through R and C as:
i(t) = Cdvout/dt = CB cos(t)

And then we can compute the voltage across the resistor, R, as:
vR(t) = Ri(t) = RCB cos(t)

Now, we can apply KVL to get the input voltage.

The input voltage is given by,


vin(t) = vR(t) +vout(t)

Or:

vin(t) = B(RC cos(t) + sin(t))


At this point step back from this. It may not be obvious, but we can
take advantage of a trigonometric identity,
sin(x+y) = sin(x)cos(y) + cos(x)sin(y)
if only we can make the things that multiply the sines and cosines in the
second bullet above look like other sines and cosines.

We know:
sin(x+y) = sin(x)cos(y) + cos(x)sin(y)

And, we know:
vin(t) = B(RC cos(t) + sin(t))

And the second expression can be put into the form of the first

We need to refer to a little geometrical


construction - at the right. "Clearly" we
have the relationships indicated below for
cos() and sin()

So, now we can write:

Which reduces to:

There are two conclusions to draw from the resulting expression for the
input voltage.

The ratio of the amplitude of the output to the input voltage is


given by:

The output voltage lags the input voltage by a phase angle, .

You can use the expressions for the gain, B/A, and the phase
shift, , to predict behvarior of circuits like this. You should note the
following in these expressions.

In the expression for B/A, there is a factor (RC) which


determines the attenuation. There is a "critical frequency" where
that factor is 1. That frequency is:
o = 1/RC
o or f = 1/2RC
When f = 1/2RC the attenuation is 0.707 (the reciprocal of the
square root of 2). You can consider that as the mid-point in
freqency where the frequency is between the high frequency range
(where the circuit does not pass a sinusoidal signal well) and the low
frequency range (where a sinusoidal signal tends to pass through the
filter unchanged).
When f = 1/2RC the phase is -45o. That is halfway between the
low frequency phase (which tends toward 0o as the frequency tends

to zero - i.e. DC) and the high frequency phase (which tends toward
-90o as the frequency gets very high).
So, there are two reasons to think of that frequency (f = 1/2RC)
as a critical frequency.
Note that that frequency is sometimes referred to as the
bandwidth of the circuit. It's one way to measure the band of
frequencies that get through the circuit relatively unscathed.

With those thoughts you can think a little more deeply using the
simulator we have just below.
A Simulation of the Circuit
Note: - This simulator is real time. However, to let you see how the
circuit behaves, we have made the signals very slow - on the order of a
few Hertz, or even a fraction of a Hertz. The time constant (the R-C
product) should be correspondingly long - on the order of a second (from
a fraction of a second to a few seconds). You won't see much if you stray
far from these limits - even though these are long time constants and the
bandwidths are quite low. That's just for purposes of illustration.
(However, note that you could get a one second time constant using R =
1.0 M, and C = 1.0f.)
Here is the simulator.

Using this simulator, you can do the following.

You can change the frequency.


o Notice that higher frequencies are attenuated more. (The
output is smaller.)
o Lower frequencies are attenuated less. (The output is larger.)
You can change the time constant (The R-C product).
o Notice that higher time constants shift the bandwidth lower,
and high frequencies are attenuated more.
o Lower time constants shift the bandwidth higher, and high
frequencies are attenuated less. (The output is larger.)
Actually, with lower time constants, the bandwidth is higher
and more frequencies get through the circuit.

Problems & Questions


P1. Here is an RC filter circuit - the same one discussed above.

In this circuit, the parameters are:

R = 10 k
C = 0.1 f

Determine the bandwidth of the


circuit.
Enter your answer in the box below,
then click the button to submit your
answer. You will get a grade on a 0 (completely wrong) to 100 (perfectly
accurate answer) scale.

Your grade is:


P2. In the RC filter circuit determine the resistance value that gives a
bandwidth ten times as large as in problem 1.

Your grade is:


P3. In the RC filter circuit , the parameters are:

R = 10 k
C = 0.1 f

Determine the phase shift between output and input at 200Hz.


Remember the sign and give your answer in degrees (not radians).

Your grade is:

P4. Assume that you have an RC filter circuit with a 0.5 second time
constant. Determine the phase shift between output and input at 1.0Hz.
Remember the sign and give your answer in degrees (not radians).
NOTE: This problem has a long time constant and a low frequency so that
you can use a real time simulator to check your answer before you submit
it. Click here to get the simulator in a separate window.

Your grade is:

P5. In an RC filter circuit , the parameters are:

R = 1 k
C = 0.047 f

Determine the frequency (in Hertz) for which the attenutation is 0.707.
In other words, the output is reduced by 29.3%.

Your grade is:


Reflections
In this lesson you have been introduced to a simple frequency
dependent circuit. We have used a very brute-force method - assuming a
voltage at the output and chasing that back to the input. That worked
here, but it won't work everywhere. Moreover, this is not the easiest way
to make such predictions, and before you attempt more complex circuits
you should learn a better way to analyze frequency dependent circuits.
You need to learn about impedance and phasors. There are links below
that will take you to many other topics.

Problem and Labs

Problems
o Problem Freq1P01
o Problem Freq1P02
o Problem Freq1P03
o Problem Freq1P04
o Problem Freq1P05
Labs
o Resistor Capacitor Filter Frequency Response

Bode' Plots
Why Bode' Plots?
What are Bode' Plots?
First Order Systems
Decibels
Bode' Plots for
Second Order Systems
Bode Plots for Larger Systems (Examples)
You are at: Basic Concepts - System Models - Frequency Response - Bode' Plots
Click here to return to the Table of Contents

Why Bode' Plots?


Bode plots are the most widely used means of displaying and communicating
frequency response information. There are many reasons for that.
Bode' plots are really log-log plots, so they collapse a wide range
of frequencies (on the horizontal axis) and a wide range of gains (on thevertical
axis) into a viewable whole.
In Bode' plots, commonly encountered frequency responses have a shape that
is simple. That simple shape means that laboratory measurementscan easily be

discerned to have the common factors that lead to those shapes. For example,
first order systems have two straight line asymptotes and if you take data and plot
a Bode' plot from the data, you can pick out first order factors in a transfer
function from the straight line asymptotes.
You may have used Bode' plots without knowing it. Stereo equipment amplifiers, speakers, microphones, headsets, etc. - often have frequency response
specifications, and when you buy that kind of equipment, you may have seen a Bode'
plot used to communicate frequency response specifications.
All in all, Bode' plots are widely used, not just to specify or show a frequency
response, but they also give useful information for designing control systems.
Stability criteria can be interpreted on Bode' plots and there are numerous design
techniques based on Bode' plots.
You need to know how to use Bode' plots when you encounter them in those
situations, so this lesson will help you to understand the basics of Bode' plots.
What do you need to learn about Bode' plots? Here is a short summary:

What is a Bode' plot?


o How is magnitude plotted? (dbs)
o How is phase plotted? (degrees)
o How is frequency plotted? (on a logarithmic scale)
Given a Transfer Function:
o Be able to plot the Bode' plot, manually or with a math analysis
program.
o Know that the Bode' plot you generated "makes sense".
Given a Bode Plot for a System:
o Determine the transfer function of thesystem represented by the
Bode' plot.

What Are Bode' Plots?


Bode' plots are:
Plots of frequency response. Gain and phase are displayed in separate plots.
Logarithmic plots.
The horizontal axis is frequency - plotted on a log scale. It can be either f

or w.
The vertical axis is gain, expressed in decibels - a logarithmic measure of gain.
Sometimes, the vertical axis is simply a gain on a logarithmic scale.
Given these characteristics, you still need to know what a Bode' plot looks
like. Our strategy in this lesson will be to examine some simple systems - first
order and second order systems - to see what Bode' plots for the frequency
response of those systems look like. We'll start with the simplest system first,
and work from there. We will end by looking at how those simple systems can be
combined to make more complex systems with more complex Bode' plots.
Remember one of our goals above.

Given a Transfer Function:


o Be able to plot the Bode' plot, manually or with a math analysis
program.
Know that the Bode' plot you generated "makes sense".
That's what we will start with for first order systems.

Bode' Plots For First Order Systems


In this section we will work on that general goal for first order systems.
Let's look at an example Bode' plot for a first order system. Here's a plot for a
sample transfer function.
G(j) = 1/(j+ 1) with = .001
Here's the Bode' plot. Examine the following points for this plot.

The low frequency asymptote,


The high frequency asymptote,
The "mid point" where
= 1

That's at f = 159 Hz.


Let's look at the low frequency asymptote first. Here's the transfer
function.
G(j) = 1/(j+ 1)

If is small, then the imaginary term in the denominator is small,


and we have:
G(j) ~= 1/(j0 + 1) = 1
The low frequency behavior of the plot shows that the plot is flat at a
value of 1.
Now, let's look at the high frequency asymptote. Here's the transfer
function.
G(jw) = 1/(j + 1)
If is large, then the imaginary term in the denominator dominates, and
we have:
G(j ) ~= 1/j
The magnitude of the gain is:
|G(j)| ~= 1/
The gain drops off inversely with frequency, but the Bode' plot
drops off as a straight line. Hmmmm? That's very interesting - that it is

a straight line. The straight line high frequency asymptote shouldn't be


cause for consternation. If we have:
|G()| ~= 1/
Remember that the Bode' plot is log gain vs log frequency, so let's
look at the logarithm of the magnitude of the gain.
log(|G(j)|) = log(1/)
= -log()
= -log() - log()
So, log gain depends linearly upon the log of frequency () for higher
frequencies.
That's an important point to remember, and it is also a reason Bode'
plots are used so much. When the asymptotic behavior - both at high
frequencies and low frequencies - is straight line behavior, it makes Bode'
plots easier to sketchand easier to understand.
Actually, we need to note that the slope of this plot - at high
frequencies - is just -1. Look again at the asymptotic high frequency
relationship between the gain and frequency.
log(|G(j)|) = -log() - log()
When frequency increases by a factor of 10, log() increases by 1.
Therefore, when frequency increases by a
factor of 10, log(|G(j)|)decreases by 1.
Therefore, when frequency increases by a
factor of 10, |G(j)| decreases by a factor of 10.
From this discussion, we need to draw a conclusion.
When frequency increases by a factor of 10, |G(j)| decreases by a
factor of 10.

Check that conclusion on the plot to be sure you understand what it


means.
Here is a plot with the lower limit extended.

Check going from f = 300 to f = 3000.


Does the gain decrease by a factor of 10 when the frequency
increases by a factor of 10?
The last point we need to examine is the behavior of the frequency
response for frequencies between high frequency and low frequency what we referred to as the mid-point earlier. If the frequency response
function is given by:
G(j) = 1/(j + 1)

If = 1/then (taking that frequency as the mid-point), we have:


G(j) = 1/(j + 1)
The magnitude of the gain is:
|G(j)| = 1/|j + 1| = 1/sqrt(2)
~= 0.707
This point is at = 1000, or f =159Hz.
There are some interesting things to note about this frequency
response. Consider the interactive graph below. On that graph you can see
the low frequency asymptote, the high frequency asymptote and the point
where the gain is .707 of the low frequency gain.
Check the intersection of the two lines.
The intersection of the two lines occurs where = 1/.
For obvious reasons, this intersection is called the corner frequency.
Problem 1 What is the corner frequency for a system with this transfer
function?
G(s) = 22/( 17 s + 1)
Enter your answer in the box below, then click the button to submit your
answer. You will get a grade on a 0 (completely wrong) to 100 (perfectly
accurate answer) scale.

Your grade is:


Problem 2
Here is a Bode' plot like the one we have been examining. Determine
the corner frequency, in Hz, for this system.

Enter your answer in the box below, then click the button to submit your
answer.

Your grade is:

Problem 3
Here is another Bode' plot like the one we have been examining.
Determine the corner frequency, in Hz, for this system.

Enter your answer in the box below, then click the button to submit your
answer.
Your grade is:

There's one last point to observe regarding first order systems. The
general first order system has a transfer function of this form.
G(j) = Gdc/(j + 1)
The point to note is that there is a DC gain term in the numerator.
This really is the DC gain. Let the frequency, , be zero:
G(j0) = Gdc/(j0 + 1) = Gdc
The effect of DC gain is to raise or lower the entire plot. You need to
understand the effect of a DC gain on a Bode' plot. Let's look at the entire
transfer function.
G(j) = Gdc/(j + 1)
log(|G(j)|) = log(Gdc) - log(1/(()2 + 1))
This really says that log(Gdc) is added at every frequency.
Here is a movie where you can set the gain and see how the gain
changes the Bode' plot.

Adding log(Gdc) at every frequency shifts the entire plot up by log(Gdc).


Phase in 1st Order Bode' Plots
We have looked exclusively at the magnitude portion of the Bode'
plots we have examined. We need to look at the phase plot as well.
The transfer function is: G(j) = Gdc/(j + 1)
The phase angle at an angular frequency is: Angle(G(j)) = - tan-1(j)
The phase plot - against frequency - is important in many systems.
We will plot the phase for this transfer function- the one used earlier in
this section:
G(j) = 1/(j + 1) with = .001

Note the following:


The phase starts at 0o at low frequencies.
The phase goes to -90o at high frequencies.
The phase is -45o at a frequency of 159 Hz - the corner frequency.
There are several things to note at this point

Any transfer function is a ratio of polynomials - and those


polynomials have real coefficients.

Polynomials with real coefficients have real roots - first order


factors - and complex conjugate pairs of roots - second order
factors.
Our discussion of this first order system model is really only
addressing systems with one pole - one real root - in the
denominator.
More interesting systems have second order factors.

We'll start with second order factors in the denominator, i.e. second
order poles. We're not done with Bode' plots. Remember:

Our Bode' plots so far were all plotted with a log scale on the
vertical (gain) axis. Decibels are more often used and you need to
learn about them.
Second order systems have interesting Bode' plots - and it is
important to know about them. Click here to look at Bode' plots for
second order systems.

Decibels And More


When we introduced Bode' plots, we noted that the vertical scale of
a Bode' plot is often in terms of decibels. It's time you got acquainted
with decibels if you haven't heard of them before. Here's a start.
Originally, decibels were used to measure power gains.
If a system had an output power, Po, and an input power, Pi, then the
ratio of output power to input power - the power gain - is:
Po/Pi
The decibel gain is proportional to the logarithm - to the base ten (10) of the power gain

The gain can be expressed as the logarithm - to the base ten (10) - of
the power gain
Gain = log10(Po/Pi)
When expressed this way, the units are bels.
A decibel is one tenth of a bel, so the gain expressed in decibels is:
Gaindb = 10 log10(Po/Pi)
The unit bel is something of a story in itself.
Alexander Graham Bell did a lot of work with the deaf, and he was
recognized for his work with an honorary doctorate in 1880 by Gaulladet
College in Washington, D.C.(and he also delivered the commencement
address) He is more famous for his founding of the National Geographic
Society, and other work he did.
Alexander Graham Bell was also honored by having a unit named in his
honor - the bel.
Today, the decibel is a commonly used unit to measure sound intensity
and it is well known that high decibel levels contribute to deafness - a
very ironic closing of the circle.
Today, power is not so much an issue. We're more interested in voltage
gain of an amplifier. There's an interesting transition from power to
voltage that will help us understand how gain - expressed in decibels - is
viewed today.
In an amplifier, if the amplifier has an input resistance R1, then the
power input to the amplifier is given by:
Power In = V12/R1
Similarly, the output power into a resistor Ro is given by:

Power Out = Vo2/Ro


Now, look at the ratio of output power to input power:
Power Out/Power In = (Vo2/Ro)/(Vi2/Ri)
Now, compute the decibel gain:
Gaindb = 10 log10(Po/Pi) = 10 log10(Po/Pi)
= 10 log10(Vo2/Ro)/(Vi2/Ri)
= 10 log10(Vo2/Vi2) + 10 log10(Ri/Ro)
= 20 log10(Vo/Vi) + 10 log10(Ri/Ro)
The final result has a term in it that depends upon the resistors.
Gaindb = 20 log10(Vo/Vi) + 10 log10(Ri/Ro)
Today, engineers are often more concerned with things like voltage
gain. The resistances and power involved are not a concern at all when
analyzing control systems, so the resistance term is ignored, and we take
the gain, in db, of a system to be:
Gaindb = 20 log10(Vo/Vi)
We should realize that we can plot gain, in db, for a system as a
function of frequency. The ratio of output voltage to input voltage is
simply the ratio of output amplitude to input amplitude at some frequency
- our old friend, frequency response.
OK! You know about decibels. But there's some other things you
need to know about Bode' plots. The vertical axis on a true Bode' plot is
scaled in db. The horizontal axis is scaled using a logarithmic frequency
scale. Here's some not-so-obvious facts about the frequency scale.

An increase of frequency by a factor of 10 is referred to as a decade.


That's a fairly obvious reference. Ten years is a decade when speaking of
time. Our currency is based on a decimal system because it's based on
factors of 10.
An increase of frequency by a factor of 2 is referred to as an octave.
We're getting into the Latin and Greek roots here. Decade is based on a
Latin root - referring to the number 10. Octave is based on a classical
root referring to the number two - or is it? Right or wrong?
Wrong! Octave refers to eight, not two. The reason a doubling of
frequency is called an octave is that the musical world defined the term
far earlier than we ever thought of it. An octave is a doubling of
frequency, but it's eight notes in the scale to go up an octave.
Ok, now we're going to put this all together. Here's a Bode' plot for
a first order system. It has a DC gain of 20db, and a corner frequency
near f = 80 Hz. Now, look at the slope of the high frequency portion of
the plot.

Every decade increase causes the same decrease in dbs.


Actually, every octave increase causes equal decreases in dbs.
The slope appears to be -20 db/decade.

Check that this is the slope for any decade, from 1000 to 10,000 or
from 3000 to 30,000 Hz.

Not so obviously, the slope could be expressed as -6 db/octave.

If we go back to the transfer function for a first order system, we


can re-examine the high frequency behavior. Here's the transfer
function.
G(j) = 1/(j+ 1)

If is large (and only if it is large!), then the imaginary term in the


denominator dominates, and we have:
G(j) ~= 1/j
|G(j)| ~= 1/

log(|G(j)|) = log(1/)
= -log()
= -log() - log()
Express things in terms of decibels.
log(|G(j)|) = -log() - log()
Gaindb = 20 log10(|G(j)|) = -20 log() - 20 log()
Now, if we start with some frequency, o, we can calculate the gain at the
frequency.
Gaindb(o) = -20 log(o) - 20 log()
Now, take a frequency one decade higher, at 10o.
Gaindb(10o) = -20 log(10o) - 20 log()
We can calculate the difference in the db gain at these two frequencies.
Gaindb(10o) - Gaindb(o)
= [-20 log(10o) - 20 log()] - [-20 log(10o) - 20 log()]
The difference is:

Gaindb(10o) - Gaindb(o)
= -20 log(10o) + 20 log(o)
= -20 log(10) - 20 log(o) + 20 log(o)
= -20 log(10) = -20 db - in one decade!
Reflecting on the derivation above, we realize that this derivation says that
the slope is -20 db/decade for the high frequency asymptote in the Bode' plot.
It's also possible to express that another way. If we consider two frequencies
that are an octave apart, we can see that the slope can also be said to be 6db/octave.
The difference in the frequency response between the two frequencies is:

Gaindb(2wo) - Gaindb(wo)
= -20 log(2o) + 20 log(o)
= - 20 log(2) - 20 log(o) + 20 log(o)
= -20 log(2) = -6.0206 db - in one decade - and it's usually just rounded
to -6db/octave.
It's time to leave this topic. However consider this. We've only looked at one
first order system. Higher order systems - even second order systems - are bound
to have some differences in their Bode' plot behavior. High frequency asymptotes
will drop off at different slopes, for example, although we'll find that they drop
off at integral multiples of -20db/decade or -6 db/octave.
There are lots of interesting things you need to know, and you can start
looking at second order systems now.

Bode' Plots For 2nd Order Systems


We've looked at first order systems. Remember our general goal:

Given a Transfer Function:


o Be able to plot the Bode' plot, manually or with a math analysis
program.

Know that the Bode' plot you generated "makes sense".


Second order systems exhibit behavior that you will never see in a first order
system. We're going to work on that goal for second order systems - systems that
have this general transfer function.
If we have this transfer function:

A little reflection will probably tell you some things.


o For example, this system could have two complex roots.
It's not obvious, but to have two complex roots, the only thing necessary is
that the damping ratio, , be less than one.

Here's a Bode' plot for a second order system. This system has the following
parameters:

- the damping ratio = 0.1

n- the undamped natural frequency = 1000.

Gdc- the DC gain of the system = 1.0.

This system also has at least one unexpected feature - the "hump" in the
frequency response between f = 100 and f = 200 - a resonant peak. It's important

to understand how that peak in the frequency response comes about. Let's look at
the transfer function of a second order system. Here's a general form for such a
system. Examine how that system behaves for different frequencies.

Substitute s = j, to get the frequency response.

For small , the gain is just Gdc.

For large , the gain is Gdc/2.

That means that the high frequency gain drops off at -40 db/decade.

There are intermediate frequencies where interesting things happen!

We will start by looking at the interesting things that happen at the


intermediate frequencies. Here's the transfer function again, with s
replaced now by j.

We will examine what happens when = n.

At the natural frequency, the (j)2 term becomes -n2, cancelling out the

last term in the denominator, the n2 term, since j2 = -1.


Now, the really interesting things start to happen. When those terms cancel
the denominator just has one term left, and we have:

Now we can find an explanation for the hump in the frequency response.

The only term that involves the damping ratio is the one left in the
denominator when = n.

The damping ratio is in the denominator, so the smaller the damping ratio,
the larger the frequency response is going to be.

At = n, the magnitude of the frequency response function is:

or G(jn) =Gdc/j2

The formula for the gain of the frequency response at = n is interesting


because:

It depends only upon the DC gain and the damping ratio, and, the smaller the
damping ratio, the higher the gain at the natural frequency.

Now, recall the other important behavior at low frequencies and high
frequencies.

For small , the gain is just Gdc.

For large , the gain Gdc/2.

For small , the gain is just Gdc, assuming Gdc = 10 (or 20 db) on the plot.

For large , the gain is Gdc/2, - dropping off at -40 db/decade.

Here we assume that the natural frequency is fn = 20.

And, we can insert the point at the resonant frequency, using our formula.
o

G(jn) =Gdc/j2

For this example, we'll assume = 0.1. Remember:


Gdc= 10, and = 0.1,

so this works out to be a gain of 50 at the resonant peak, the equivalent of 34 db.
Do we have a problem here?

The peak is well above either of the asymptotes at the natural frequency.
We should believe all of the math we've done.
Is there really a problem here? Should we look at the actual frequency
response? Here it is. There's the peak. It does exist.

Let's examine the parameters here again to be sure that his all
hangs together. The system parameters were:

Gdc= 10,
= 0.1,

n = 2 20, (since that natural frequency was 20 Hz.)

With these paramters, note the following in the plot.

The DC gain is 20 db which corresponds to a gain of 10.


The resonant peak is pretty much right at 20 Hz as it should be.
The resonant peak is about 13 or 14 db high.
o A gain of 50 would be 14 db, do that also checks!
The high frequency slope looks to be around -12 db/octave or -20
db/decade.

All of these observations confirm the calculations, and they really


point out that it can be important to understand how the resonant peak
depends upon the damping ratio.
To make that correspondence between resonant peak and damping ratio as
clear as possible, we have here an example of a frequency response for another
system. We'll let you control the damping ratio, but we're going to set the DC gain

and the natural frequency. Hopefully, you'll see how this peak depends upon the
system's damping ratio. Use the right and left arrow controls to step the movie a
single step forward or backward.

Gdc = 1.0.
Natural frequency = 159 Hz.
Damping ratio - variable and controllable by user.
What should we note about the second order system response in the movie?

There is a resonant peak in the second order system response.


The size of the resonant peak depends upon the damping ratio.
For damping ratios less than about 0.5 the peak is relatively insignificant.

Finally, we have to deal with the phase. A Bode' plot isn't complete until you
have the phase plot. Here's a phase plot for a system with:

A damping ratio of 0.1


An undamped natural frequency of 159 Hz. (1000 rad/sec.)

Notice the following for this plot.

The phase starts at zero degrees for low frequencies.

The phase asymptotically approaches -180o for high frequencies.

How the phase plot depends upon damping ratio is something you should know.
Next, we have a movie of phase shift as a function of damping ratio.
For the system in the plot, the parameters are:

Gdc = 1.0.
Natural frequency = 159 Hz.
Damping ratio - variable.

Now, at this point you've seen Bode' plots for second order system with
complex poles. Second order systems with real poles are really combinations of
two first order systems, and they will be covered in the next section.
At this point, one direction to continue would be to continue to the next
section. However, you might want to go in the direction of looking at Nyquist plots
for the systems discussed above. In that case, use this link to go to the lesson on
Nyquist plots.
Nyquist Plots

Sketching Bode' Plots For Larger Systems - Examples


There will be times when you will need to have some sense of what a Bode'
plot looks like for a larger system. A useful skill is to be able to sketch what the
plot should look like so that you can anticipate what you'll get. That's particularly
helpful when you have a complex system and you enter a large transfer function.
It's not only helpful. You can often gain insight by playing "What if?" games with a
notepad and pencil.
In this section, we will look at some larger systems and examine some overall
properties of Bode' plots for those systems.
We will start with a system that is not all that large - a second order system
with two real poles. Just for discussion, we'll use the system with the transfer
function shown below.

If we wanted to sketch this Bode' plot we could start by looking at the DC


gain.
Remember that the DC gain is just G(jw) with = 0.
Letting = 0 in G(j), we get:
o G(0) = 10. (20db)
At low frequencies, the (.002s + 1) term in the denominator will still look
pretty much like 1.0.
However, as we go up in frequency, the (.01s + 1) term will have an effect.
The (.01s + 1) term introduces a corner frequency which we discussed
earlier in the section on Bode' plots for first order systems.
The corner frequency is at:
o f = 100/2 = 15.9Hz.

At slightly higher frequencies, the (.002s + 1) term will start to have an effect.

The (.002s + 1) term will add another -20db/decade slope to the plot, for a
total of -40 db.decade.
We get -40 db/decade because we now have two poles contributing to the
roll-off, and 2*(-20db/dec) = -40 db/dec.
The second corner frequency is at f = 500/2p = 79.5Hz.

The straight line approximation is high at the corners, but gives a pretty
good idea of where the actual Bode' plot lies.

Now, let us make this slightly more complicated. Here's another transfer
function.

Start by looking at the DC gain - as before.

Remember that the DC gain is just G(jw) with w = 0.


Letting w = 0 in G(j), we get:
o G(0) = 10. (20db)
As we go up in frequency from DC, the (.01s + 1) term will have an effect.

The (.01s + 1) term introduces a corner frequency - as before.


The corner frequency is at f = 100/2p = 15.9Hz.
Check the slope. It should be -20 db/decade.
At slightly higher frequencies, the (.002s + 1) term will start to have an
effect.
The (.002s + 1) term will add another -20db/decade - or wait a minute - is
that +20 db/decade?
Because it is a zero, it is +20db/dec, and the corner frequency is at:
o f = 500/2 = 79.5Hz.
For frequencies above 79.5 Hz, the gain would be 10*.002/.01 = 2 or 6db.

And don't forget we still have one more corner frequency. so let's add the last
corner frequency.

We have another corner frequency at:


o f = (1/.0001)/2 = 1590Hz. - Call that 1600 Hz.
Above 400 Hz, we have another -20 db/decade added, but the total will now
be -20 db/decade.

You might also like