Sei sulla pagina 1di 24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Unit 6: Series AC Circuits


In this unit we'll pull together a lot of things from earlier units and use them to analyze series circuits containing AC voltage
sources. You'll need to remember what you've learned about AC fundamentals, capacitors, inductors, complex numbers, and
phasors. Once you take all of that into account, though, you'll find that to analyze a series AC circuit, you follow the same
steps that you follow to analyze a series DC circuit.
Also, the same rules that hold for series DC circuits (such as Ohm's law, Kirchhoff's Voltage Law, and the Voltage-Divider
Rule) also hold for series AC circuits. But the math is a little more complicated, because each step involves phasors (complex
numbers) instead of ordinary numbers.
First, you should read the following sections of Thomas Floyd's Principles of Electric Circuits (8th edition):
Sinusoidal Response of Series RC Circuits (Section 15-2)
Impedance of Series RC Circuits (Section 15-3)
Analysis of Series RC Circuits (Section 15-4)
Sinusoidal Response of Series RL Circuits (Section 16-1)
Impedance of Series RL Circuits (Section 16-2)
Analysis of Series RL Circuits (Section 16-3)
Impedance of Series RLC Circuits (Section 17-1)
Analysis of Series RLC Circuits (Section 17-2)
Then work through the e-Lesson and Self-Test questions below.
After you finish the e-Lesson, you'll be ready to take Quiz #6, perform Lab #6, and do Homework #6.

Units 4 and 5 Review

This unit will build on material that you studied in Unit 4 and Unit 5 . So let's begin by taking these two self-tests to
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

1/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

review what you learned in those units.

Review of Series DC Circuits

In EET 150 you learned how to analyze series DC circuits like the one shown below. Let's do a quick review of what
you learned there.

You should recall that the basic steps in analyzing a circuit like this one are:
1. Add the resistance values to find the circuit's total resistance, RT. For the circuit shown, this means that
RT = R1 + R2 + R3
2. Apply Ohm's law to the entire circuit to find the circuit's total current:
IT = VS RT
3. Recognize that, since we're dealing with a series circuit, each resistor's current is equal to the total current. For the
circuit shown:
I1 = I2 = I3 = IT
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

2/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

4. Apply Ohm's law to each resistor to find the voltage drops:


V1 = I1 R1

and

V2 = I2 R2

and

V3 = I3 R3

Review: Kirchhoff's Voltage Law in Series DC Circuits

In EET 150 you also learned that Kirchhoff's Voltage Law (KVL) says that the sum of the voltage drops around any
closed loop equals the sum of the voltage rises around that loop.
In terms of a simple series DC circuit like the one you just analyzed, this means that the sum of all the resistor voltage
drops must equal the source voltage.

Review: Voltage Divider Rule in Series DC Circuits

In EET 150 you also learned that the voltage-divider rule is a shortcut rule that you can use to find the voltage drop
across a resistor in a series circuit.
The rule says that the voltage across any resistance in a series circuit is equal to the ratio of that resistance to the
circuit's total resistance, multiplied by the source voltage.
In equation form, this rule is expressed as:
Vx = VS(RxRT)

Review: Troubleshooting Series DC Circuits


http://people.sinclair.edu/nickreeder/eet155/mod06.htm

3/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Troubleshooting a non-working circuit means finding the problem that is preventing the circuit from working
correctly.
The two most common types of problems are open circuits and short circuits.
An open circuit, or "open," is a break in a circuit path.
The most important thing to remember about opens is that no current can flow through an open.
Therefore, no current can flow anywhere in a series circuit containing an open.
Since no current flows through an open, you can think of the open as having infinite resistance (R = ).
Usually, an open will not have a voltage drop of 0 V. In fact, in a series DC circuit that contains an open, the
entire source voltage will appear across the open, and no voltage will appear across any of the other
resistors.
So if you measure the voltage between any two points in a series circuit containing an open, you'll measure 0 V if
the two points are on the same side of the open, but you'll measure the entire source voltage if the points are on
opposite sides of the open.
For example, suppose R3 is open in the circuit shown below. Then there will be 0 V across R1, across R2,
and across R4. Also, Vab = 0 V. But there will be 9 V across R3. Also, Vac = 9 V, and Vbc = 9 V.

A short circuit, or "short," is a path of zero resistance connecting two points in a circuit that are not supposed to be
connected.
Since a short has zero resistance, the voltage across it must be zero. This follows from Ohm's law, V = I R.
A component is said to be short-circuited, or "shorted out," when there is a short circuit connected in parallel with
it. No current flows through a short-circuited component. Instead, current is diverted through the short itself.
For example, suppose that in the circuit shown below there is a short between points a and b, perhaps
caused by a loose wire clipping that connects these two points. Then R2 is short-circuited. No current will
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

4/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

flow through R2; instead, current will follow the path of zero resistance through the short itself (the wire
clipping).

A short in a series DC circuit reduces the circuit's total resistance, causing more current to flow out of the voltage
source.
For example, in the circuit shown above, if R2 is short-circuited by a wire clipping that connects points a
and b, then R2's resistance disappears from the circuit, and the circuit's total resistance is equal to
R1 + R3 + R4.

That ends our quick review of series DC circuits. If you'd like a more thorough review, go to Unit 2 of EET 150. Now
let's get back to AC circuits.

Sinusoidal Response of AC Circuits (Floyd, p. 610)

Here's an important point that we've mentioned a couple of times before and that is worth repeating: When a sinusoidal
voltage is applied to any circuit containing resistors, capacitors, and inductors, all of the circuit's current
waveforms and voltage waveforms are sinusoids and have the same frequency as the source voltage.
So, for example, if you're given the circuit shown below, and if you're told that the source voltage is a sinusoid having a
frequency of 5 kHz, then you can say immediately that the current through every component is a 5-kHz sinusoidal
current, and the voltage drop across every component is a 5-kHz sinusoidal voltage.

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

5/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Where things get a bit tricky is figuring out the peak values and phase shifts of these current and voltage waveforms.
But we'll be able to do it, thanks to complex numbers.
Impedance (Floyd, p. 611)

Recall that resistance (R), capacitive reactance (XC), and inductive reactance (XL) are all measured in ohms.
Each one of these quantities represents a component's opposition to the flow of AC current.
Impedance is the general term that applies to all three of these quantities, or to a combination of them.
For example, if you combine a resistor and a capacitor in series, then their combined opposition to the flow of AC
current is called their total impedance. This total impedance is a combination of the resistor's resistance and the
capacitor's reactance.
But you can't find total impedance simply by adding resistances and reactances together. For example, if you've
got a 1-k resistor in series with a capacitor whose reactance is 2 k, their total impedance is not 3 k. As we'll
see soon, you have to use complex numbers to find their total impedance.
The symbol for impedance is Z. It is measured in ohms.
Boldface Notation for Phasors (Floyd, p. 611)

Up to now we have used italicized, non-boldface letters to represent various quantities. In particular:
R represents resistance.
XC represents capacitive reactance.
XL represents inductive reactance.
Z represents impedance.
V represents voltage.
I represents current.
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

6/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

From this point onward, we will usually treat all of these quantities as phasors, which means we'll treat them as
complex numbers with both a magnitude and an angle. We'll use the italic letters listed above to denote the magnitude
of the phasor, and we'll use boldface letters to represent the total phasor quantity (which includes both the magnitude
and the angle). In particular:
R represents a phasor resistance, which has both a magnitude (R) and an angle.
XC represents a phasor capacitive reactance, which has both a magnitude (XC) and an angle.
XL represents a phasor inductive reactance, which has both a magnitude (XL) and an angle.
Z represents a phasor impedance, which has both a magnitude (Z) and an angle.
V represents a phasor voltage, which has both a magnitude (V) and an angle.
I represents a phasor current, which has both a magnitude (I) and an angle.
For example, suppose the total impedance in a circuit has a magnitude of 7.36 k and an angle of 26.5. At times we
might be interested in talking just about the magnitude, in which case we'll write
Z = 7.36 k
At other times we might be interested in talking about the total phasor quantity (magnitude and angle), in which case
we'll write
Z = 7.36 26.5 k
Phasor Form of Resistance

We treat resistance as a phasor R whose magnitude R is the resistance in ohms and whose angle is 0.
We use 0 because voltage and current are in phase in resistors. (In other words, there is a 0 phase angle between
a resistor's current and its voltage.)
So in polar notation, the phasor form of resistance is
R = R0
We can easily convert this to rectangular notation, to get
R = R + j0
or simply
R=R
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

7/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Thus, the phasor R for any resistor has a real part but no imaginary part. In the complex plane, resistance lies along the
positive real axis, as shown in the following diagram of a phasor representing a resistance of 50 :

Phasor Form of Capacitive Reactance (Floyd, p. 611)

We treat capacitive reactance as a phasor XC whose magnitude XC is 1 (2pfC), and whose angle is 90.
We use 90 because voltage lags current by 90 in a capacitor.
So in polar notation, the phasor form of capacitive reactance is
XC = XC 90
We can easily convert this to rectangular notation, to get
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

8/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

XC = 0 jXC
or simply
XC = jXC
Remember, in each of these equations XC = 1 (2pfC), which is also equal to 1 (C).
Thus, the phasor XC for any capacitive reactance has a negative imaginary part but no real part. In the complex plane,
capacitive reactance lies along the negative imaginary axis, as shown in the following diagram of a phasor representing
a capacitive reactance of 50 :

Phasor Form of Inductive Reactance (Floyd, p. 679)


http://people.sinclair.edu/nickreeder/eet155/mod06.htm

9/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

We treat inductive reactance as a phasor XL whose magnitude XL is 2pfL and whose angle is +90.
We use +90 because voltage leads current by 90 in an inductor.
So in polar notation, the phasor form of inductive reactance is
XL = XL 90
We can easily convert this to rectangular notation, to get
XL = 0 + jXL
or simply
XL = jXL
Remember, in each of these equations XL = 2pfL, which is also equal to L.
Thus, the phasor XL for any inductive reactance has a positive imaginary part but no real part. In the complex plane,
inductive reactance lies along the positive imaginary axis, as shown in the following diagram of a phasor representing
an inductive reactance of 50 :

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

10/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Phasor Form of Ohm's Law (Floyd, p. 614)

Once you know a component's resistance or reactance, you can use Ohm's law to find the component's current if you
knows its voltage, or to find the component's voltage if you knows its current.
To use Ohm's law in AC circuits, we'll treat voltages, currents, resistances, and reactances as phasors.
For resistors, V = I R.
For capacitors, V = I XC.
For inductors, V = I XL.
Recall that impedance is the general term applied to resistance, capacitive reactance, inductive reactance, or any
combination of these. The symbol for impedance is Z, so we can write the most general form of Ohm's law as:
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

11/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

V=IZ
Of course, you can also rearrange this equation (and the ones given above) to solve for current if you know voltage and
impedance, or to solve for impedance if you know current and voltage:
I=VZ
Z=VI
Remember, in each case, all quantities are phasors, not ordinary numbers.

Series Impedance (Floyd, p. 612)

Suppose we have a resistance in series with a reactance. We'd like to find the total impedance of these two components,
but we can't simply add resistances and reactances as ordinary numbers. For example, a 1 k resistance in series with
a 2 k capacitive reactance does not add up to 3 k.
Instead, we must add them as phasors. So, to find the total impedance of a 1 k resistance in series with a 2 k
capacitive reactance, we must add 10 k plus 290 k, which gives us a total of 2.2463.4 k.
It may seem strange that you can combine a 1 k resistance with a 2 k reactance and come up with a total of only
2.24 k, but that's how it works.
Analyzing Series AC Circuits

Now that you know how to treat resistances and reactances as phasors, and how to use the phasor form of Ohm's law,
and how to use phasors to find total impedance, you're ready to analyze any series AC circuit, such as the series RLC
circuit shown below.

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

12/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Here are the steps to follow:


1. Use phasor addition to find the circuit's total impedance, ZT.
2. Apply Ohm's law (phasor form) to the entire circuit to find the circuit's total current.
3. Recognize that, since we're dealing with a series circuit, each component's current is equal to the total current.
4. Apply Ohm's law (phasor form) to each component to find the voltage drops.
These are very similar to the steps that you followed in EET 150 to analyze a simple series DC circuit containing
resistors. The big difference is that throughout this procedure, we must now use phasors (complex numbers) instead
of ordinary numbers.
Let's look at each step in more detail.
Step 1: Find Total Impedance

For n impedances in series, total impedance is given by


ZT = Z1 + Z2 + + Zn
Each Z on the right-hand side of this equation may be a resistance, an inductive reactance, or a capacitive reactance. So
this step will require you first to find the reactances of any capacitors or inductors in the circuit.
Remember: we're adding phasors here, not ordinary numbers.

Step 2: Find Total Current


http://people.sinclair.edu/nickreeder/eet155/mod06.htm

13/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Knowing the source voltage VS and the total impedance ZT, you can use Ohm's Law to find the current:
IT = VS ZT
Again, remember that we're dividing phasors, not ordinary numbers.

Step 3: Find Individual Currents

This step is the easiest. It simply requires you to remember that in any series circuit (DC or AC), every component's
current is equal to the total current:
IT = I1 = I2 = = In

Step 4: Find Voltage Drops

Now that you know the current through each component, use Ohm's Law to find the voltage drop across each
component:
V1 = I 1 Z 1

and

V2 = I 2 Z 2

and

V3 = I 3 Z 3

and ...

Remember, Z1 is the first component's impedance, which will be a resistance if the first component is a resistor, or a
capacitive reactance if the first component is a capacitor, or an inductive reactance if the first component is an inductor.
Similarly for Z2, Z3, and so on for however many components the circuit contains.
Again, remember that we're multiplying phasors, not ordinary numbers.

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

14/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Phasor Diagrams

After you've analyzed a circuit by finding currents and voltage drops, you can draw a phasor diagram that shows in
graphical form how these quantities relate to each other. A phasor diagram simply shows each voltage and current as a
vector in the complex plane, drawn with the appropriate angle and magnitude.
Here is a simple example showing the current and voltage phasors for a single component:

Here's another example showing the phasors for all voltages and currents in a particular series RC circuit.

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

15/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

From the diagram you can quickly see the relationship between the circuit's current and voltages.
In every phasor diagram for a series circuit, you should find that the current and the resistor's voltage drop have the
same angle, since current and voltage in a resistor are always in phase with each other.
Also, you should find that inductor voltage and capacitor voltage are always at a 90 angle to the current, which
should make sense. (Remember ELI the ICEman from Unit 4?)

Kirchhoff's Voltage Law

As in DC circuits, Kirchhoff's Voltage Law (KVL) says that the sum of the voltage drops around any closed loop
equals the sum of the voltage rises around that loop.
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

16/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

As we'll see in later units of this course, KVL applies to all circuits, whether series, parallel, or series-parallel.
In this unit we're restricting our attention to series circuits containing a single voltage source. In these circuits,
KVL simplifies to the following form: the source voltage in a series circuit is equal to the sum of the voltage
drops across the circuit's resistors, capacitors, and inductors.
Whenever you apply KVL to an AC circuit, you must use phasors, not ordinary numbers. If you just add the
magnitudes of the voltages, instead of adding the magnitudes along with their angles, you won't get good results.

Voltage-Divider Rule

As in DC circuits, this is a shortcut rule for finding voltage drops in a series circuit.
The voltage-divider rule says that the voltage Vx across any impedance Zx in a series circuit with source voltage VS
is given by:
Vx = (Zx ZT) VS
Again, use phasors, not ordinary numbers.

Troubleshooting Series AC Circuits

Above we reviewed the basics of troubleshooting series DC circuits. Almost all of these same points apply to series AC
circuits. (But there's one important difference noted below.) In particular:
No current flows through an open.
Therefore, no current flows anywhere in a series circuit containing an open.
Since no current flows through an open, you can think of the open as having infinite resistance (R = ).
Usually, an open will not have a voltage drop of 0 V. In fact, in a series circuit that contains an open, the entire
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

17/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

source voltage will appear across the open, and no voltage will appear across any of the other resistors,
capacitors, or inductors.
So if you measure the voltage between any two points in a series circuit containing an open, you'll measure 0 V if
the two points are on the same side of the open, but you'll measure the entire source voltage if the points are on
opposite sides of the open.
A short has zero resistance and zero voltage.
A component is said to be short-circuited, or "shorted out," when there is a short connected in parallel with it. No
current flows through a short-circuited component. Instead, current is diverted through the short itself.
So far, everything we've said about opens and shorts in series AC circuits is the same as what we said earlier about
opens and shorts in series DC circuits. But here's a difference:
A short in a series DC circuit will always reduce the circuit's total resistance, increasing the circuit's total current.
But in a series AC circuit, a short could either increase or decrease the circuit's total impedance, and therefore
could either decrease or increase the total current.
Why is there this difference between shorts in DC and AC circuits? It's because of the difference between regular
numbers and complex numbers.
To find a series DC circuit's total resistance, you add two or more regular numbers. If one of the circuit's resistors
is shorted out, then you replace one of these numbers with zero, and this will decrease the total.
For example, suppose a series circuit contains a 100 resistor, a 150 resistor, and a 200 resistor. Then
the total resistance is 450 . But if the 200 resistor is shorted out, then you replace its resistance with
zero, and so the total resistance decreases to 250 . This decrease in total resistance will increase the
circuit's total current.
But to find a series AC circuit's total impedance, you add two or more complex numbers (phasors). If one of the
circuit's components is shorted out, then you replace one of these complex numbers with zero. When you're
adding complex numbers, replacing one of them by zero could either decrease or increase the total.
For example, suppose a series AC circuit contains a 1000 resistance, a 15090 inductive
reactance, and a 20090 capacitive reactance. Then the total impedance is 11226.6 . If the
capacitor is shorted out, then you replace its reactance with zero, and so the total impedance increases to
18056.3 . This increase in total impedance will decrease the circuit's total current.
On the other hand, suppose that in the same circuit the capacitor is okay and the resistor is shorted out.
Then the circuit's total impedance is 5090 , which is a decrease from the original total impedance.
This decrease in total impedance will increase the circuit's total current.
So a short in a series AC circuit may either increase or decrease the total impedance.

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

18/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

"Systems View" of a Circuit

In your electronics courses up to now, you've considered a circuit as a collection of individual components such as
resistors, capacitors, and inductors. Most of the circuits you've studied have been very small, with just a few
components.
But real-world circuits are usually much more complex, with hundreds or even thousands of components. To study and
analyze such circuits, you have to shift your way of looking at circuits. Instead of concentrating on individual
components, it's more useful to think of the circuit as a system made up of parts that perform certain functions. The
"parts" that I'm referring to here are not individual components. Rather, they are sub-circuits that contain many
components connected together to perform some function.
For example, in your later courses you'll study amplifier circuits. These circuits contain transistors as well as resistors,
capacitors, and inductors, so we're not ready to understand the details now. But there are a number of standard designs
for amplifier circuits, and rather than focusing on the details you might just want to consider the entire amplifier circuit
as a "box." This box has two input terminals to which you can connect an input voltage, and two output terminals at
which the amplifier's output voltage will appear. Here's a diagram:

The amplifier circuit contains many components (resistors, capacitors, transistors). But in this diagram we're not
showing those details.
A more complete circuit might consist of an oscillator connected to an amplifier connected to a power amplifier, as
shown here:

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

19/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

In this diagram we have three boxes representing three complicated sub-circuits. We're not showing the details of these
sub-circuits, but we are showing how the sub-circuits are connected to each other. (For example, the two lines between
the oscillator and the amplifier show that the oscillator's output voltage is also the amplifier's input voltage.)
Next we'll look at a few simple series RC and RL circuits that can be thought of as "boxes" that perform a certain
function.
Lag Circuits and Lead Circuits

In some applications, a designer needs to shift a voltage's phase angle by a certain amount. In such cases the designer
uses a circuit that introduces a phase shift between the circuit's output voltage and its input voltage.
There are two basic possibilities here. Either:
The circuit's output voltage lags its input voltage, in which case we're dealing with a lag circuit. Or:
The circuit's output voltage leads its input voltage, in which case we're dealing with a lead circuit.

RC Lag Circuits and Lead Circuits (Floyd, pp. 620-623)

A simple series RC circuit can serve as either a lag circuit or a lead circuit, depending on whether you take the output
voltage across the resistor or across the capacitor. In particular:
In any series RC circuit, the capacitor's voltage lags the source voltage, so you'll have a lag circuit if you take the
output voltage across the capacitor, as shown here:

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

20/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Also, in any series RC circuit, the resistor's voltage leads the source voltage, so you'll have a lead circuit if you
take the output voltage across the resistor, as shown here:

If you remember ELI the ICEman, you'll be able to quickly identify circuits like the ones above as either lead circuits
or lag circuits.
ICE reminds you that a capacitor's voltage tends to lag everything else in the circuit, so you've got a lag circuit if
you're taking the output voltage across the capacitor.
Whenever you encounter circuits like these, you can always use the general techniques you learned above to analyze the
circuit and figure out how far the output voltage is shifted from the input voltage. Or you can remember the following
formulas.
For an RC lag circuit, the phase angle between the input and output is
= tan1(R XC)
For an RC lead circuit, the phase angle between the input and output is
= tan1(XC R)
Using these formulas, and by choosing appropriate values of R and C, you can design a lead circuit or lag circuit to shift
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

21/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

the voltage by any desired angle.


Here's a nice memory trick to help you remember those two formulas: Notice that the order of the R and the XC in each
formula is the same as the order of the resistor and the capacitor in the corresponding schematic diagram.

RL Lag Circuits and Lead Circuits (Floyd, pp. 685-689)

A simple series RL circuit can also serve as either a lag circuit or a lead circuit, depending on whether you take the
output voltage across the resistor or across the inductor. In particular:
In any series RL circuit, the resistor's voltage lags the source voltage, so you'll have a lag circuit if you take the
output voltage across the resistor, as shown here:

Also, in any series RL circuit, the inductor's voltage leads the source voltage, so you'll have a lead circuit if you
take the output voltage across the inductor, as shown here:

If you remember ELI the ICEman, you'll be able to quickly identify circuits like the ones above as either lead circuits
or lag circuits.
http://people.sinclair.edu/nickreeder/eet155/mod06.htm

22/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

ELI reminds you that an inductor's voltage tends to lead everything else in the circuit, so you've got a lead circuit
if you're taking the output voltage across the inductor.
Whenever you encounter circuits like these, you can always use the general techniques you learned above to analyze the
circuit and figure out how far the output voltage is shifted from the input voltage. Or you can remember the following
formulas.
For an RL lag circuit, the phase angle between the input and output is
= tan1(XL R)
For an RL lead circuit, the phase angle between the input and output is
= tan1(R XL)
Here's a nice memory trick to help you remember those two formulas: Notice that the order of the R and the XL in each
formula is the same as the order of the resistor and the inductor in the corresponding schematic diagram.

Practice Problems

Want more practice analyzing series AC circuits? Here are a couple of learning objects that will generate as many
practice problems as you want, and then let you check your answers against the correct answers.
The first one covers series RC circuits:

And the second one covers series RL circuits:

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

23/24

9/5/2014

EET 155 Unit 6: Series AC Circuits

Unit 6 Review

This e-Lesson has covered several important topics, including:


series AC circuits
phasor diagrams
Kirchhoff's Voltage Law
voltage-divider rule
troubleshooting series AC circuits
lag circuits and lead circuits.
To finish the e-Lesson, take this self-test to check your understanding of these topics.

Congratulations! You've completed the e-Lesson for this unit. What's next?
Take Online Quiz #6.
Perform Lab #6 and turn in a typed short lab report. (You may wish to review my instructions on writing short
reports.)
Do Homework #6.
Keep practicing your skills by playing the games on the Games page.
Prepare for Unit 7 by reading Sections 15-5, 15-6, 16-4, 16-5, 17-4, and 17-5 of Thomas Floyd's Principles of Electric
Circuits (8th edition).
Then you'll be ready to go on to Unit 7 .
Nick Reeder | Electronics Engineering Technology | Sinclair Community College
Send comments to nick.reeder@sinclair.edu

http://people.sinclair.edu/nickreeder/eet155/mod06.htm

24/24

Potrebbero piacerti anche