You are on page 1of 34

Alternating Current Circuits

Alternating currents and voltages are sinusoidal and vary with time. Alternating currents produce different responses in resistors, capacitors, and inductors than direct currents do.

Alternating currents and voltages


Figure 1 shows the plot of alternating voltage and alternating current as a function of time in a circuit that has only a resistor and a source of alternating current an ac generator.

Figure 1 Current and voltage from an ac source through a simple resistor.


Because the voltage and current reach their maximum values at the same time, they are in phase. Ohm's law and the previous expressions for power are valid for this circuit if the root mean square (rms) of the voltage and the rms of the current, sometimes called the effective value, are used. These relationships are

Ohm's law is expressed thus: VR = IR, where VR is the rms voltage across the resistor and I is the rms in the circuit.

Resistor-capacitor circuits
A circuit with a resistor, a capacitor, and an ac generator is called an RC circuit. A capacitor is basically a set of conducting plates separated by an insulator; thus, a steady current cannot pass through the capacitor. A time-varying current can add or remove charges from the capacitor plates. A simple circuit for charging a capacitor is shown in Figure 2 .

Figure 2 An RC circuit for charging a capacitor.


Initially, at time t = 0, the switch (S) is open, and there is no charge on the capacitor. When the switch is closed, a current will pass through the resistor and charge the capacitor. The current will cease when voltage drop across the capacitor equals the potential of the battery (V). Once the capacitor reaches the maximum charge, the current will decrease to zero. The current is at maximum immediately after the switch is closed and decreases exponentially with time. Thecapacitive time constant (), the Greek letter tau) is the time for the charge to decay to 1/ e of its initial value, where e is the natural logarithm. A capacitor with a large time constant will change slowly. The capacitive time constant is = RC. From Kirchhoff's rules, the following expressions for the potential difference across the capacitor (VC) and the current (I) in the circuit are derived:

where V is the potential of the battery.

Resistor-inductance circuits
A circuit with a resistor, an inductor, and an ac generator is an RL circuit. When the switch is closed in an RL circuit, a back emf is induced in the inductor coil. The current, therefore, takes time to reach its maximum value, and the time constant, called the inductive time constant, is given by

The equations for the current as a function of time and for the potential across the inductor are

A switch was used in the above discussions of RC and RL circuits for simplicity. Opening and closing a switch gives a response similar to that of an ac current. The RC and RL circuits are similar to each other because an increase in voltage yields a current that changes exponentially in each circuit, but the responses are different in other ways. These different behaviors, described below, lead to different responses in ac circuits.

Reactance
Now consider an ac circuit consisting only of a capacitor and an ac generator. The plots of current and voltage across the capacitor as a function of time are shown in Figure 3 . The curves are not in phase as they were for the circuit of a resistor and an ac generator. (Refer to Figure 1 .) The curves indicate that, for a capacitor, the voltage reaches its maximum value one quarter of a cycle after the current reaches its maximum value. Thus, the voltage lags the current through the capacitor by 90 degrees.

Figure 3 Current and voltage from an ac source through a capacitor.


The capacitive reactance (Xc) expresses the impeding effect of the capacitor on the current and is defined as

where C is in farads and the frequency (f) is in units of hertz. Ohm's law yields V c= IX c , where V c is the rms voltage across the capacitor and I is the rms current in the circuit. Consider a circuit with only an inductor and an ac generator. Figure 4 shows the plots of the current and voltage as a function of time for the inductor. Note again that the voltage and current are not in phase. The voltage for this circuit reaches its maximum value one quarter of a cycle before the current reaches its maximum; thus, the voltage leads the current by 90 degrees.

Figure 4 Current and voltage from an ac source through an inductor.


The current in the circuit is impeded by the back emf of the inductor coil. The effective resistance is called the inductive reactance (XL) defined by (XL) =2 fL, where L is measured in henries and f is in hertz. Ohm's law yields (VL) = IXL , where (VL) is the rms voltage across the inductor and I is the rms in the inductor.

Resistor-inductor-capacitor circuit
A circuit with a resistor, an inductor, a capacitor, and an ac generator is called anRLC circuit. The phase relationships of these elements can be summarized as follows: The instantaneous voltage across the resistor VR is in phase with the instantaneous current. The instantaneous voltage across the inductor VL leads the instantaneous current by 90 degrees. The instantaneous voltage across the capacitor Vc lags the instantaneous current. Because the voltages across the different elements are not in phase, the individual voltages cannot be simply added in ac circuits. The equations for the total voltage and the phase angle are

where all voltages are rms values. Ohm's law for the general case of ac circuits is now expressed V = IZ, where R is replaced by impedance ( Z), measured in ohms. The impedance is defined as

http://www.cliffsnotes.com/study_guide/Alternating-Current-Circuits.topicArticleId10453,articleId-10437.html

Alternating Current (AC) Electricity


by Ron Kurtus (revised 2 June 2009) Alternating current (AC) electricity is the type of electricity commonly used in homes and businesses throughout the world. While direct current (DC) electricity flows in one direction through a wire, AC electricity alternates its direction in a back-andforth motion. The direction alternates between 50 and 60 times per second, depending on the electrical system of the country. AC electricity is created by an AC electric generator, which determines the frequency. What is special about AC electricity is that the voltage can be readily changed, thus making it more suitable for long-distance transmission than DC electricity. But also, AC can employ capacitors and inductors in electronic circuitry, allowing for a wide range of applications.
Note: We usually say AC electricity instead of simply saying AC, since that is also the abbreviation for air conditioning. You need to be exact in science to avoid any misunderstandings.

Questions you may have include: What is the difference between AC and DC electricity? Why do we use AC instead of DC? What are advantages of AC electricity?

This lesson will answer those questions.


Useful tool: Metric-English Conversion

Difference between AC and DC electricity


Electrons have negative () electrical charges. Since opposite charges attract, they will move toward an area consisting of positive (+) charges. This movement is made easier in an electrical conductor, such as a metal wire.

Electrons move direct with DC electricity


With DC electricity, connecting a wire from the negative () terminal of a battery to the positive (+) terminal will cause the negative charged electrons to rush through the wire toward the positive charged side. The same thing happens with a DC generator, where the motion of coiled wire through a magnetic field pushes electrons out of one terminal and attracts electrons to the other terminal.

Electrons alternate directions in AC electricity


With an AC generator, a slightly different configuration alternates the push and pull of each generator terminal. Thus the electricity in the wire moves in one direction for

a short while and then reverses its direction when the generator armature is in a different position. This illustration gives an idea of how the electrons move through a wire in AC electricity. Of course, both ends of the wire extend to the AC generator or source of power.
AC movement of electrons in wire

The charge at the ends of the wire alternates between negative () and positive (+). If the charge is negative (), that pushes the negatively charged electrons away from that terminal. If the charge is positive (+), the electrons are attracted in that direction. Rate of change AC electricity alternates back-and-forth in direction 50 or 60 times per second, according to the electrical system in the country. This is called the frequency and is designated as either 50 Hertz (50Hz) or 60 Hertz (60Hz). (See Worldwide AC Voltages and Frequencies for more information.)

Light bulbs
Many electrical deviceslike light bulbsonly require that the electrons move. They don't care if the electrons flow through the wire or simply move back-and-forth. Thus a light bulb can be used with either AC or DC electricity.

AC is periodic motion
The regular back-and-forth motion of the electrons in a wire when powered by AC electricity is periodic motion, similar to that of a pendulum. (See Periodic Motion and Pendulum for more information.) Because of this periodic motion of the electrons, the voltage and current follow a sine waveform, alternating between positive (+) and negative (), as measured with a voltmeter or multimeter.

Waveform varies between positive and negative as it travels in time

The rate that the voltage or current peaks pass a given point is the frequency of the AC electricity.

Advantages of AC electricity
There are distinct advantages of AC over DC electricity. The ability to readily transform voltages is the main reason we use AC instead of DC in our homes.

Transforming voltages
The major advantage that AC electricity has over DC electricity is that AC voltages can be readily transformed to higher or lower voltage levels, while it is difficult to do that with DC voltages. Since high voltages are more effecient for sending electricity great distances, AC electricity has an advantage over DC. This is because the high voltages from the power station can be easily reduced to a safer voltage for use in the house. Changing voltages is done by the use of a transformer. This device uses properties of AC electromagnets to change the voltages. (See AC Transformers for more information.)

Tuning circuits
AC electricity also allows for the use of a capacitor and inductor within an electrical or electronic circuit. These devices can affect the way the alternating current passes through a circuit. They are only effective with AC electricity. A combination of a capacitor, inductor and resistor is used as a tuner in radios and televisions. Without those devices, tuning to different stations would be very difficult.

Summary
We commonly use AC electricity to power our television, lights and computers. In AC electricity, the current alternates in direction. AC electricity was proven to be better for supplying electricity than DC, primarily because the voltages can be transformed. AC also allows for other devices to be used, opening a wide range of applications.

http://www.school-for-champions.com/science/ac.htm

AC circuits: alternating current electricity


Alternating current (AC) circuits explained using time and phasor animations. Impedance, phase relations, resonance and RMS quantities. A resource page from Physclips: a multi-level, multimedia introduction to physics. AC electricity is ubiquitous not only in the supply of power, but in electronics and signal processing. Here are some of the questions we'll answer:

The voltage across a resistor (upper trace) and an inductor (lower trace) in series. Note the phase difference.

If impedance is voltage over current, why is it different from resistance? What are RMS values? How can you have 100 volts across a capacitor, 100 volts across an inductor, and 10 volts across the two of them in series? What does it mean to say that current and voltage are out of phase? Why does the ratio of voltage to current in capacitors and inductors depend on frequency?

We introduce the voltage-current relations for resistors, capacitors and inductors separately using animations to show the time-varying nature, and why frequency is important. Then we combine the components in series and parallel.
Sub-headings on this page are:

There are separate pages on


Resistors in AC circuits What are impedance and reactance? Capacitors and charging

Ohm's law and drift velocity in conductors RC filters, integrators and differentiators LC oscillations

Inductors and the Farady emf Impedance of components RC Series combinations RL Series combinations RLC Series combinations Resonance Bandwidth and Q factor Complex Impedance (more to come)

Power, RMS values and three-phase circuits Transformers, and Motors and generators

But first, why study AC circuits? You probably live in a house or appartment with sockets that deliver AC. Your radio, television and portable phone receive it, using (among others) circuits like those below. As for the computer you're using to read this, its signals are not ordinary sinusoidal AC, but, thanks to Fourier's theorem, any varying signal may be analysed in terms of its sinusoidal components. So AC signals are almost everywhere. And you can't escape them, because even the electrical circuits in your brain use capacitors and resistors.

Some terminology

For brevity, we shall refer to electrical potential difference as voltage. Throughout this page, we shall consider voltages and currents that vary sinusoidally with time. We shall use lower case v and i for the voltage and current when we are considering their variation with time explicitly. The amplitude or peak value of the sinusoidal variation we shall represent by Vm and Im, and we shall use V = Vm/2 and I = Im/2 without subscripts to refer to the RMS values. For an explanation of RMS values, see Power and RMS values. For the origin of the sinusoidally varying voltage in the mains supply, see Motors and generators. So for instance, we shall write: v = v(t) = Vm sin (t + ) i = i(t) = Im sin (t).

where is the angular frequency. = 2f, where f is the ordinary or cyclic frequency. f is the number of complete oscillations per second. is the phase difference between the voltage and current. We shall meet this and the geometrical significance of later.

Resistors and Ohm's law in AC circuits

The voltage v across a resistor is proportional to the current i travelling through it. (See the page on drift velocity and Ohm's law.) Further, this is true at all times: v = Ri. So, if the current in a resistor is i = Im . sin (t) , we write: v = R.i = R.Im sin (t) v = Vm. sin (t) where Vm = R.Im So for a resistor, the peak value of voltage is R times the peak value of current. Further, they are in phase: when the current is a maximum, the voltage is also a maximum. (Mathematically, = 0.) The first animation shows the voltage and current in a resistor as a function of time. The rotating lines in the right hand part of the animation are a very simple case of a phasor diagram (named, I suppose, because it is a vector representation of phase). With respect to the x and y axes, radial vectors or phasors representing the current and the voltage across the resistance rotate with angular velocity . The lengths of these phasors represent the peak current Im and voltage Vm. The y components are Im sin (t) = i(t) and voltage Vm sin (t)= v(t). You can compare i(t) and v(t) in the animation with the vertical components of the phasors. The animation and phasor diagram here are simple, but they will become more useful when we consider components with different phases and with frequency dependent behaviour. (See Physclips for a comparison of simple harmonic motion and circular motion)

What are impedance and reactance?

Circuits in which current is proportional to voltage are called linear circuits. (As soon as one inserts diodes and transistors, circuits cease to be linear, but that's another story.) The ratio of voltage to current in a resistor is its resistance. Resistance does not depend on frequency, and in resistors the two are in phase, as we have seen in the animation. However, circuits with only resistors are not very interesting. In general, the ratio of voltage to current does depend on frequency and in general there is a phase difference. So impedance is the general name we give to the ratio of voltage to current. It has the symbol Z. Resistance is a special case of impedance. Another special case is that in which the voltage and current are out of phase by 90: this is an important case because when this happens, no power is lost in the circuit. In this case where the voltage and current are out of phase by 90, the ratio of voltage to current is called the reactance, and it has the symbol X. We return to summarise these terms and give expressions for them below in the section Impedance of components, but first let us see why there are frequency dependence and phase shifts for capacitors and for inductors.

Capacitors and charging

The voltage on a capacitor depends on the amount of charge you store on its plates. The current flowing onto the positive capacitor plate (equal to that flowing off the negative plate) is by definition the rate at which charge is being stored. So the charge Q on the capacitor equals the integral of the current with respect to time. From the definition of the capacitance, vC = q/C, so

Now remembering that the integral is the area under the curve (shaded blue), we can see in the next animation why the current and voltage are out of phase. Once again we have a sinusoidal current i = Im . sin (t), so integration gives

(The constant of integration has been set to zero so that the average charge on the capacitor is 0). Now we define the capacitive reactance XC as the ratio of the magnitude of the voltage to magnitude of the current in a capacitor. From the equation above, we see that XC= 1/C. Now we can rewrite the equation above to make it look like Ohm's law. The voltage is proportional to the current, and the peak voltage and current are related by Vm = XC.Im. Note the two important differences. First, there is a difference in phase: the integral of the sinusoidal current is a negative cos function: it reaches its maximum (the capacitor has maximum charge) when the current has just finished flowing forwards and is about to start flowing backwards. Run the animation again to make this clear. Looking at the relative phase, the voltage across the capacitor is 90, or one quarter cycle, behind the current. We can see also see how the = 90 phase difference affects the phasor diagrams at right. Again, the vertical component of a phasor arrow represents the instantaneous value of its quanitity. The phasors are rotating counter clockwise (the positive direction) so the phasor representing VC is 90 behind the current (90 clockwise from it). Recall that reactance is the name for the ratio of voltage to current when they differ in phase by 90. (If they are in phase, the ratio is called resistance.) Another difference between reactance and resistance is that the reactance is frequency dependent. From the algebra above, we see that the capacitive reactance XC decreases with frequency . This is shown in the next animation: when the frequency is halved but the current amplitude kept constant, the capacitor has twice as long to charge up, so it generates twice the potential difference. The blue shading shows q, the integral under the current curve (light for positive, dark for negative). The second and fourth curves show V C = q/C .

See how the lower frequency leads to a larger charge (bigger shaded area before changing sign) and therefore a larger VC. Thus for a capacitor, the ratio of voltage to current decreases with frequency. We shall see later how this can be used for filtering different frequencies.

Inductors and the Farady emf

An inductor is usually a coil of wire. In an ideal inductor, the resistance of this wire is negligibile, as is its capacitance. The voltage that appears across an inductor is due to its own magnetic field and Faraday's law of electromagnetic induction. The current i(t) in the coil sets up a magnetic field, whose magnetic flux B is proportional to the field strength, which is proportional to the current flowing. (Do not confuse the phase with the flux B.) So we define the (self) inductance of the coil thus: B(t) = L.i(t) Faraday's law gives the emf EL = - dB/dt. Now this emf is a voltage rise, so for the voltage drop vL across the inductor, we have:

Again we define the inductive reactance XL as the ratio of the magnitudes of the voltage and current, and from the equation above we see that XL = L. Again we note the analogy to Ohm's law: the voltage is proportional to the current, and the peak voltage and currents are related by Vm = XL.Im. Remembering that the derivative is the local slope of the curve (the purple line), we can see in the next animation why voltage and current are out of phase in an inductor.

Again, there is a difference in phase: the derivative of the sinusoidal current is a cos function: it has its maximum (largest voltage across the inductor) when the current is changing most rapidly, which is when the current is intantaneously zero. The animation should make this clear. The voltage across the ideal inductor is 90 ahead of the current, (ie it reaches its peak one quarter cycle before the current does). Note how this is represented on the phasor diagram. Again we note that the reactance is frequency dependent XL = L. This is shown in the next animation: when the frequency is halved but the current amplitude kept constant, the current is varying only half as quickly, so its derivative is half as great, as is the Faraday emf. For an inductor, the ratio of voltage to current increases with frequency, as the next animation shows.

Impedance of components

Let's recap what we now know about voltage and curent in linear components. The impedance is the general term for the ratio of voltage to current. Resistance is the special case of impedance when = 0, reactance the special case when = 90. The table below summarises the impedance of the different components. It is easy to remember that the voltage on the capacitor is behind the current, because the charge doesn't build up until after the current has been flowing for a while.

The same information is given graphically below. It is easy to remember the frequency dependence by thinking of the DC (zero frequency) behaviour: at DC, an inductance is a short circuit (a piece of wire) so its impedance is zero. At DC, a capacitor is an open circuit, as its circuit diagram shows, so its impedance goes to infinity.

RC Series combinations

When we connect components together, Kirchoff's laws apply at any instant. So the voltage v(t) across a resistor and capacitor in series is just vseries(t) = vR(t) + vC(t) however the addition is complicated because the two are not in phase. The next animation makes this clear: they add to give a new sinusoidal voltage, but the amplitude isless than VmR(t) + VmC(t). Similarly, the AC voltages (amplitude times 2) do not add up. This may seem confusing, so it's worth repeating: vseries = vR + vC Vseries < VR + VC. but

This should be clear on the animation and the still graphic below: check that the voltages v(t) do add up, and then look at the magnitudes. The amplitudes and the RMS voltages V do not add up in a simple arithmetical way. Here's where phasor diagrams are going to save us a lot of work. Play the animation again (click play), and look at the projections on the vertical axis. Because we have sinusoidal variation in time, the vertical component (magnitude times the sine of the angle it makes with the x axis) gives us v(t). But the y components of different vectors, and therefore phasors, add up simply: if

rtotal = r1 + r2, then ry total = ry1 + ry2. So v(t), the sum of the y projections of the component phasors, is just the y projection of the sum of the component phasors. So we can represent the three sinusoidal voltages by their phasors. (While you're looking at it, check the phases. You'll see that the series voltage is behind the current in phase, but the relative phase is somewhere between 0 and 90, the exact value depending on the size of VR and VC. We'll discuss phase below.) Now let's stop that animation and label the values, which we do in the still figure below. All of the variables (i, vR, vC, vseries) have the same frequency f and the same angular frequency , so their phasors rotate together, with the same relative phases. So we can 'freeze' it in time at any instant to do the analysis. The convention I use is that the x axis is the reference direction, and the reference is whatever is common in the circuit. In this series circuit, the current is common. (In a parallel circuit, the voltage is common, so I would make the voltage the horizontal axis.) Be careful to distinguish v and V in this figure!

(Careful readers will note that I'm taking a shortcut in these diagrams: the size of the arrows on the phasor diagrams are drawn the same as the amplitudes on the v(t) graphs. However I am just calling them VR, VC etc, rather than VmR, VmR etc. The reason is that the peak values (VmR etc) are rarely used in talking about AC: we use the RMS values, which are peak values times 0.71. Phasor diagrams in RMS have the same shape as those drawn using amplitudes, but everything is scaled by a factor of 0.71 = 1/2.)

The phasor diagram at right shows us a simple way to calculate the series voltage. The components are in series, so the current is the same in both. The

voltage phasors (brown for resistor, blue for capacitor in the convention we've been using) add according to vector or phasor addition, to give the series voltage (the red arrow). By now you don't need to look at v(t), you can go straight from the circuit diagram to the phasor diagram, like this:

From Pythagoras' theorem: V2mRC = V2mR + V2mC If we divide this equation by two, and remembering that the RMS value V = Vm/2, we also get:

Now this looks like Ohm's law again: V is proportional to I. Their ratio is the series impedance, Zseries and so for this series circuit,

Note the frequency dependence of the series impedance ZRC: at low frequencies, the impedance is very large, because the capacitive reactance 1/C is large (the capacitor is open circuit for DC). At high frequencies, the capacitive reactance goes to zero (the capacitor doesn't have time to charge up) so the series impedance goes to R. At the angular frequency = o = 1/RC, the capacitive reactance 1/C equals the resistance R. We shall show this characteristic frequency on all graphs on this page. Remember how, for two resistors in series, you could just add the resistances: Rseries = R1 + R2 to get the resistance of the series combination. That simple result comes about because the two voltages are both in phase with the current, so their phasors are parallel. Because the phasors for reactances are 90 out of phase with the current, the series impedance of a resistor R and a reactance X are given by Pythagoras' law: Zseries2 = R2 + X2 . Ohm's law in AC. We can rearrange the equations above to obtain the current flowing in this circuit. Alternatively we can simply use the Ohm's Law analogy and say that I = Vsource/ZRC. Either way we get

where the current goes to zero at DC (capacitor is open circuit) and to V/R at high frequencies (no time to charge the capacitor). So far we have concentrated on the magnitude of the voltage and current. We now derive expressions for their relative phase, so let's look at the phasor diagram again.

From simple trigonometry, the angle by which the current leads the voltage is tan-1 (VC/VR) = tan-1 (IXC/IR) = tan-1 (1/RC) = tan-1 (1/2fRC). However, we shall refer to the angle by which the voltage leads the current. The voltage is behind the current because the capacitor takes time to charge up, so is negative, ie = tan-1 (1/RC) = tan-

(1/2fRC).

(You may want to go back to the RC animation to check out the phases in time.) At low frequencies, the impedance of the series RC circuit is dominated by the capacitor, so the voltage is 90 behind the current. At high frequencies, the impedance approaches R and the phase difference approaches zero. The frequency dependence of Z and are important in the applications of RC circuits. The voltage is mainly across the capacitor at low frequencies, and mainly across the resistor at high frequencies. Of course the two voltages must add up to give the voltage of the source, but they add up as vectors. V2RC = V2R + V2C. At the frequency = o = 1/RC, the phase = 45 and the voltage fractions are VR/VRC = VC/VRC = 1/2V1/2 = 0.71.

So, by chosing to look at the voltage across the resistor, you select mainly the high frequencies, across the capacitor, you select low frequencies. This brings us to one of the very important applications of RC circuits, and one which merits its own page: filters, integrators and differentiators where we use sound files as examples of RC filtering.

RL Series combinations

In an RL series circuit, the voltage across the inductor is aheadof the current by 90, and the inductive reactance, as we saw before, is XL = L. The resulting v(t) plots and phasor diagram look like this.

It is straightforward to use Pythagoras' law to obtain the series impedance and trigonometry to obtain the phase. We shall not, however, spend much time on RL circuits, for three reasons. First, it makes a good exercise for you to do it yourself. Second, RL circuits are used much less than RC circuits. This is because inductors are always* too big, too expensive and the wrong value, a claim you can check by looking at an electronics catalogue. If you can use a circuit involving any number of Rs, Cs, transistors, integrated circuits etc to replace an inductor, one usually does. The third reason why we don't look closely at RL circuits on this site is that you can simply look at RLC circuits (below) and omit the phasors and terms for the capacitance.

* Exceptions occur at high frequencies (~GHz) where only small value Ls are required to get substantial L. In such circuits, one makes an inductor by twisting copper wire around a pencil and adjusts its value by squeezing it with the fingers.

RLC Series combinations

Now let's put a resistor, capacitor and inductor in series. At any given time, the voltage across the three components in series, vseries(t), is the sum of these: vseries(t) = vR(t) + vL(t) + vC(t), The current i(t) we shall keep sinusoidal, as before. The voltage across the resistor, vR(t), is in phase with the current. That across the inductor, vL(t), is 90 ahead and that across the capacitor, vC(t), is 90 behind. Once again, the time-dependent voltages v(t) add up at any time, but the RMS voltages V do not simply add up. Once again they can be added by phasors representing the three sinusoidal voltages. Again, let's 'freeze' it in time for the purposes of the addition, which we do in the graphic below. Once more, be careful to distinguish v and V.

Look at the phasor diagram: The voltage across the ideal inductor is antiparallel to that of the capacitor, so the total reactive voltage (the voltage which is 90 ahead of the current) is VL - VC, so Pythagoras now gives us: V2series = V2R + (VL - VC)2 Now VR = IR, VL = IXL = IL and VC = IXC= I/C. Substituting and taking the common factor I gives:

where Zseries is the series impedance: the ratio of the voltage to current in an RLC series ciruit. Note that, once again, reactances and resistances add according to Pythagoras' law: Zseries2 = R2 + Xtotal2 = R2 + (XL XC)2. Remember that the inductive and capacitive phasors are 180 out of phase, so their reactances tend to cancel. Now let's look at the relative phase. The angle by which the voltage leads the current is = tan-1 ((VL - VC)/VR). Substituting VR = IR, VL = IXL = IL and VC = IXC= I/C gives:

The dependence of Zseries and on the angular frequency is shown in the next figure. The angular frequency is given in terms of a particular value o, the resonant frequency (o2 = 1/LC), which we meet below.

(Setting the inductance term to zero gives back the equations we had above for RC circuits, though note that phase is negative, meaning (as we saw above) that voltage lags the current. Similarly, removing the capacitance terms gives the expressions that apply to RL circuits.) The next graph shows us the special case where the frequency is such that VL = VC.

Because vL(t) and vC are 180 out of phase, this means that vL(t) = vC(t), so the two reactive voltages cancel out, and the series voltage is just equal to that across the resistor. This case is called series resonance, which is our next topic.

Resonance

Note that the expression for the series impedance goes to infinity at high frequency because of the presence of the inductor, which produces a large emf if the current varies rapidly. Similarly it is large at very low frequencies because of the capacitor, which has a long time in each half cycle in which to charge up. As we saw in the plot of Zseries above, there is a minimum value of the series impedance, when the voltages across capacitor and inductor are equal and opposite, ie vL(t) = vC(t) so VL(t) = VC, so L = 1/C so the frequency at which this occurs is

where o and fo are the angular and cyclic frequencies of resonance, respectively. At resonance, series impedance is a minimum, so the voltage for a given current is a minimum (or the current for a given voltage is a maximum).
This phenomenon gives the answer to our teaser question at the beginning. In an RLC series circuit in which the inductor has relatively low internal resistance r, it is possible to have a large voltage across the the inductor, an almost equally large voltage across capacitor but, as the two are nearly 180 degrees out of phase, their voltages almost cancel, giving a total series voltage that is quite small. This is one way to produce a large voltage oscillation with only a small voltage source. In the circuit diagram at right, the coil corresponds to both the inducance L and the resistance r, which is why they are drawn inside a box representing the physical component, the coil. Why are they in series? Because the current flows through the coil and thus passes through both the inductance of the coil and its resistance.

You get a big voltage in the circuit for only a small voltage input from the power source. You

are not, of course, getting something for nothing. The energy stored in the large oscillations is gradually supplied by the AC source when you turn on, and it is then exchanged between capacitor and inductor in each cycle. For more details about this phenomenon, and a discussion of the energies involved, go toLC oscillations.

Bandwidth and Q factor At resonance, the voltages across the capacitor and the pure inductance cancel out, so the series impedance takes its minimum value: Zo = R. Thus, if we keep the voltage constant, the current is a maximum at resonance. The current goes to zero at low frequency, because XC becomes infinite (the capacitor is open circuit for DC). The current also goes to zero at high frequency because XL increases with (the inductor opposes rapid changes in the current). The graph shows I() for circuit with a large resistor (lower curve) and for one with a small resistor (upper curve). A circuit with low R, for a given L and C, has a sharp resonance. Increasing the resistance makes the resonance less sharp. The former circuit is more selective: it produces high currents only for a narrow bandwidth, ie a small range of or f. The circuit with higher R responds to a wider range of frequencies and so has a larger bandwidth. The bandwidth (indicated by the horiztontal bars on the curves) is defined as the difference between the two frequencies + and - at which the circuit converts power at half the maximum rate.

Now the electrical power converted to heat in this circuit is I2R, so the maximum power is converted at resonance, = o. The circuit converts power at half this rate when the current is Io/2. The Q value is defined as the ratio
Q = o/.

Complex impedance

You have perhaps been looking at these phasor diagrams, noticing that they are all two-dimensional, and thinking that we could simply use the complex

plane. Good idea! But not original: indeed, that is the most common way to analyse such circuits. The only difference from the presentation here is to consider cosusoids, rather than sinusoids. In the animations above, we used sin waves so that the vertical projection of the phasors would correspond to the height on the v(t) graphs. In complex algebra, we use cos waves and take their projections on the (horizontal) real axis. The phasor diagrams have now become diagrams of complex numbers, but otherwise look exactly the same. They still rotate at t, but in the complex plane. The resistor has a real impedance R, the inductor's reactance is a positive imaginary impedance XL = jL and the capacitor has a negative imaginary impedance XC = j.1/C = 1/jC. Consequently, using bold face for complex quantities, we may write: Zseries = (R2 + (jL + 1/jC)2)1/2 and so on. The algebra is relatively simple. The magnitude of any complex quantity gives the magnitude of the quantity it represents, the phase angle its phase angle. Its real component is the component in phase with the reference phase, and the imaginary component is the component that is 90 ahead.
http://www.animations.physics.unsw.edu.au/jw/AC.html

What is Alternating Current?

Alternating Current vs. Direct Current

The figure to the right shows the schematic diagram of a very basic DC circuit. It consists of nothing more than a source (a producer of electrical energy) and a load (whatever is to be powered by that electrical energy). The source can be any electrical source: a chemical battery, an electronic power supply, a mechanical generator, or any other possible continuous source of electrical energy. For simplicity, we represent the source in this figure as a battery. At the same time, the load can be any electrical load: a light bulb, electronic clock or watch, electronic instrument, or anything else that must be driven by a continuous source of electricity. The figure here represents the load as a simple resistor. Regardless of the specific source and load in this circuit, electrons leave the negative terminal of the source, travel through the circuit in the direction shown by the arrows, and eventually return to the positive terminal of the source. This action continues for as long as a complete electrical circuit exists.

Now consider the same circuit with a single change, as shown in the second figure to the right. This time, the energy source is constantly changing. It begins by building up a voltage which is positive on top and negative on the bottom, and therefore pushes electrons through the circuit

in the direction shown by the solid arrows. However, then the source voltage starts to fall off, and eventually reverse polarity. Now current will still flow through the circuit, but this time in the direction shown by the dotted arrows. This cycle repeats itself endlessly, and as a result the current through the circuit reverses direction repeatedly. This is known as analternating current. This kind of reversal makes no difference to some kinds of loads. For example, the light bulbs in your home don't care which way current flows through them. When you close the circuit by turning on the light switch, the light turns on without regard for the direction of current flow. Of course, there are some kinds of loads that require current to flow in only one direction. In such cases, we often need to convert alternating current such as the power provided at your wall socket to direct current for use by the load. There are several ways to accomplish this, and we will explore some of them in later pages in this section.

Properties of Alternating Current

A DC power source, such as a battery, outputs a constant voltage over time, as depicted in the top figure to the right. Of course, once the chemicals in the battery have completed their reaction, the battery will be exhausted and cannot develop any output voltage. But until that happens, the output voltage will remain essentially constant. The same is true for any other source of DC electricity: the output voltage remains constant over time.

By contrast, an AC source of electrical power changes constantly in amplitude and regularly changes polarity, as shown in the second figure to the right. The changes are smooth and regular, endlessly repeating in a succession of identical cycles, and form a sine wave as depicted here. Because the changes are so regular, alternating voltage and current have a number of properties associated with any such waveform. These basic properties include the following list:

Frequency. One of the most important properties of any regular waveform identifies the number of complete cycles it goes through in a fixed period of time. For standard measurements, the period of time is one second, so the frequency of the wave is commonly measured in cycles per second (cycles/sec) and, in normal usage, is expressed in units of Hertz (Hz). It is represented in mathematical equations by the letter 'f.' In North America (primarily the US and Canada), the AC power system operates at a frequency of 60 Hz. In Europe, including the UK, Ireland, and Scotland, the power system operates at a frequency of 50 Hz. Period. Sometimes we need to know the amount of time required to complete one cycle of the waveform, rather than the number of cycles per second of time. This is logically the reciprocal of frequency. Thus, period is the time duration of one cycle of the waveform, and is measured in seconds/cycle. AC power at 50 Hz will have a period of 1/50 = 0.02 seconds/cycle. A 60 Hz power system has a period of 1/60 = 0.016667 seconds/cycle. These are often expressed as 20 ms/cycle or 16.6667 ms/cycle, where 1 ms is 1 millisecond = 0.001 second (1/1000 of a second).

Wavelength. Because an AC wave moves physically as well as changing in time, sometimes we need to know how far it moves in one cycle of the wave, rather than how long that cycle takes to complete. This of course depends on how fast the wave is moving as well. Electrical signals travel through their wires at nearly the speed of light, which is very nearly 3 108 meters/second, and is represented mathematically by the letter 'c.' Since we already know the frequency of the wave in Hz, or cycles/second, we can perform the division of c/f to obtain a result in units of meters/cycle, which is what we want. The Greek letter (lambda) is used to represent wavelength in mathematical expressions. Thus, = c/f. As shown in the figure to the right, wavelength can be measured from any part of one cycle to the equivalent point in the next cycle. Wavelength is very similar to period as discussed above, except that wavelength is measured in distance per cycle where period is measured in time per cycle. Amplitude. Another thing we have to know is just how positive or negative the voltage is, with respect to some selected neutral reference. With DC, this is easy; the voltage is constant at some measurable value. But AC is constantly changing, and yet it still powers a load. Mathematically, the amplitude of a sine wave is the value of that sine wave at its peak. This is the maximum value, positive or negative, that it can attain. However, when we speak of an AC power system, it is more useful to refer to the effective voltage or current. This is the rating that would cause the same amount of work to be done (the same effect) as the same value of DC voltage or current would cause. We won't cover the mathematical derivations here; for the present, we'll simply note that for a sine wave, the effective voltage of the AC power system is 0.707 times the peak voltage. Thus, when we say that the AC line voltage in the US is 120 volts, we are referring to the voltage amplitude, but we are describing the effective voltage, not the peak voltage of nearly 170 volts. The effective voltage is also known as the rms voltage.

When we deal with AC power, the most important of these properties are frequency and amplitude, since some types of electrically powered equipment must be designed to match the frequency and voltage of the power lines. Period is sometimes a consideration, as we'll discover when we explore electronic power supplies. Wavelength is not generally important in this context, but becomes much more important when we start dealing with signals at considerably higher frequencies.

Why Use Alternating Current?

Since some kinds of loads require DC to power them and others can easily operate on either AC or DC, the question naturally arises, "Why not dispense entirely with AC and just use DC for everything?" This question is augmented by the fact that in some ways AC is harder to handle as well as to use. Nevertheless, there is a very practical reason, which overrides all other considerations for a widely distributed power grid. It all boils down to a question of cost. DC does get used in some local commercial applications. An excellent example of this is the electric trolley car and trolley bus system used in San Francisco, for public transportation. Trolley cars are electric train cars with power supplied by an overhead wire. Trolley busses are like any other bus, except they are electrically powered and get their power from two overhead wires. In both cases, they operate on 600 volts DC, and the overhead wires span the city. The drawback is that most of the electrical devices on each car or bus, including all the light bulbs inside, are quite standard and require 110 to 120 volts. At the same time, however, if we were to reduce the system voltage, we would have to increase the amount of current drawn by each car or bus in order to provide the same amount of power to it. (Power is equal to the product of the applied voltage and the resulting current: P = I E.) But those overhead wires are not perfect conductors; they exhibit some resistance. They will absorb some energy from the electrical current and dissipate it as waste heat, in accordance with Ohm's Law (E = I R). With a small amount of algebra, we can note that the lost power can be expressed as: Plost = IR Now, if we reduce the voltage by a factor of 5 (to 120 volts DC), we must increase the current by a factor of 5 to maintain the same power to the trolley car or bus. But lost power is a function of the square of the current, so we will lose not five times as much power in the resistance of the wires, but twenty-five times as much power. To offset and minimize that loss, we

would have to use much larger wires, and pay a high price for all that extra copper. A cheaper solution is to mount a motor-generator set in each trolley car and bus, using a 600 volt dc motor and a lower-voltage generator to power all the equipment aboard that car. The same reality of Ohm's Law and resistive losses holds true in the country-wide power distribution system. We need to keep the voltage used in homes to a reasonable and relatively safe value, but at the same time we need to minimize resistive losses in the transmission wires, without bankrupting ourselves buying heavy-gauge copper wire. At the same time, we can't use motor-generator pairs all across the country; they would need constant service and would break down far too often. We need a system that allows us to raise the voltage (and thus reduce the current) for longdistance transmission, and then reduce the voltage again (to a safe value) for distribution to individual homes and businesses. And we need to do this without requiring any moving parts to break down or need servicing. The answer is to use an AC power system and transformers. (We'll learn far more about transformers in a later page; for now, a transformer is an electrical component that can convert incoming AC power at one voltage to outgoing power at a different voltage, higher or lower, with only very slight losses.) Thus, we can generate electricity at a reasonable voltage for practical AC generators (sometimes called alternators), then use transformers to step that voltage up to very high levels for long-distance transmission, and then use additional transformers to step that high voltage back down for local distribution to individual homes. In practice, this is done in stages. The really high-voltage transmission lines hanging from long glass insulators on the arms of tall steel towers carry electricity cross-country at several hundred thousand volts. This is stepped down to about 22,000 volts for distribution to multiple neighborhoods these are the wires you see at the top of the telephone poles in many areas. Additional transformers mounted on some of these telephone poles step this voltage down again for distribution to several homes each. The design of the system minimizes the overall cost by balancing the cost of transformers against the cost of heavier-gauge copper wire, as well as the cost of maintaining the system and repairing damage. This is how the cost of electricity delivered to your home or business is kept to a minimum, while maintaining a very high level of service.

A Note on Nomenclature

When we examined DC circuit theory and discussed DC power losses above, we used capital letters to represent all quantities. This is standard nomenclature; capital letters are used to represent fixed, static values. Thus, we use a capital I to represent DC current and a capital V to represent DC voltage. By the same token, we use the capital letters R, L, and C to represent the values of circuit components containing resistance, inductance, and capacitance, respectively. These are fixed values that are set at the time the component is manufactured. On the other hand, AC circuits use constantly-changing voltages and currents. In addition, some circuits contain both AC and DC components, which must often be considered separately. Therefore, we typically use lower-case letters to designate instantaneous values of voltage, current and power. Thus, if you see an equation written as: P=IE

you know it refers to DC power, current, and voltage. On the other hand, an equation written as: p=ie

refers to the instantaneous power, current, and voltage of some AC signal at some specified instant in time. There will also be cases where we must refer to an overall AC signal rather than one instantaneous value from it. In such cases, the use of upperand lower-case letters may be adjusted so we don't have too many of one or the other in a single discussion. We will try to avoid unnecessary confusion by including subscripts in equations, or specifying in the text exactly what we are discussing.

http://www.play-hookey.com/ac_theory/

You might also like