Sunteți pe pagina 1din 123

PRINCIPLES OF ALTERNATING CURRENT

COMPLEX NUMBERS PASSIVE COMPONENTS IN AC CIRCUITS

A sinusoidal voltage can be described by the equation:

v (t) = VM sin ( t + ) or v (t) = VM cos ( t + )

where v (t) Instantaneous value of the voltage, in volts (V).


VM Maximum or peak value of the voltage, in volts (V)
T Period: The time taken for one cycle, in seconds
Frequency – the number of periods in 1 second, in Hz (Hertz) or
f
1/s. f = 1/T
Angular frequency, expressed in radians/s

 = 2**f or  = 2* / T.
Initial phase given in radians or degrees. This quantity determines

the value of the sine or cosine wave at t = 0.
Note: The amplitude of a sinusoidal voltage is sometimes
expressed as VEff, the effective or RMS value. This is related to
VM according to the relationship VM=2VEff, or approximately VEff =
0.707 VM

Here are a few examples to illustrate the terms above.

The properties of the 220 V AC voltage in household electrical outlets in Europe:

Effective value: VEff = 220 V


Peak value: VM = 2 220 V = 311 V
Frequency: f = 50 1/s = 50 Hz
Angular frequency: = 2**f = 314 1/s = 314 rad/s
Period: T = 1/f = 20 ms
Time function: v(t)=311 sin (314 t)

Let’s see the time function using TINA’s Analysis/AC Analysis/Time Function command.

Click here to load or save this circuit

You can check that the period is T=20m and that VM = 311 V.
The properties of the 120 V AC voltage in the household electrical outlet in the US:

Effective value: VEff = 120 V


Peak value: VM = 2 120 V = 169.68 V » 170 V
Frequency: f = 60 1/s = 60 Hz
Angular frequency: = 2**f = 376.8 rad/s  377 rad/s
Period: T = 1/f = 16.7 ms
Time function: v(t)=170 sin (377 t)

Note that in this case the time function could be given either as v(t)=311 sin (314 t+ ) or v(t)=311 cos (314
t+), since in the case of the outlet voltage we do not know the initial phase.

The initial phase plays an important role when several voltages are present simultaneously. A good practical example is
the three-phase system, where three voltages of the same peak value, shape and frequency are present, each of which
has a 120° phase shift relative to the others. In a 60 Hz network, the time functions are:

vA(t)=170 sin (377 t)

vB(t)=170 sin (377 t - 120°)

vC(t)=170 sin (377 t + 120°)

The following figure made with TINA shows the circuit with these time functions as TINA’s voltage generators.

Click here to load or save this circuit


The voltage difference vAB= vA(t) - vB(t) is shown as solved by TINA’s Analysis/AC Analysis/Time Function command.

Note that the peak of vAB (t) is approximately 294 V, larger than the 170 V peaks of the vA(t) or vB(t) voltages, but also not
simply the sum of their peak voltages. This is due to the phase difference. We will discuss how to calculate the resulting
voltage (which is 3 * 170  294 in this case) later in this chapter and also in the separate Three-phase Systems chapter.
Characteristic values of sinusoidal signals

Though an AC signal continuously varies during its period, it is easy to define a few characteristic values for comparing
one wave with another: These are the peak, average and root-mean-square (rms) values.

We have already met the peak value VM , which is simply the maximum value of the time function, the
amplitude of the sinusoidal wave.

Sometimes the peak-to-peak (p-p) value is used. For sinusoidal voltages and currents, the peak-to-peak value is double
the peak value.

The average value of the sine wave is the arithmetic average of the values for the positive half cycle. It is also
called absolute average since it is the same as the average of the absolute value of the waveform. In practice, we
encounter this waveform by rectifying the sine wave with a circuit called a full wave rectifier.

It can be shown that the absolute average of a sinusoidal wave is:

VAV = 2 /  VM  0.637 VM

Note that the average of a whole cycle is zero.


The rms or effective value of a sinusoidal voltage or current corresponds to the equivalent DC value producing the same
heating power. For example, a voltage with an effective value of 120 V produces the same heating and lighting power in a
light bulb as does 120 V from a DC voltage source. It can be shown that the rms or effective value of a sinusoidal wave is:

Vrms = VM / 2  0.707 VM

These values can be calculated the same way for both voltages and currents.

The rms value is very important in practice. Unless indicated otherwise, power line AC voltages (e.g. 110V or 220V) are
given in rms values. Most AC meters are calibrated in rms and indicate the rms level.

Example 1 Find the peak value of the sinusoidal voltage in an electrical network with 220 V rms value.
VM = 220/0.707 = 311.17 V

Example 2 Find the peak value of the sinusoidal voltage in an electrical network with 110 V rms value.

VM = 110/0.707 = 155.58 V

Example 3 Find the (absolute) average of the sinusoidal voltage if its rms value is 220 V.

Va = 0.637 * VM = 0.637*311.17 = 198.26 V

Example 4 Find the absolute average of the sinusoidal voltage if its rms value is 110 V.

The peak of the voltage from Example 2 is155.58 V and hence:

Va = 0.637 * VM = 0.637 * 155.58 = 99.13 V

Example 5 Find the ratio between the absolute average (V a) and rms (V) values for the sinusoidal waveform.

V/Va = 0.707/0.637 = 1.11

Note that you cannot add average values in an AC circuit because it leads to improper results.

PHASORS

As we have already seen in the previous section, it is often necessary in AC circuits to add sinusoidal voltages and
currents of the same frequency. Though it is possible to add the signals numerically using TINA, or by employing
trigonometric relations, it is more convenient to use the so-called phasor method. A phasor is a complex number
representing the amplitude and phase of a sinusoidal signal. It is important to note that the phasor does not represent the
frequency, which must be the same for all phasors.

A phasor can be handled as a complex number or represented graphically as a planar arrow in the complex plane. The
graphic representation is called a phasor diagram. Using phasor diagrams, you can add or subtract phasors in a complex
plane by the triangle or parallelogram rule.
There are two forms of complex numbers: rectangular and polar.

The rectangular representation is in the form a + jb, where j = -1 is the imaginary unit.

The polar representation is in the form Ae j , where A is the absolute value (amplitude) and f is the angle of the phasor
from the positive real axis, in the counterclockwise direction.

We will use bold letters for complex quantities.

Now let’s see how to derive the corresponding phasor from a time function.

First, assume that all the voltages in the circuit are expressed in the form of cosine functions. (All voltages can be
converted to that form.) Then the phasorcorresponding to the voltage of v(t) = VM cos(t+) is: VM = VMe j, which is also
called the complex peak value.

For example, consider the voltage: v(t) = 10 cos(t+30°)

The corresponding phasor is: V

We can calculate the time function from a phasor in the same way. First we write the phasor in polar form e.g. VM =
VMe j and then the corresponding time function is

v(t)=VM (cos(t+).

For example, consider the phasor VM =10 - j20 V

Bringing it to polar form:

And hence the time function is: v(t) = 22.36 cos(t – 63.5°) V
Phasors are often used to define the complex effective or rms value of the voltages and currents in AC circuits. Given v(t)
= VMcos(t+)= 10cos(t+30°)

Numerically:

v(t) = 10*cos (t-30°)

The complex effective (rms) value: V = 0.707*10* e- j30° = 7.07 e- j30° = 6.13 – j 3.535

Vice versa: if the complex effective value of a voltage is:

V = - 10 + j 20 = 22.36 e j 116.5°

then the complex peak value:

and the time function: v(t) = 31.63 cos ( t + 116.5° ) V

A short justification of the above techniques is as follows. Given a time function


VM (cos(t+r), let’s define the complex time function as:

v(t) =VM e j e jt = VMe jt = VM (cos() + j sin())e jt

where VM =VM e j t = VM (cos() + j sin()) is just the phasor introduced above.

For example, the complex time function of v(t)=10 cos( t+30°)

v(t) = VMe jt = 10 e j30 e jt = 10e jt (cos(30) + j sin(30))= e jt (8.66+j5)
By introducing the complex time function, we have a representation with both a real part and an imaginary part. We can always recover the original
real function of time by taking the real part of our result: v(t) = Re {v(t)}

However the complex time function has the great advantage that, since all the complex time functions in the AC circuits under consideration have
the same ejt multiplier, we can factor this out and just work with the phasors. Moreover, in practice we do not use the e jt part at all--just the
transformations from the time functions to the phasors and back.

To demonstrate the advantage of using phasors, let’s see the following example.

Example 6 Find the sum and the difference of the voltages:

v1 = 100 cos (314*t) and v2 = 50 cos (314*t-45°)

First write the phasors of both voltages:

V1M = 100 V2M= 50 e – j 45° = 35.53 – j 35.35

Hence:

Vadd = V1M + V2M = 135.35 – j 35.35 = 139.89 e- j 14.63°

Vsub = V1M – V2M = 64.65 + j35.35 = 73.68 e j 28.67°

and then the time functions:

vadd(t) = 139.89 * cos(t – 14.63°)

vsub(t) = 73.68 * cos(t + 28.67°)

As this simple example shows, the method of phasors.is an extremely powerful tool for solving AC problems.

Let’s solve the problem using the tools in TINA’s interpreter.

{calculation of v1+v2}
v1:=100
v2:=50*exp(-pi/4*j)
v2=[35.3553-35.3553*j]
v1add:=v1+v2
v1add=[135.3553-35.3553*j]
abs(v1add)=[139.8966]
radtodeg(arc(v1add))=[-14.6388]

{calculation of v1-v2}
v1sub:=v1-v2
v1sub=[64.6447+35.3553*j]
abs(v1sub)=[73.6813]
radtodeg(arc(v1sub))=[28.6751]

The amplitude and phase results confirm the hand


calculations.

Now lets check the result using TINA’s AC analysis.

Before performing the analysis, let’s make sure that


the Base function for AC ia set to cosine in the Editor
Options dialog box from the View/Option menu. We will
explain the role of this parameter at Example 8.

The circuits and the results:

Click here to load or save this circuit


Again the result is the same. Here are the time function graphs:
Example 7 Find the sum and the difference of the voltages:

v1 = 100 sin (314*t) and v2 = 50 cos (314*t-45°)

This example brings up a new question. So far we have required that all time functions be given as cosine functions. What
shall we do with a time function given as a sine? The solution is to transform the sine function to a cosine function. Using
the trigonometric relation sin(x)=cos(x-/2)=cos(x-90°), our example can be rephrased as follows:

v1 = 100 cos(314t - 90°) and v2 = 50 cos (314*t - 45°)

Now the phasors of the voltages are:

V1M = 100 e – j 90° = -100 j V2M= 50 e – j 45° = 35.53 – j 35.35

Hence:
Vadd = V1M + V2M = 35.53 – j 135.35

Vsub = V1M – V2M = - 35.53 – j 64.47

and then the time functions:

vadd(t) = 139.8966 cos(t-75.36°)

vsub(t) = 73.68 cos(t-118.68°)

Let’s solve the problem using the tools in TINA’s interpreter.

{calculation of v1+v2}
v1:=-100*j
v2:=50*exp(-pi/4*j)
v2=[35.3553 - 35.3553*j]
v1add:=v1+v2
v1add=[35.3553-135.3553*j]
abs(v1add)=[139.8966]
radtodeg(arc(v1add))=[-75.3612]
{calculation of v1-v2}
v1sub:=v1-v2
v1sub=[-35.3553 - 64.6447*j]
abs(v1sub)=[73.6813]
radtodeg(arc(v1sub))=[-118.6751]

Let’s check the result with TINA’s AC Analysis

Click here to load or save this circuit


Example 8 Find the sum and the difference of the voltages:

v1 = 100 sin (314*t ) and v2 = 50 sin (314*t-45°)

This example brings up one more issue. What if all voltages are given as sine waves and we also wish to see the result as
a sine wave?. We could of course convert both voltages to cosine functions, compute the answer, and than convert the
result back to a sine function--but this isn't necessary. We can create phasors from the sine waves in the same way that
we did from cosine waves and then simply use their amplitude and phases as amplitude and phase of sine waves in the
result.

This will obviously give the same result as transforming the sine waves to cosine waves. As we could see in the previous example, this is
equivalent to multiplying by –j and then using the cos(x) = sin (x-90°) relation to transform it back to a sine wave. This is equivalent to multiplying
by j. In other words, since –j × j = 1, we could use the phasors derived directly from the amplitudes and phases of sine waves to represent the
function and then return to them directly. Also, reasoning in the same manner about the complex time functions, we could consider sine waves as
the imaginary parts of the complex time functions and supplement them with the cosine function to create the full complex time function.

Let's see the solution to this example using the sine functions as the base of the phasors (transforming sin(t) to the real
unit phasor (1) ).

V1M = 100 V2M= 50 e – j 45° = 35.53 - j 35.35

Hence:

Vadd = V1M + V2M = 135.53 – j 35.35

Vsub = V1M – V2M = 64.47+ j 35.35

Note that the phasors are exactly the same as in Example 6 but not the time functions:

v3(t) = 139.9sin(t - 14.64°)

v4(t) = 73.68sin(t + 28.68°)

As you can see, it is very easy to obtain the result using sine functions, especially when our initial data are sine waves.
Many textbooks prefer to use the sine wave as the base function of phasors. In practice, you can use either method, but
don't confuse them.

When you create the phasors, it is very important that all time functions are first converted either to sine or
cosine. If you started from sine functions, your solutions should be represented with sine functions when
returning from phasors to time functions. The same is true if you start with cosine functions.

Let’s solve the same problem using TINA’s interactive mode. Since we want to use sine functions as the base for creating
the phasors, make sure that theBase function for AC is set to sine in the Editor Options dialog box from
theView/Option menu.
The circuits for making the sum and difference
of the waveforms and the result:
and the time funtions:
PASSIVE COMPONENTS IN AC CIRCUITS

PRINCIPLES OF ALTERNATING CURRENT


USING IMPEDANCE AND
ADMITTANCE

As we move from our study of DC circuits to AC circuits, we must consider two other types of passive component, ones that behave
very differently from resistors--namely, inductors and capacitors. Resistors are characterized only by their resistance and by Ohm’s
law. Inductors and capacitors change the phase of their current relative to their voltage and have impedances that depend upon
frequency. This makes AC circuits much more interesting and powerful. In this chapter, you will see how the use of phasors will
permit us to characterize all passive components (resistor, inductor, and capacitor) in AC circuits by their impedance and
the generalized Ohm’s law.

Resistor

When a resistor is used in an AC circuit, the variations of the current through and the voltage across the resistor are in
phase. In other words, their sinusoidal voltages and currents have the same phase. This in phase relationship can be
analyzed using the generalized Ohm’s law for the phasors of the voltage and current:

VM = R*IM or V = R*I

Obviously, we can use Ohm’s law simply for the peak or rms values (the absolute values of the complex phasors)--

VM = R*IM or V = R*I

but this form does not contain the phase information , which plays such an important role in AC circuits.

Inductor

An inductor is a length of wire, sometimes just a short trace on a PCB, sometimes a longer wire wound in the shape of a
coil with a core of iron or air.

The symbol of the inductor is L, while its value is called inductance. The unit of inductance is the henry (H), named after
the famous American physicist Joseph Henry. As inductance increases, so too does the inductor's opposition to the flow of
AC currents.
It can be shown that the AC voltage across an inductor leads the current by a quarter of a period. Viewed as phasors, the
voltage is 90 ahead (in a counterclockwise direction) of the current. In the complex plane, the voltage phasor is
perpendicular to the current phasor, in the positive direction (with respect to the reference direction, counterclockwise).
You can express this by complex numbers using an imaginary factor j as a multiplier.

The inductive reactance of an inductor reflects its opposition to the flow of AC current at a particular frequency, is
represented by the symbol XL, and is measured in ohms. Inductive reactance is calculated by the relationship X L = *L =
2**f*L. The voltage drop across an inductor is XL times the current. This relationship is valid for both the peak or rms
values of the voltage and current. In the equation for inductive reactance (X L ), f is frequency in Hz,  the angular
frequency in rad/s (radians/second), and L the inductance in H (Henry). So we have two forms of the generalized Ohm’s
law:

1. For the peak (VM, IM ) or effective (V,I) values of the current and the voltage:

VM = XL*IM or V = XL*I

2. Using complex phasors:

VM = j * X L I M or V = j * XL * I

The ratio between the voltage and current phasors of the inductor is its complex inductive impedance:

ZL= V/I = VM / IM = j  L

The ratio between the phasors of the current and voltage of the inductor is its complex inductive admittance:

YL= I/V = IM /VM =1/ (j  L)

You can see that the three forms of the generalized Ohm's law--ZL= V / I, I = V / ZL, and V = I * ZL--are very similar to
Ohm’s law for DC, except that they use impedance and complex phasors. Using impedance, admittance, and the
generalized Ohm’s law, we can treat AC circuits very similarly to DC circuits.
We can use Ohm’s law with the magnitude of inductive reactance just as we did for resistance. We simply relate the peak
(VM, IM) and rms (V, I) values of the current and voltage by XL, the magnitude of inductive reactance:

VM = XL IM or V = XL * I

However, since these equations do not include the phase difference between the voltage and current, they
shouldn’t be used unless phase is of no interest or is taken into account otherwise.

Proof

The time function of the voltage across a pure linear inductor (an inductor with zero internal
resistance and no stray capacitance) can be found by considering the time function that relates
voltage and current of the inductor:

Using the complex time function concept introduced in the previous chapter

Using complex phasors:

VL = j  L* IL

or with real time functions


vL (t) =  L iL (t+90)

so the voltage is 90 ahead of the current.

Let us demonstrate the proof above with TINA and show the voltage and the current as time functions and as phasors, in
a circuit containing a sinusoidal voltage generator and an inductor. First we will calculate the functions by hand.

The circuit we will study consists of a 1mH inductor connected to a voltage generator with sinusoidal voltage of 1Vpk and
a frequency of 100Hz (vL=1sin (t)=1sin(6.28*100t) V).

Using the generalized Ohm’s law, the complex phasor of the current is:

ILM= VLM/(jL) =1/(j6.28*100*0.001)=-j1.59A

and consequently the time function of the current:

iL(t)=1.59sin (t-90) A.

Now let’s demonstrate the same functions with TINA. The results are shown in the next figures.

Note on the use of TINA: We derived the time function using Analysis/AC Analysis/Time Function, while the phasor
diagram was derived usingAnalysis/AC Analysis/Phasor Diagram. We then used copy and paste to put the analysis
results on the schematic diagram. To show the amplitude and phase of the instruments on the schematic, we used AC
Interactive Mode.
The circuit diagram with the embedded time function and phasor diagram

Click here to load or save this


circuit

Time functions
Phasor diagram
Example 1

Find the inductive reactance and the complex impedance of an inductor with L = 3mH inductance, at a frequency f = 50
Hz.

XL = 2**f*L = 2*3.14*50*0.003 = 0.9425 ohm = 942.5 mohms

The complex impedance:

ZL= j  L = j 0.9425 = 0.9425 j ohms

You can check these results using TINA’s impedance meter. Set the frequency to 50Hz in the property box of the
impedance meter, which appears when you double click on the meter. The impedance meter will show the inductive
reactance of the inductor if you press the AC Interactive mode button as shown in the figure, or if you select
the Analysis/AC Analysis/Calculate nodal voltages command.

Using the Analysis/AC


Analysis/Calculate nodal
voltages command, you can
also check the complex
impedance measured by the
meter. Moving the pen-like tester
that appears after this command
and clicking on the inductor, you
will see the following table
showing the complex impedance
and admittance.

Note that both the impedance and the


admittance have a very small (1E-16) real part due to rounding errors in the calculation.

You can also show the complex impedance as a complex phasor using TINA’s AC Phasor Diagram. The result is shown in
the next figure. Use the Auto Label command to put the label showing the inductive reactance on the figure. Note that you
may need to change the automatic settings of the axes by double clicking to achieve the scales shown below.
Example 2

Find the inductive reactance of the 3mH inductor again, but this time at a frequency f = 200kHz.

XL = 2**f*L = 2*3.14*200*3 = 3769.91 ohms

As you can see, the inductive reactance rises with frequency.

Using TINA you can also plot the reactance as a function of the frequency.

Select the Analysis /AC Analysis/AC transfer and set the Amplitude and Phase checkbox. The following diagram will
appear:
In this diagram the Impedance is shown on a linear scale against frequency on a logarithmic scale. This conceals the fact
that the impedance is a linear function of frequency. To see this, double click on the upper frequency axis and set Scale to
Linear and Number of Ticks to 6. See the dialog box below:
Note that in some older version of TINA the phase diagram may show very small oscillations around 90
degrees due to rounding errors. You can eliminate this from the diagram by setting the vertical axis limit
similar to those shown in the figures above.

Capacitor
A capacitor consists of two conducting electrodes of metal separated by a dielectric (insulating) material. The capacitor
stores electric charge.

The symbol of the capacitor is C, and its capacity (or capacitance) is measured in farads (F), after the famous English
chemist and physicist Michael Faraday. As capacitance increases, the capacitor's opposition to the flow of AC
currents decreases. Furthermore, as frequency increases, the capacitor's opposition to the flow of AC currents decreases.

The AC current through a capacitor leads the AC voltage across the


capacitor by a quarter of period. Viewed as phasors, the voltage is 90 behind (in a counterclockwise direction) the
current. In the complex plane, the voltage phasor is perpendicular to the current phasor, in the negative direction (with
respect to the reference direction, counterclockwise). You can express this by complex numbers using an imaginary factor
-j as a multiplier.

The capacitive reactance of a capacitor reflects its opposition to the flow of AC current at a particular frequency, is
represented by the symbol XC, and is measured in ohms. Capacitive reactance is calculated by the relationship XC = 1/
(2**f*C) = 1/C. The voltage drop across a capacitor is XC times the current. This relationship is valid for both the peak or
rms values of the voltage and current. Note: in the equation for capacitive reactance (XC ), f is frequency in Hz,  the
angular frequency in rad/s (radians/second), C is the

in F (Farad), and XC is the capacitive reactance in ohms. So we have two forms of the generalized Ohm’s law:

1. For the absolute peak or effective values of the current and the voltage:

or V = XC*I
2. For the complex peak or effective values of the current and the voltage:

VM = -j * XC*IM or V = - j*XC*I

The ratio between the voltage and current phasors of the capacitor is its complex capacitive impedance:

ZC = V/I = VM / IM = - j*XC = - j / C

The ratio between the phasors of the current and voltage of the capacitor is its complex capacitive admittance:

YC= I/V = IM / VM = j C)

Proof:

The time function of the voltage across a pure linear capacitance (a capacitor with no parallel or
series resistance and no stray inductance) can be expressed using the time functions of the
capacitor’s voltage (vC), charge (qC) and current (iC ):

If C does not depend on time, using complex time functions:

iC(t) = j  C vC(t) or vC(t) = (-1/jC)*iC(t)

or using complex phasors:


or with real time functions

vc (t) = ic (t-90)/( C)

so the voltage is 90 behind the current.

Let us demonstrate the proof above with TINA and show the voltage and the current as functions of time, and as phasors.
Our circuit contains a sinusoidal voltage generator and a capacitor. First we will calculate the functions by hand.

The capacitor is 100nF and is connected across a voltage generator with sinusoidal voltage of 2V and a frequency of
1MHz : vL=2sin (t)=2sin(6.28*106t) V

Using the generalized Ohm’s law, the complex phasor of the current is:

ICM= jCVCM =j6.28*10610-7 *2)=j1.26A,

and consequently the time function of the current is:

iL(t)=1.26sin (t+90) A

so the current is ahead of the voltage by 90.

Now let us demonstrate the same functions with TINA. The results are shown in the next figures.
The circuit diagram with the embedded time function and phasor diagram

Click here to load or save this circuit


Time diagram
Phasor diagram

Example 3

Find the capacitive reactance and the complex impedance of a capacitor with C = 25 F capacitance, at a frequency f =
50 Hz.

XC = 1/ (2**f*C) = 1/(2*3.14*50*25*10-6) = 127.32 ohms

The complex impedance:

Z-C= 1/ (j  C) = - j 127.32 = -127.32 j ohms

Let’s check these results with TINA as we did for the inductor earlier.
You can also show the complex impedance as a complex phasor using TINA’s AC Phasor Diagram. The result is shown in the next
figure. Use the Auto Label command to put the label showing the inductive reactance on the figure. Note that you may need to change
the automatic settings of the axes by double clicking to achieve the scales shown below.
Example 4

Find the capacitive reactance of a 25 F capacitor again, but this time at frequency f = 200 kHz.

XC = 1/ (2**f*C) = 1/(2*3.14*200*103*25*10-6) = 0.0318 = 31.8 mohms.

You can see that the capacitive reactance decreases with frequency.
To see the frequency dependence of the impedance of a capacitor, let’s use TINA as we did earlier with the inductor.

Summarizing what we have


covered in this chapter,

The generalized Ohm’s law:

Z = V / I = VM/IM

The complex impedance for the


basic RLC components:

ZR =
R; ZL = j  L and Z
C = 1 / (j  C) = -j / C

We have seen how the


generalized form of Ohm's law
applies to all components--
resistors, capacitors, and
inductors. Since we have already
learned how to work with
Kirchoff's laws and Ohm's law for
DC circuits, we can build upon
them and use very similar rules
and circuit theorems for AC circuits. This will be described and demonstrated in the next chapters.

SING IMPEDANCE AND ADMITTANCE


PASSIVE COMPONENTS IN AC CIRCUITS
VOLTAGE AND CURRENT
DIVISION

As we saw in the previous chapter, impedance and admittance can be manipulated using the same rules as are used for
DC circuits. In this chapter we will demonstrate these rules by calculating total or equivalent impedance for series, parallel
and series-parallel AC circuits.

Example 1

Find the equivalent impedance of the following circuit:

R = 12 ohm, L = 10 mH, f = 159 Hz

Click here to load or save this circuit

The elements are in series, so we realise that their complex impedances should be added:

Zeq = ZR + ZL = R + j  L = 12 + j*2**159*0.01 = (12 + j 9.99) ohm = 15.6 ej39.8 ohm.

Yeq = 1/Zeq = 0.064 e- j 39.8 S = 0.0492 – j 0.0409 S


We can illustrate this result using impedance meters and the Phasor Diagram in
TINA v6. Since TINA’s impedance meter is an active device and we are going to use two of them, we must arrange the
circuit so that the meters don’t influence each other.
We have created another circuit just for the measurement of the part impedances. In this circuit, the two meters do not
“see” each other’s impedance.

The Analysis/AC Analysis/Phasor diagram command will draw the three phasors on one diagram. We used the Auto
Label command to add the values and the Line command of the Diagram Editor to add the dashed auxiliary lines for the
parallelogram rule.

The circuit for measuring the impedances of the parts

Click here to load or save this circuit


Phasor diagram showing the construction of Zeq with the parallelogram rule

As the diagram shows, the total impedance, Zeq, can be considered as a complex resultant vector derived
using the parallelogram rulefrom the complex impedances ZR and ZL .

Example 2

Find the equivalent impedance and admittance of this parallel circuit:

Click here to load or save this circuit


R =20 ohm, C = 5 F, f = 20 kHz

The admittance:

The impedance using the Ztot= Z1 Z2 / (Z1 + Z2 ) formula for parallel impedances:

Check your calculations using TINA’s Analysis menu Calculate nodal voltages. When you click on the Impedance meter, TINA
presents both the impedance and admittance and gives the results in algebraic and exponential forms.

Another way TINA can solve this problem is with its Interpreter:

om:=2*pi*20000;
Z:=Replus(R,(1/j/om/C))
Z=[125.8545m-1.5815*j]
Y:=1/R+j*om*C;
Y=[50m+628.3185m*j]

Example 3

Find the equivalent impedance of this parallel circuit. It uses the same elements as in Example 1:
R = 12 ohm and L = 10 mH, at f = 159 Hz frequency.

For parallel circuits, it’s often easier to calculate the admittance first:

Yeq = YR + YL = 1/R + 1/ (j*2**f*L) = 1/12 – j /10 = 0.0833 – j 0.1 = 0.13 e-j 50 S

Zeq = 1 / Yeq = 7.68 e j 50 ohm.


Click here to load or save this circuit

Another way TINA can solve this problem is with its Interpreter:

f:=159;
om:=2*pi*f;
Zeq:=replus(R,j*om*L);
Zeq=[4.9124+5.9006*j]

Example 4

Find the impedance of a series circuit with R = 10 ohm, C = 4 F, and L = 0.3 mH, at an angular frequency  = 50
krad/s (f =  / 2 = 7.957 kHz ).

Z = R + j  L - j / C = 10 + j 5*104 * 3*10-4 – j / (5*104 *4 * 10-6 ) = 10 + j 15 – j 5

Z = (10 + j 10) ohm = 14.14


ej 45 ohms.

The circuit for measuring the


impedances of the parts

Click here to load or save this circuit


Click here to load or save this circuit

The phasor diagram as generated by TINA

Starting with the phasor diagram above, let’s use the triangle or geometric construction rule to find the equivalent
impedance. We start by moving the tail ofZR to the tip of ZL. Then we move the tail of ZC to the tip of ZR. Now the
resultant Zeq will exactly close the polygon starting from the tail of the first ZR phasor and ending at the tip of ZC.
The phasor diagram showing the geometric construction of Zeq

{Solution by TINA's Interpreter}


om:=50k;
ZR:=R;
ZL:=om*L;
ZC:=1/om/C;
Z:=ZR+j*ZL-j*ZC;
Z=[10+10*j]
abs(Z)=[14.1421]
radtodeg(arc(Z))=[45]
{other way}
Zeq:=R+j*om*L+1/j/om/C;
Zeq=[10+10*j]
Abs(Zeq)=[14.1421]
fi:=arc(Z)*180/pi;
fi=[45]

Check your calculations using TINA’s Analysis menu Calculate nodal voltages. When you click on the Impedance
meter, TINA presents both the impedance and admittance, and gives the results in algebraic and exponential forms.

Since the circuit’s impedance has a positive phase like an inductor, we can call it an inductive circuit--at least at this
frequency!

Example 5

Find a simpler series network that could replace the series circuit of example 4 (at the given frequency).

We noted in Example 4 that the network is inductive, so we can replace it by a 4 ohm resistor and a 10 ohm inductive
reactance in series:

XL = 10 = *L = 50*103 L

 L = 0.2 mH
Click here to load or save this circuit

Don’t forget that, since inductive reactance depends upon frequency, this equivalence is valid only for one frequency.

Example 6

Find the impedance of three components connected in parallel: R = 4 ohm, C = 4 F, and L = 0.3 mH, at an angular
frequency  = 50 krad/s (f =  / 2 = 7.947 kHz).

Click here to load or save this circuit

Noting that this is a parallel circuit, we solve first for the admittance:

1/Z = 1/R +1/ j  L + jC = 0.25 – j/15 +j0.2 = 0.25 +j 0.1333

Z = 1/(0.25 + j 0.133) = (0.25 – j 0.133)/0.0802 = 3.11 – j 1.65 =3.5238 e-j 28.1 ohms.

{Solution by TINA's Interpreter}


om:=50k;
ZR:=R;
ZL:=om*L;
ZC:=1/om/C;
Z:=1/(1/R+1/j/ZL-1/j/ZC);
Z=[3.1142-1.6609*j]
abs(Z)=[3.5294]
fi:=radtodeg(arc(Z));
fi=[-28.0725]

The Interpreter calculates phase in radians. If you want phase in degrees, you can convert from radians to degrees by
multiplying by 180 and dividing by . In this last example, you see a simpler way—use the Interpreter’s built in function,
radtodeg. There is an inverse function as well, degtorad. Note that this network’s impedance has a negative phase like a
capacitor, so we say that—at this frequency—it is a capacitive circuit.

In Example 4 we placed three passive components in series, while in this example we placed the same three elements in
parallel. Comparing the equivalent impedances calculated at the same frequency, reveals that they are totally different,
even their inductive or capacitive character.

Example 7

Find a simple series network that could replace the parallel circuit of example 6 (at the given frequency).

This network is capacitive because of the negative phase, so we try to replace it with a series connection of a resistor and
a capacitor:

Zeq = (3.11 – j 1.66) ohm = Re –j / Ce


Click here to load or save this circuit

Re = 3.11 ohm *C = 1/1.66 = 0.6024

hence

Re = 3.11 ohm
C = 12.048 F

You could, of course, replace the parallel circuit with a simpler parallel circuit in both examples

Example 8

Find the equivalent impedance of the following more complicated circuit at frequency f=50 Hz:
Click here to load or save this circuit

{Solution by TINA's interpreter}


om:=2*pi*50;
Z1:=R3+j*om*L3;
Z2:=replus(R2,1/j/om/C);
Zeq:=R1+Replus(Z1,Z2);
Zeq=[55.469-34.4532*j]
abs(Zeq)=[65.2981]
radtodeg(arc(Zeq))=[-31.8455]

We need a strategy before we begin. First we’ll reduce C and R2 to an equivalent impedance, Z RC. Then, seeing that ZRC is in parallel
with the series-connected L3 and R3, we’ll compute the equivalent impedance of their parallel connection, Z 2. Finally, we calculate
Zeq as the sum of Z1 and Z2.

Here’s the calculation of ZRC:

Here’s the calculation of Z2:


And finally:

Zeq = Z1 + Z2 = (55.47 – j 34.45) ohm = 65.3 e-j31.8 ohm

according to TINA’s result.

VOLTAGE AND CURRENT DIVISION

USING IMPEDANCE AND ADMITTANCE


SUPERPOSITION IN AC
CIRCUITS

We have already shown how the elementary methods of DC circuit analysis can be extended and used in AC circuits to
solve for the complex peak or effective values of voltage and current and for complex impedance or admittance. In this
chapter, we'll solve some examples of voltage and current division in AC circuits.

Example 1

Find the voltages v1(t) and v2(t), given that vs(t)=110cos(250t).


Click here to load or save this circuit

Let's first obtain this result by hand calculation using the voltage division formula.

The problem can be considered as two complex impedances in series: the impedance of the resistor R1, Z1=R1 ohms (which is a real
number), and the equivalent impedance of R2 and L2 in series, Z2 = R2 + j  L2.

Substituting the equivalent impedances, the circuit can be redrawn in TINA as follows:
Note that we have used a new component, a complex impedance, now available in TINA v6. You can define the frequency dependence
of Z by means of a table that you can reach by double clicking the impedance component. In the first row of the table you can define
either the DC impedance or a frequency independent complex impedance (we have done the latter here, for the inductor and resistor in
series, at the given frequency).

Using the formula for voltage division:

V1 = Vs*Z1 / (Z1 + Z2)

V2 = Vs*Z2 / (Z1 + Z2)

Numerically:

Z1 = R1 = 10 ohms

Z2 = R2 + j  L = 15 + j 2** 50*0.04 =15 + j 12.56 ohms

V1= 110*10/ (25+j12.56) = 35.13-j17.65 V = 39.31 e -j26.7  V

V2= 110*(15+j12.56)/ (25+j12.56) = 74.86+j17.65 V = 76.92 e j 13.3 V

The time function of the voltages:

v1(t) = 39.31 cos (t - 26.7) V

v2(t) = 76.9 cos (t + 13.3) V

Let's check the result with TINA using Analysis/AC Analysis/Calculate nodal voltages

V1 V2
Next let's check these results with TINA's Interpreter:

{Solution by TINA's Interpreter}


f:=50;
om:=2*pi*f;
VS:=110;
v1:=VS*R1/(R1+R2+j*om*L2);
v2:=VS*(R2+j*om*L2)/(R1+R2+j*om*L2);
v1=[35.1252-17.6559*j]
v2=[74.8748+17.6559*j]
abs(v2)=[76.9283]
radtodeg(arc(v2))=[13.2683]
abs(v1)=[39.313]
radtodeg(arc(v1))=[-26.6866]

Note that when using the Interpreter we did not have to declare the values of the passive components. This is because we are using the
Interpreter in a work session with TINA in which the schematic is in the schematic editor. TINA's Interpreter looks in this schematic
for the definition of the passive component symbols entered into the Interpreter program.

Finally, let's use TINA's Phasor Diagram to demonstrate this result. Connecting a voltmeter to the voltage generator,
selecting the Analysis/AC Analysis/Phasor Diagram command, setting the axes, and adding the labels, will yield the
following diagram. Note that View/Vector label style was set to Amplitude for this diagram.
The diagram shows that Vs is the sum of the phasors V1 and V2, Vs = V1 + V2.

By moving the phasors we can also demonstrate that V2 is the difference between Vs and V1, V2 = Vs – V1.
This figure also demonstrates the subtraction of vectors. The resultant vector should start from the tip of the second vector, V1.

In a similar way we can demonstrate that V1 = Vs – V2. Again, the resultant vector should start from the tip of the second vector, V1.

Of course, both phasor diagrams can be considered as a simple triangle rule diagram for Vs = V1 + V2 .

The phasor diagrams above also demonstrate Kirchhoff's voltage law (KVL).

As we have learned in our study of DC circuits, the applied voltage of a series circuit equals the sum of the voltage drops
across the series elements. The phasor diagrams demonstrate that KVL is also true for AC circuits, but only if we use
complex phasors!
Example 2

In this circuit, R1 represents the DC resistance of the coil L; together they model a real world inductor with its loss component. Find
the voltage across the capacitor and the voltage across the real world coil.

L = 1.32 h, R1 = 2 kohms, R2 = 4 kohms, C = 0.1 F, vS(t) = 20 cos (t) V, f = 300Hz.

Click here to load or save this circuit

V1 V2

Solving by hand using voltage division:


= 13.91 e j 44.1 V

and

v1(t) = 13.9 cos (t + 44) V

= 13.93 e -j 44.1 V

and

v2(t) = 13.9 cos(t - 44.1) V

Notice that at this frequency, with these component values, the magnitudes of the two voltages are nearly the same, but the phases are
of opposite sign.

Once again, let's have TINA do the tedious work by solving for V1 and V2 with the Interpreter:

{Solution by TINA's Interpreter!}


om:=600*pi;
V:=20;
v1:=V*(R1+j*om*L)/(R1+j*om*L+replus(R2,(1/j/om/C)));
abs(v1)=[13.9301]
180*arc(v1)/pi=[44.1229]
v2:=V*(replus(R2,1/j/om/C))/(R1+j*om*L+replus(R2,(1/j/om/C)));
abs (v2)=[13.9305]
180*arc(v2)/pi=[-44.1211]

And finally, take a look at this result using TINA's Phasor Diagram. Connecting a voltmeter to the voltage generator, invoking
the Analysis/AC Analysis/Phasor Diagramcommand, setting the axes, and adding the labels will yield the following diagram (note
that we have set View/Vector label style to Real+j*Imag for this diagram):

Example 3
The current source iS(t) = 5 cos (t) A, the resistor R = 250 mohm, the inductor L = 53 uH, and the frequency f = 1
kHz. Find the current in the inductor and the current in the resistor.

Click here to load or save this circuit

IR IL

Using the formula for current division:

iR(t) = 4 cos (t + 37.2) A

Similarly:
iL(t) = 3 cos(t - 53.1)

And using the Interpreter in TINA:

{Solution by TINA's Interpreter}


om:=2*pi*1000;
is:=5;
iL:=is*R/(R+j*om*L);
iL=[1.8022-2.4007*j]
iR:=is*j*om*L/(R+j*om*L);
iR=[3.1978+2.4007*j]
abs(iL)=[3.0019]
radtodeg(arc(iL))=[-53.1033]
abs(iR)=[3.9986]
radtodeg(arc(iR))=[36.8967]

We can also demonstrate this solution with a phasor diagram:


The phasor diagram shows that the generator current IS is the resultant vector of the complex currents IL and IR. It also
demonstrates Kirchhoff's current law (KCL), showing that the current IS entering the upper node of the circuit equals
the sum of IL and IR, the complex currents leaving the node.

Example 4

Determine i0(t), i1(t) and i2(t). The component values and the source voltage, frequency, and phase are given on the schematic below.
Click here to load or save this circuit

i0

i1
i2

In our solution, we will use the principle of current division. First we find the expression for the total current i0:

I0M = 0.315 e j 83.2 A and i0(t) = 0.315 cos (t + 83.2) A

Then using current division, we find the current in the capacitor C:

I1M = 0.524 e j 91.4 A and i1(t) = 0.524 cos(t + 91.4) A

And the current in the inductor:


I2M = 0.216 e-j 76.6 A and i2(t) = 0.216 cos(t - 76.6) A

With anticipation, we seek confirmation of our hand calculations using TINA's Interpreter.

{Solution by TINA's Interpreter}


V:=10;
om:=2*pi*1000;
I0:=V/((1/j/om/C1)+replus((1/j/om/C),(R+j*om*L)));
I0=[37.4671m+313.3141m*j]
abs(I0)=[315.5463m]
180*arc(I0)/pi=[83.1808]
I1:=I0*(R+j*om*L)/(R+j*om*L+1/j/om/C);
I1=[-12.489m+523.8805m*j]
abs(I1)=[524.0294m]
180*arc(I1)/pi=[91.3656]
I2:=I0*(1/j/om/C)/(R+j*om*L+1/j/om/C);
I2=[49.9561m-210.5665m*j]
abs(I2)=[216.4113m]
180*arc(I2)/pi=[-76.6535]
{Control: I1+I2=I0}
abs(I1+I2)=[315.5463m]

Another way of solving this would be to first find the voltage across the parallel complex impedance of Z LR and ZC.
Knowing this voltage, we could find the currents i1 and i2 by then dividing this voltage first by ZLR and then by ZC. We will
show next the solution for voltage across the parallel complex impedance of Z LR and ZC. We will have to use the voltage
division principal along the way:
VRLCM = 8.34 e j 1.42 V

and
IC = I1= VRLCM*jC = 0.524 e j 91.42 A

and hence

iC (t) = 0.524 cos (t + 91.4) A.

SUPERPOSITION IN AC CIRCUITS

VOLTAGE AND CURRENT

THÉVENIN AND NORTON EQUIVALENT


DIVISION CIRCUITS

We have already studied the superposition theorem for DC circuits. In this chapter we will show its application for AC
circuits.

The superposition theorem states that in a linear circuit with several sources, the current and voltage for any element in the circuit is
the sum of the currents and voltages produced by each source acting independently. The theorem is valid for any linear circuit. The
best way to use superposition with AC circuits is to calculate the complex effective or peak value of the contribution of each source
applied one at a time, and then to add the complex values. This is much easier than using superposition with time functions, where one
has to add the individual time functions.

To calculate the contribution of each source independently, all the other sources must be removed and replaced without affecting the
final result.

When removing a voltage source, its voltage must be set to zero, which is equivalent to replacing the voltage source with a short
circuit.

When removing a current source, its current must be set to zero, which is equivalent to replacing the current source with an open
circuit.

Now let's explore an example.


In the circuit shown below"

Ri = 100 ohm, R1 = 20 ohm, R2 = 12 ohm, L = 10 uH, C = 0.3 nF, vS(t)=50cos(t) V, iS(t)=1cos(t+30°) A, f=400 kHz.

Notice that both sources have the same frequency: we will only work in this chapter with sources all having the same
frequency. Otherwise, superposition must be handled differently.

Find the currents i(t) and i1(t) using the superposition theorem.

Click here to load or save this circuit

Let's use TINA and hand calculations in parallel to solve the problem.

First substitute an open circuit for the current source and calculate the complex phasors I', I1' due to the contribution only
from VS.

The currents in this case are equal:

I' = I1' = VS/(Ri + R1 + j**L) = 50/(120+j2**4*105*10-5) = 0.3992-j0.0836

I' = 0.408 ej 11.83 A


Next substitute a short-circuit for the voltage source and calculate the complex phasors I'', I1'' due to the contribution only
from IS.

In this case we can use the current division formula:

I'' = -0.091 - j 0.246 A


and

I1" = 0.7749 + j 0.2545 A

The sum of the two steps:

I = I' + I" = 0.3082 - j 0.3286 = 0.451 e- j 46.9 A

I1 = I1" + I1' = 1.174 + j 0.1709 = 1.1865 ej 8.28 A

These results correspond well with the values calculated by TINA:

The time functions of the currents:

i(t) = 0.451 cos (t - 46.9) A


i1(t) = 1.1865 cos (t + 8.3) A

Similarly, the results given by TINA's Interpreter also agree:

{Solution by TINA's Interpreter}


f:=400000;
Vs:=50;
IG:=1*exp(j*pi/6);
om:=2*pi*f;
sys I,I1
I+IG=I1
Vs=I*Ri+I1*(R1+j*om*L)
end;
I=[308.093m-329.2401m*j]
abs(I)=[450.9106m]
radtodeg(arc(I))=[-46.9004]
abs(I1)=[1.1865]
radtodeg(arc(I1))=[8.2749]

As we said in the DC chapter on superposition, it gets pretty complicated using the superposition theorem for circuits
containing more then two sources. While the superposition theorem can be useful for solving simple practical problems,
its main use is in the theory of circuit analysis, where it is employed in proving other theorems.

THÉVENIN AND NORTON EQUIVALENT CIRCUITS

SUPERPOSITION IN AC CIRCUITS
KIRCHHOFF'S LAWS IN AC
CIRCUITS
Thévenin's Theorem for AC circuits with sinusoidal sources is very similar to the theorem we have learned for DC circuits. The only
difference is that we must considerimpedance instead of resistance. Concisely stated, Thévenin's Theorem for AC circuits says:

Any two terminal linear circuit can be replaced by an equivalent circuit consisting of a voltage source (V Th) and a series
impedance (ZTh).

In other words, Thévenin's Theorem allows one to replace a complicated circuit with a simple equivalent circuit containing only a
voltage source and a series connected impedance. The theorem is very important from both theoretical and practical viewpoints.

It is important to note that the Thévenin equivalent circuit provides equivalence at the terminals only. Obviously, the internal structure
of the original circuit and the Thévenin equivalent may be quite different. And for AC circuits, where impedance is frequency
dependent, the equivalence is valid at one frequency only.

Using Thévenin's Theorem is especially advantageous when:

 we want to concentrate on a specific portion of a circuit. The rest of the circuit can be replaced by a simple Thévenin
equivalent.

 we have to study the circuit with different load values at the terminals. Using the Thévenin equivalent we can avoid having to
analyze the complex original circuit each time.

We can calculate the Thévenin equivalent circuit in two steps:

1. Calculate ZTh. Set all sources to zero (replace voltage sources by short circuits and current sources by open circuits) and then
find the total impedance between the two terminals.

2. Calculate VTh. Find the open circuit voltage between the terminals.

Norton's Theorem, already presented for DC circuits, can also be used in AC circuits. Norton's Theorem applied to AC circuits states
that the network can be replaced by acurrent source in parallel with an impedance.

We can calculate the Norton equivalent circuit in two steps:


1. Calculate ZTh. Set all sources to zero (replace voltage sources by short circuits and current sources by open circuits) and then
find the total impedance between the two terminals.

2. Calculate ITh. Find the short circuit current between the terminals.

Now let's see some simple examples.

Example 1

Find the Thévenin equivalent of the network for the points A and B at a frequency: f = 1 kHz, vS(t) = 10 cost V.

Click here to load or save this circuit

The first step is to find the open circuit voltage between points A and B:

The open circuit voltage using voltage division:

= -0.065 - j2.462 = 2.463 e-j91.5º V

Checking with TINA:


The second step is to replace the voltage source by a short circuit and to find the impedance between points A and B:

Of course, we can check our ZT solution using TINA's impedance meter (note that we have replaced the voltage source with a short
circuit):
Here is the Thévenin equivalent circuit, valid only at a frequency of 1kHz. We must first, however, solve for CT's capacitance. Using
the relationship 1/CT = 304 ohm, we find CT = 0.524 uF

Now we have the solution: RT = 301 ohm and CT = 0.524  F:

Next, we can use TINA's interpreter to check our calculations of the Thévenin equivalent circuit:

{Solution by TINA's Interpreter}


VM:=10;
f:=1000;
om:=2*pi*f;
Z1:=R1+j*om*L;
Z2:=R2/(1+j*om*C*R2);
VT:=VM*Z2/(Z1+Z2);
VT=[-64.0391m-2.462*j]
abs(VT)=[2.4629]
abs(VT)/sqrt(2)=[1.7415]
radtodeg(arc(VT))=[-91.49]
ZT:=Replus((R1+j*om*L),replus(R2,(1/j/om/C)));
ZT=[301.7035-303.4914*j]
Abs(ZT)=[427.9393]
radtodeg(arc(ZT))=[-45.1693]
Ct:=-1/im(ZT)/om;
Ct=[524.4134n]

Note that in the listing above we used a function "replus.' Replus solves for the parallel equivalent of two impedances; i.e., it finds the
product over the sum of the two parallel impedances.

Example 2

Find the Norton equivalent of the circuit in Example 1.

f = 1 kHz, vS(t) = 10 cost V.

Click here to load or save this circuit

The equivalent impedance is the same:


ZN=(0.301-j0.304) k

Next, find the short-circuit current:

IN = (3.97-j4.16) mA

And we can check our hand calculations against TINA's results. First the open circuit impedance:
Then the short-circuit current:

And finally the Norton equivalent:

Next, we can use TINA's interpreter to find the Norton equivalent circuit components:

{Solution by TINA's Interpreter}


VM:=10;
f:=1000;
om:=2*pi*f;
Z1:=R1+j*om*L;
Z2:=R2/(1+j*om*C*R2);
IN:=VM/Z1;
IN=[3.9746m-4.1622m*j]
abs(IN)=[5.7552m]
abs(IN)/sqrt(2)=[4.0695m]
radtodeg(arc(IN))=[-46.3207]
ZN:=Replus((R1+j*om*L),replus(R2,(1/j/om/C)));
ZN=[301.7035-303.4914*j]
Abs(ZN)=[427.9393]
radtodeg(arc(ZN))=[-45.1693]
CN:=-1/im(ZN)/om;
CN=[524.4134n]

Example 3

In this circuit, the load is the series-connected RL and CL. These load components are not part of the circuit whose equivalent we are
seeking. Find the current in the load using the Norton equivalent of the circuit.

v1(t) = 10 cos t V; v2(t) = 20 cos (t+30) V; v3(t) = 30 cos (t+70) V;

v4(t) = 15 cos (t+45) V; v5(t) = 25 cos (t+50) V; f = 1 kHz.


Click here to load or save this circuit

First find the open circuit equivalent impedance Zeq by hand (without the load).

Numerically
ZN = Zeq = (13.93 - j5.85)
ohm.

Below we see TINA's solution. Note that we replaced all the voltage sources with short circuits before we used the
meter.

Now the short-circuit current:

The calculation of the short-circuit current is quite complicated. Hint: this would be a good time to use Superposition. An approach
would be to find the load current (in rectangular form) for each voltage source taken one at a time. Then sum the five partial results to
get the total.

We will just use the value provided by TINA:

iN(t) = 2.77 cos (t-118.27) A


Putting it all together (replacing the network with its Norton equivalent, reconnecting the load components to the
output, and inserting an ammeter in the load), we have the solution for the load current that we sought:

By hand calculation, we could find the load current using current division:
Finally

I = (- 0.544 - j 1.41) A

and the time function

i(t) = 1.51 cos (t - 111.1) A

KIRCHHOFF'S LAWS IN AC CIRCUITS

THÉVENIN AND NORTON

NODE POTENTIAL AND MESH


EQUIVALENT CIRCUITS CURRENT METHOD IN AC CIRCUITS

As we have already seen, circuits with sinusoidal excitation can be solved using complex impedances for the elements
and complex peak or complex rms values for the currents and voltages. Using the complex values version of Kirchhoff's
laws, nodal and mesh analysis techniques can be employed to solve AC circuits in a manner similar to DC circuits. In this
chapter we will show this through examples of Kirchhoff's laws.

Example 1

Find the amplitude and phase angle of the current i vs(t) if


vS(t) = VSM cos 2ft; i(t) =ISM cos 2ft; VSM = 10 V; ISM = 1 A; f = 10 kHz;

R = 5 ohm; L =0.2 mH; C1 = 10 F; C2 = 5 F


Click here to load or save this circuit

Altogether we have 10 unknown voltages and currents, namely: i, iC1, iR, iL, iC2, vC1, vR, vL, vC2 and vIS. (If we
use complex peak or rms values for the voltages and currents, we have altogether 20 real equations!)

The equations:

Loop or mesh equations: for M1 - VSM +VC1M+VRM = 0

M2 - VRM + VLM = 0

M3 - VLM + VC2M = 0

M4 - VC2M + VIsM = 0

Ohm's laws VRM = R*IRM

VLM = j**L*ILM

IC1M = j**C1*VC1M

IC2M = j**C2*VC2M

Nodal equation for N1 - IC1M - ISM + IRM + ILM +IC2M = 0


for series elements I = IC1M

Solving the system of equations you can find the unknown current:

ivs (t) = 1.81 cos (t + 79.96) A

Solving such a large system of complex equations is very complicated, so we haven't shown it in
detail. Each complex equation leads to two real equations, so we show the solution only by the values
calculated with TINA's Interpreter.

The solution using TINA's Interpreter:


{Solution by TINA's Interpreter}
om:=20000*pi;
Vs:=10;
Is:=1;
Sys Ic1,Ir,IL,Ic2,Vc1,Vr,VL,Vc2,Vis,Ivs
Vs=Vc1+Vr {M1}
Vr=VL {M2}
Vr=Vc2 {M3}
Vc2=Vis {M4}
Ivs=Ir+IL+Ic2-Is {N1}
{Ohm's rules}
Ic1=j*om*C1*Vc1
Vr=R*Ir
VL=j*om*L*IL
Ic2=j*om*C2*Vc2
Ivs=Ic1
end;
Ivs=[3.1531E-1+1.7812E0*j]
abs(Ivs)=[1.8089]
fiIvs:=180*arc(Ivs)/pi
fiIvs=[79.9613]

The solution using TINA:


Click here to load or save this circuit

Click here to load or save this circuit

To solve this problem by hand, work with the complex impedances. For example, R, L and C2 are connected
in parallel, so you can simplify the circuit by computing their parallel equivalent. || means the parallel
equivalent of the impedances:
Numerically:

Click here to load or save this circuit

The simplified circuit using the impedance:

The equations in ordered form : I + IG1 = IZ

VS = VC1 +VZ

VZ = Z · IZ

I = j  C1· VC1

There are four unknowns- I; IZ; VC1; VZ - and we have four equations, so a solution is possible.

Express I after substituting the other unknowns from the equations:


Numerically

Click here to load or save this circuit

According to TINA's Interpreter's result.


{Solution using the impedance Z}
om:=20000*pi;
Vs:=10;
Is:=1;
Z:=replus(R,replus(j*om*L,1/j/om/C2));
Z=[2.1046E0-2.4685E0*j]
sys I
I=j*om*C1*(Vs-Z*(I+Is))
end;
I=[3.1531E-1+1.7812E0*j]
abs(I)=[1.8089]
180*arc(I)/pi=[79.9613]

The time function of the current, then, is:

i(t) = 1.81 cos (t + 80) A

You can check Kirchhoff's current rule using phasor diagrams. The picture below was developed by
checking the node equation in iZ = i + iG1 form. The first diagram shows the phasors added by parallelogram
rule, the second one illustrates the triangular rule of the phasor addition.
Now let's demonstrate KVR using TINA's phasor diagram feature. Since the source voltage is negative in the
equation, we connected the voltmeter "backwards." The phasor diagram illustrates the original form of the
Kirchhoff's voltage rule.
Click here to load or save this circuit

The first phasor diagram uses the parallelogram rule, while the second uses the triangular rule.
To illustrate KVR in the form VC1 + VZ - VS = 0, we again connected the voltmeter to the voltage source backwards. You
can see that the phasor triangle is closed.
Note that TINA lets you use either sine or cosine function as a base function. Depending on the function chosen, the
complex amplitudes seen in phasor diagrams may differ by 90º. You can set the base function under 'View' 'Options' 'Base
function for AC'. In our examples we always used cosine function as a base.

Example 2

Find the voltages and currents of all the components if:

vS(t) = 10 cos t V, iS(t) = 5 cos ( t + 30°) mA;

C1 = 100 nF, C2 = 50 nF, R1 = R2 = 4 k; L = 0.2 H, f = 10 kHz.


Click here to load or save this circuit

Let the unknowns be the complex peak values of the voltages and currents of 'passive' elements, as well as the current of
the voltage source ( iVS ) and the voltage of the current source ( vIS ). Altogether, there are twelve complex unknowns. We
have three independent nodes, four independent loops ( marked as M I), and five passive elements which can be
characterized by five "Ohm's laws" - altogether there are 3+4+5 = 12 equations:

Nodal equations for N1 IVsM = IR1M + IC2M

for N2 IR1M = ILM + IC1M

for N3 IC2M + ILM + IC1M +IsM = IR2M

Loop equations for M1 VSM = VC2M + VR2M

for M2 VSM = VC1M + VR1M+ VR2M


for M3 VLM = VC1M

for M4 VR2M = VIsM

Ohm's laws VR1M = R1*IR1M

VR2M = R2*IR2M

IC1m = j**C1*VC1M

IC2m = j**C2*VC2M

VLM = j**L*ILM

Don't forget that any complex equation might lead to two real equations, so Kirchhoff's method requires many
calculations. It's much simpler to solve for the time functions of the voltages and currents using a system of differential
equations (not discussed here). First we show the results calculated by TINA's Interpreter:
{Solution by TINA's Interpreter}
f:=10000;
Vs:=10;
s:=0.005*exp(j*pi/6);
om:=2*pi*f;
sys ir1,ir2,ic1,ic2,iL,vr1,vr2,vc1,vc2,vL,vis,ivs
ivs=ir1+ic2 {1}
ir1=iL+ic1 {2}
ic2+iL+ic1+Is=ir2 {3}
Vs=vc2+vr2 {4}
Vs=vr1+vr2+vc1 {5}
vc1=vL {6}
vr2=vis {7}
vr1=ir1*R1 {8}
vr2=ir2*R2 {9}
ic1=j*om*C1*vc1 {10}
ic2=j*om*C2*vc2 {11}
vL=j*om*L*iL {12}
end;
abs(vr1)=[970.1563m]
abs(vr2)=[10.8726]
abs(ic1)=[245.6503u]
abs(ic2)=[3.0503m]
abs(vc1)=[39.0965m]
abs(vc2)=[970.9437m]
abs(iL)=[3.1112u]
abs(vL)=[39.0965m]
abs(ivs)=[3.0697m]
180+radtodeg(arc(ivs))=[58.2734]
abs(vis)=[10.8726]
radtodeg(arc(vis))=[-2.3393]
radtodeg(arc(vr1))=[155.1092]
radtodeg(arc(vr2))=[-2.3393]
radtodeg(arc(ic1))=[155.1092]
radtodeg(arc(ic2))=[-117.1985]
radtodeg(arc(vc2))=[152.8015]
radtodeg(arc(vc1))=[65.1092]
radtodeg(arc(iL))=[-24.8908]
radtodeg(arc(vL))=[65.1092]

Now try to simplify the equations by hand using substitution. First substitute eq.9. into eq 5.

VS = VC2 + R2 IR2 a.)

then eq.8 and eq.9. into eq 5.

VS = VC1 + R2 IR2 + R1 IR1 b.)

then eq 12., eq. 10. and IL from eq. 2 into eq.6.

VC1 = VL = jL IL = jL(IR1 - IC1) = jL IR1 - jL jC1 VC1

Express VC1
c.)

Express VC2 from eq.4. and eq.5. and substitute eq.8., eq.11. and V C1:

d.)

Substitute eq.2., 10., 11. and d.) into eq.3. and express I R2

IR2 = IC2 + IR1 + IS = jC2 VC2 + IR1 + IS

e.)

Now substitute d.) and e.) into eq.4 and express IR1

Numerically:
Click here to load or save this circuit

According to TINA's results.

The time function of iR1 is the following:

iR1(t) = 0.242 cos (t+155.5) mA

The measured voltages:


Click here to load or save this circuit

Click here to load or save this circuit

NODE POTENTIAL AND MESH CURRENT METHOD IN AC CIRCUITS


KIRCHHOFF'S LAWS IN AC CIRCUITS COUPLED INDUCTORS

In the previous chapter, we've seen that the use of Kirchhoff's laws for AC circuit analysis not only results in many
equations (as too with DC circuits), but also (due to the use of complex numbers) doubles the number of unknowns. To
reduce the number of equations and unknowns there are two other methods we can use: the node potential and the mesh
(loop) current methods. The only difference from DC circuits is that in the AC case, we have to work withcomplex
impedances (or admittances) for the passive elements and complex peak or effective (rms) values for the voltages and
currents.

In this chapter we will demonstrate these methods by two examples.

Let's first demonstrate the use of the node potentials method.

Example 1

Find the amplitude and phase angle of the current i(t) if R = 5 ohm; L = 2 mH; C1 = 10 F; C2 = 20 F; f = 1 kHz; vS(t) =
10 cos t V and iS(t) = cos t A

Click here to load or save this circuit


Here we have only one independent node, N 1 with an unknown potential:  = vR = vL = vC2 = vIS . The best method is the
node potential method.

The node equation:

Express M from the equation:

Now we can calculate IM (the complex amplitude of the current i(t)):

The time function of the current:

i(t) = 0.3038 cos (t + 86.3) A

Using TINA
Click here to load or save this circuit
{Solution by TINA's Interpreter}
om:=2000*pi;
V:=10;
Is:=1;
Sys fi
(fi-V)*j*om*C1+fi*j*om*C2+fi/j/om/L+fi/R1-Is=0
end;
I:=(V-fi)*j*om*C1;
abs(I)=[303.7892m]
radtodeg(arc(I))=[86.1709]

Now an example of the mesh current method

Click here to load or save this circuit


Example 2

Find the current of the voltage generator V = 10 V, f = 1 kHz, R = 4 kohm, R2 = 2 kohm, C = 250 nF, L = 0.5 H, I = 10
mA, vS(t) = V cos t, iS(t) = I sin t

Although we could again use the method of node potential with only one unknown, we will demonstrate the solution
with the mesh current method.

Let's first calculate the equivalent impedances of R2,L (Z1) and R,C (Z2) to simplify the work:

and

Click here to load or save this circuit

We have two independent meshes (loops).The first is: v S, Z1 and Z2 and the second: iS and Z2. The direction of the mesh
currents are: I1 clockwise, I2counterclockwise.

The two mesh equations are: VS = J1*(Z1 + Z2) + J2*Z2 J2 = Is

You must use complex values for all the impedances, voltages and currents.
The two sources are: VS = 10 V; IS = -j*0.01 A.

We calculate the voltage in volts and the impedance in kohm so we get the current in mA.

Hence:

j1(t) = 10.5 cos (t -7.1) mA

Solution by TINA:
{Solution by TINA's Interpreter}
Vs:=10;
Is:=-j*0.01;
om:=2000*pi;
Z1:=R2*j*om*L/(R2+j*om*L);
Z2:=R/(1+j*om*R*C);
Sys I
Vs=I*(Z1+Z2)+Is*Z2
end;
I=[10.406m-1.3003m*j]
abs(I)=[10.487m]
radtodeg(arc(I))=[-7.1224]

Finally, let's check the results using TINA.


Click here to load or save this circuit

COUPLED INDUCTORS

NODE POTENTIAL AND MESH CURRENT METHOD IN AC

PERIODIC
CIRCUITS WAVEFORMS

Two inductors or coils that are linked by electromagnetic induction are said to be coupled inductors. When an alternating
current flows through one coil, the coil sets up a magnetic field which is coupled to the second coil and induces a voltage
in that coil. The phenomenon of one inductor inducing a voltage in another inductor is known as mutual inductance.

Coupled coils can be used as a basic model for transformers, an important part of power distribution systems and
electronic circuits. Transformers are used for changing alternating voltages, currents, and impedances, and to isolate
one part of a circuit from another.
Three parameters are required to characterize a pair of coupled inductors: two self inductances, L1 and L2, and the mutual
inductance, L12 = M. The symbol for coupled inductors is:

Circuits which contain coupled inductors are more complicated than other circuits because we can only express the
voltage of the coils in terms of their currents. The following equations are valid for the circuit above with the dot locations
and reference directions shown:

Using impedances instead:

The mutual inductance terms can have a negative sign if the dots have different positions. The governing
rule is that the induced voltage on a coupled coil has the same direction relative to its dot as the inducing
current has to its own dot on the coupled counterpart.

The T - equivalent circuit


is very useful when solving circuits with coupled coils.

Writing the equations you can easily check the equivalence.

Let us illustrate this through some examples.

Example 1

Find the amplitude and initial phase angle of the current.

vs (t) = 1cos (t ) V =1kHz

Click here to load or save this circuit

The equations: VS = I1*j  L1 - I*j  M

0 = I*j  L2 - I1*j  M

Hence: I1 = I*L2/M; and


i(t) = 0.045473 cos (t - 90) A

Click here to load or save this circuit

{Solution by TINA's Interpreter}


om:=2*pi*1000;
Sys I1,I
1=I1*j*om*0.001-I*j*om*0.0005
0=I*j*om*0.002-I1*j*om*0.0005
end;

abs(I)=[45.4728m]
radtodeg(arc(I))=[-90]

Example 2

Find the equivalent impedance of the two-pole at 2 MHz!


Click here to load or save this circuit

First we show the solution obtained by solving the loop equations. We suppose that the impedance meter current is 1 A
so that the meter voltage equals the impedance. You can see the solution in TINA's Interpreter.

{Solution by TINA's Interpreter}


{Use loop equations}
L1:=0.0001;
L2:=0.00001;
M:=0.00002;
om:=2*pi*2000000;
Sys Vs,J1,J2,J3
J1*(R1+j*om*L1)+J2*j*om*M-Vs=0
J1+J3=1
J2*(R2+j*om*L2)+J1*om*j*M-J3*R2=0
J3*(R2+1/j/om/C)-J2*R2-Vs=0
end;
Z:=Vs;
Z=[1.2996k-1.1423k*j]

We could also solve this problem using the T-equivalent of the transformer in TINA:
Click here to load or save this circuit

If we wanted to calculate the equivalent impedance by hand, we'd need to use wye to delta conversion. While this is
feasible here, in general circuits can be very complicated, and it is more convenient to use the equations for coupled coils.

PERIODIC WAVEFORMS

COUPLED INDUCTORS BODE PLOTS

The Fourier theorem states that any periodic waveform can be synthesized by adding appropriately weighted sine and cosine terms
of various frequencies. The theorem is well-covered in other textbooks, so we will only summarize the results and show some
examples.

Let our periodic function be f (t) = f (t nT) where T is the time of one period and n is an integer number.

w0= 2/ T the fundamental angular frequency.

By the Fourier theorem, the periodic function can be written as the following sum:
where

An and Bn are the Fourier coefficients and the sum is the Fourier series.

Another form, probably a bit more practical:

where

A0 = C0 is the DC or average value, A1, B1 and C1 are the fundamental components, and the others are the harmonic terms.

While only a few terms may be required to approximate some waveforms, others will require many terms.

Generally, the more terms included, the better the approximation, but for waveforms containing steps, such as rectangular impulses,
the Gibbs phenomenon comes into play. As the number of terms increases, the overshoot becomes concentrated in an ever smaller
period of time.

An even function f(t) = f(-t) (axis symmetry) requires only cosine terms.

An odd function f(t) = - f(-t) (point symmetry) requires only sine terms.

A waveform with mirror or half-wave symmetry has only odd harmonics in its Fourier representation.

Here we will not deal with the Fourier series expansion, but will only use a given sum of sines and cosines as an excitation for a
circuit.
In the earlier chapters of this book, we dealt with sinusoidal excitation. If the circuit is linear, the superposition theorem is valid. For a
network with nonsinusoidal periodic excitation, superposition allows us to calculate the currents and voltages due to each Fourier
sinusoid term one at a time. When all are calculated, we finally summarize the harmonic components of the response.

It is a bit complicated to determine the different terms of the periodic voltages and currents and, in fact, it may yield an overload of
information. In practice, we would like to simply make measurements. We can measure the different harmonic terms using
a harmonic analyzer, spectrum analyzer, wave analyzer or Fourier analyzer. All these are complicated and probably yield more data
than needed. Sometimes it is sufficient to describe a periodic signal only by its average values. But there are several kinds of
average measurements.

AVERAGE VALUES

Simple average or DC term was seen in the Fourier representation as A0

This average can be measured with instruments such as the Deprez DC instruments.

Effective value or rms (root mean square) has the following definition:

This is the most important average value because the heat dissipated in resistors is proportional to the effective value.
Many digital and some analog voltmeters can measure the effective value of voltages and currents.

Absolute average
This average is no longer important; earlier instruments measured this form of average.

If we know the Fourier representation of a voltage or current waveform, we can also calculate the average values as follows:

Simple average or DC term was seen in the Fourier representation as A0 = C0

Effective value or rms (root mean square) is, after integrating the Fourier series of the voltage:

The klirr factor is a very important ratio of the average values:

It is the ratio of the effective value of the higher harmonic terms to the effective value of the fundamental harmonic:

There seems to be a contradiction here--we solve network in terms of harmonic components, but we measure average
quantities.

Let us illustrate the method by simple examples:


Example 1
Find the time function and the effective (rms) value of the voltage v C(t)

Click here to load or save this circuit

if R = 5 ohm, C = 10 F and v(t)=(100 + 200 cos(0t) + 30 cos(3 0t - 90°)) V, where the fundamental angular frequency is 0= 30
krad/s.

Try using the superposition theorem to solve the problem.

The first step is to find the transfer function as a function of the frequency. For simplicity, use the substitution: s = j 

Now substitute the component values and s = j k 0where k = 0; 1; 3 in this example and 0= 30 krad/s. In V, A, ohm, F and Mrad/s
units:
It is helpful to use a table to organise the steps of the numerical solution:

k
W(jk) =

We can summarise the steps of the superposition solution in another table. As we have already seen, to find
the complex peak value of a component, we should multiply the complex peak value of the component of the
excitation by the value of the complex transfer function:

k VSk W(jk) VCk


0 100 1 100
1 200 0.55 e-j56.3 110 e-j56.3
3 30 e-j90 0.217 e-j77.5 6.51 e-j167.5

And finally we can give the time function knowing the complex peak values of the components:

vC(t) = 100 + 110 cos(0t - 56.3) + 6.51 cos(30t - 167.5) V

The rms (effective) value of the voltage is:

As you can see, TINA's measuring instrument measures this rms value.
Example 2

Find the time function and the effective (rms) value of the current i(t)

Click here to load or save this circuit

if R = 5 ohm, C = 10 F and v(t)=(100 + 200 cos(0t) + 30 cos(30t - 90°)) V where the fundamental angular frequency is 0= 30
krad/s.

Try to solve the problem using the superposition theorem.

The steps of the solution are similar to Example 1, but the transfer function is different.

Now substitute the numerical values and s = j k 0,where k = 0; 1; 3 in this example.

In V, A, ohm, F and Mrad/s units:


It's helpful to use a table during the numerical solution:

k
W(jk) =

We can summarise the steps of the superposition in another table. As we have already seen, to find the
peak value of a component, we should multiply the complex peak value of that component of the excitation
by the value of the complex transfer function. Use the complex peak values of the components of the
excitation:

k VSk W(jk) Ik
0 100 0 0
1 200 0.162 ej33.7 32.4 ej33.7
3 30 e-j90 0.195 ej12.5 5.85 e-j77.5

And finally, knowing the complex peak values of the components we can state the time function:

i(t) = 32.4 cos (0t + 33.7) + 5.85 cos (30t - 77.5) [A]

The rms value of the current:


You can often do a sanity check for part of the solution. For example, a capacitor can have a DC voltage but not a DC
current.

Example 3

Obtain the time function of the voltage Vab if R1= 12 ohm, R2 = 14 ohm, L = 25 mH, and

Click here to load or save this circuit

C = 200 F. The generator voltage is v(t)=(50 + 80 cos(0t) + 30 cos(2 0t+60°)) V, where the fundamental frequency is f0 = 50 Hz.

The first step is to find the transfer function:

Substituting numerical values in V, A, ohm, mH, mF, kHz units:


Merging the two tables:

k VSk Vabk

0 50
50

1 80
79.3 e-j66.3°

2 30 ej60
29.7 e-j44.7°

Click here to load or save this circuit

Finally the time function:

vab(t) = 50 + 79.3 cos(1t - 66.3) + 29.7 cos(21t - 44.7) [V]


and the rms value:

S-ar putea să vă placă și