Electrical

Published on February 2017 | Categories: Documents | Downloads: 63 | Comments: 0 | Views: 620
of 35
Download PDF   Embed   Report

Comments

Content

Two Current Flow theories exist. The first is: ELECTRON THEORY The Electron Theory states that current flows from NEGATIVE to POSITIVE. Electrons move from atom to atom as they move through the conductor towards positive.

The second Current Flow theory is: CONVENTIONAL THEORY Conventional theory, also known as HOLE THEORY, states that current flows from POSITIVE to NEGATIVE. Protons or the lack of electrons (the holes) move towards the negative. (Current flow direction in Hole Theory is the opposite of that in Electron Theory.) VOLTAGE Voltage is the electrical force that moves electrons through a conductor. Voltage is electrical pressure also known as EMF (Electro Motive Force) that pushes electrons. The greater the difference in electrical potential push (difference between positive and negative), the greater the voltage force potential.

CURRENT (AMPERES) CURRENT is the quantity or flow rate of electrons moving past a point within one second. Current flow is also known as amperage, or amps for short. Higher voltage will produce higher current flow, and lower

voltage will produce lower current flow.

RESISTANCE Resistance is the force that reduces or stops the flow of electrons. It opposes voltage. Higher resistance will decrease the flow of electrons and lower resistance will allow more electrons to flow.

TYPES OF ELECTRICITY Two basic types of Electricity classifications: STATIC ELECTRICITY is electricity that is standing still. Voltage potential with NO electron flow. DYNAMIC ELECTRICITY is electricity that is in motion. Voltage potential WITH electron flow. Two types of Dynamic electricity exist: Direct Current (DC) Electron Flow is in only one direction.

Alternating Current (AC) Electron flow alternates and flows in both directions (back and forth). DIRECT CURRENT (DC) Electricity with electrons flowing in only one direction is called Direct Current or DC. DC electrical systems are used in cars.

ALTERNATING CURRENT (AC) Electricity with electrons flowing back and forth, negative - positive- negative, is called Alternating Current, or AC. The electrical appliances in your home use AC power.

AN ELECTRICAL CIRCUIT The circuit shown below has a power source, fuse, switch, two lamps and wires connecting each into a loop or circle. When the connection is complete, current flows from the positive terminal of the battery through the wire, the fuse, the switch, another wire, the lamps, a wire and to the negative terminal of the battery. The route along which the electricity flows is called an electrical circuit. ELECTRICAL CIRCUIT REQUIREMENTS A complete Electrical Circuit is required in order to make electricity practical. Electrons must flow from and return to the power source. There are three different circuit types, all require the same basic components: 1. Power Source is needed to supply the flow of electrons (electricity). 2. Protection Device prevents damage to the circuit in the event of a short.

3. Load Device converts the electricity into work. 4. Control Device allows the user control to turn the circuit on or off 5. Conductors provide an electrical path to and from the power source.

BASIC CIRCUIT CONSTRUCTION 1. Power Source (Battery, Alternator, Generator, etc.) 2. Protection Device (Fuse, Fusible Link, or Circuit Breaker) 3. Load Device (Lamp, Motor, Winding, Resistor, etc. 4. Control (Switch, Relay, or Transistor) 5. Conductors (A Return Path, Wiring to Ground)

OHM'S LAW FORMULA When voltage is applied to an electrical circuit, current flows in the circuit. The following special relationship exists among the voltage, current and resistance within the circuit: the

size of the current that flows in a circuit varies in proportion to the voltage which is applied to the circuit, and in inverse proportion to the resistance through which it must pass. This relationship is called Ohm's law, and can be expressed as follows:

E=IR
Voltage = Current x Resistance E Voltage applied to the circuit, in volts (V) I Current flowing in the circuit, in amperes (A) R Resistance in the circuit, in ohms

In practical terms "V = I x R" which means "Voltage = Current x Resistance". 1 volt will push one amp through 1 ohm of resistance. NOTE: E = IR, V=AR, or V=IR are all variations of the same formula. How you learned Ohm's law will determine which one you will use. Personal preference is the only difference; anyone will get you the correct answer. OHM'S LAW SYMBOL SHORTCUT Mathematical formulas can be difficult for many who don't use them regularly. Most people can remember a picture easier than a mathematical formula. By using the Ohms law symbol below, anyone can remember the correct formula to use. By knowing any two values you can figure out the third. Simply put your finger over the portion of the symbol you are trying to figure out and you have your formula.

APPLICATIONS OF OHM'S LAW As an application of Ohm's law, any voltage V, current I or resistance R in an electrical circuit can be determined without actually measuring it if the two others values are known. This law can be used to determine the amount of current I flowing in the circuit when voltage V is applied to resistance R. As stated previously, Ohm's law is:

Current = Voltage / Resistance.

In the following circuit, assume that resistance R is 2 and voltage V that is applied to it is 12 V. Then, current I flowing in the circuit can be determined as follows:

This law can also be used to determine the voltage V that is needed to permit current I to pass through resistance R: V = I x R (Voltage= Current x Resistance). In the following circuit, assume that resistance R is 4 ohms. The voltage V that is necessary to permit a current I of 3 A to pass through the resistance can be determined as follows:

SERIES CIRCUIT CALCULATIONS If, for example, two or more lamps (resistances R1 and R2, etc.) are connected in a circuit as follows, there is only one route that the current can take. This type of connection is called a series connection. The value of current I is always the same at any point in a series circuit.

The combined resistance RO in this circuit is equal to the sum of individual resistance R1 and R2. In other words: The total resistance(RO) is equal to the sum of all resistances (R1 + R2 + R3 + .......)

Therefore, the strength of current (I) flowing in the circuit can be found as follows: Resistance R0 (a combination of resistances R1 and R2, which are connected in series in the circuit as illustrated) and current I flowing in this circuit can be determined as follows:

VOLTAGE DROP TOTAL When more than one load exists in a circuit, the voltage divides and will be shared among the loads. The sum of the voltage drops equal source voltage. The higher the resistance the higher the voltage drop. Depending on the resistance, each load will have a different voltage drop. 0V + 5V + 7V + 0V = 12V

The voltage drop across resistances R1 and R2 in the following circuit can be determined as follows:

PARALLEL CIRCUIT In parallel connection, two or more resistances (R1, R2, etc.) are connected in a circuit as follows, with one end of each resistance connected to the high (positive) side of the circuit, and one end connected to the low (negative) side. Full battery voltage is applied to all resistances within a circuit having a parallel connection.

Resistance R0 (a combination of resistances R1 and R2) in a parallel connection can be determined as follows:

From the above, the total current I flowing in this circuit can be determined from Ohm's law as follows:

The total current I is also equal to the sum of currents I1 and I2 flowing through individual resistances R1 and R2

Since battery voltage V is applied equally to all resistances, the strength of currents I1 and I2 can be determined from Ohm's law as follows:

SERIES PARALLEL CIRCUIT A series-parallel circuit has some components in series and others in parallel. The power source and control or protection devices are usually in series; the loads are usually in parallel. The same current flows in the series portion, different currents in the parallel portion. The same voltage is applied to parallel devices, different voltages to series devices. If the series portion is broken, current stops flowing in the entire circuit. If a parallel branch is broken, current continues flowing in the series portion and the remaining branches.

SERIES-PARALLEL CIRCUIT A resistance and lamps may be connected in a circuit as illustrated below. This type of connecting method is called series-parallel connection, and is a combination of series and parallel connections. The interior dash board lights are a good example. By adjusting the rheostat, you can increase or decrease the brilliance of the lights.

The combined resistance R02 in this series-parallel connection can be determined in the following order: a. Determine combined resistance R01, which is a combination of resistances R2 and R3 connected in parallel. b. Then, determine resistance R02, which is a combination of resistance R1 and combined resistance R01 connected in series.

Total current I flowing in the circuit can be determined from Ohm's law as follows:

The voltage applied to R2 and R3 can be found by the following formula:

Currents I1, I2 and I flowing through resistances R1, R2 and R3 in the series-parallel connection, as shown below, can be determined as follows:

SOLENOIDS - PULLING TYPE

A solenoid is an electromagnetic switch that converts current flow into mechanical movement. As current flows through the winding a magnetic field is created. The magnetic field will pull the moveable iron core into the center of the winding. This type of solenoid is called a "pulling" type solenoid, as the magnetic field pulls the moveable iron core into the coil. A common use for pulling solenoids are in the starting system. The starter solenoid engages the starter with the flywheel.

SOLENOIDS - PULLING TYPE OPERATION As current flows through the winding a magnetic field is created. These magnetic lines of force want to be as small as possible. If an iron core is placed near the coil that has current flowing through it, the magnetic field will stretch out like a rubber band, reaching out and pulling the iron bar into the center of the coil.

The force law for charge is similar to the gravitational force law. For two charges, q1 and q2, the force between them is:
y y

y

Proportional to the product of the two charges q1 and q2, and Inversely proportional to the square of the distance, r (in meters), between them. So, the force is given by an expression:

F1,2 = q1 q2/(4TIo r2)

Here, Io is a fundamental constant of nature, = ~8.885419x10-12 F/m. Like every other physical quantity, when you deal with charge you must account for units.
y y

y

y

Charge is measured in coulombs. Coulombs are named after Charles Augustin Coulomb who was the first person to determine that the force law for charges was an inverse square law. Charge not only comes in two varieties, it also comes in discrete sizes. Electrons and protons each have the same size charge (but of opposite polarity) of magnitude 1.6x 10-19 coulombs (+ or - as appropriate), where a coulomb is the MKS unit of charge. Note, the electron's charge is usually counted as negative, and the proton's charge as positive, although that is only a convention and there is no "lack" or "surplus" associated with negative and positive charges.

When you use the force law expression: F1,2 = q1 q2/(4TIo r2)
y y y

Charge is measured in couloumbs (for q1 and q2). Force is measured in newtons (for F1,2 ). Distance is measured in meters (for r).

Why Bode' Plots? Bode plots are the most widely used means of displaying and communicating frequency response information. There are many reasons for that.

Bode' plots are really log-log plots, so they collapse a wide range of frequencies (on the horizontal axis) and a wide range of gains (on thevertical axis) into a viewable whole. In Bode' plots, commonly encountered frequency responses have a shape that is simple. That simple shape means that laboratory measurementscan easily be discerned to have the common factors that lead to those shapes. For example, first order systems have two straight line asymptotes and if you take data and plot

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

y

y

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

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

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

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

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

The low frequency asymptote,

The high frequency asymptote, The "mid point" where [X = 1 That's at f = 159 Hz. Let's look at the low frequency asymptote first. Here's the transfer function. G(j[) = 1/(j[X+ 1) If [ is small, then the imaginary term in the denominator is small, and we have: G(j[) ~= 1/(j0 + 1) = 1 The low frequency behavior of the plot shows that the plot is flat at a value of 1. Now, let's look at the high frequency asymptote. Here's the transfer function. G(jw) = 1/(j [X+ 1) If [ is large, then the imaginary term in the denominator dominates, and we have: G(j [) ~= 1/j [X The magnitude of the gain is: |G(j[)| ~= 1/[X The gain drops off inversely with frequency, but the Bode' plot drops off as a straight line. Hmmmm? That's very interesting - that it is a straight line. The straight line high frequency asymptote shouldn't be cause for consternation. If we have: |G([)| ~= 1/[X Remember that the Bode' plot is log gain vs log frequency, so let's look at the logarithm of the magnitude of the gain. log(|G(j[)|) = log(1/[X) = -log([X)

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

Check that conclusion on the plot to be sure you understand what it means. Here is a plot with the lower limit extended.

Check going from f = 300 to f = 3000. Does the gain decrease by a factor of 10 when the frequency increases by a factor of 10? The last point we need to examine is the behavior of the frequency response for frequencies between high frequency and low frequency - what we referred to as the mid-point earlier. If the frequency response function is given by: G(j[) = 1/(j[X + 1) If [ = 1/Xthen (taking that frequency as the mid-point), we have: G(j[) = 1/(j + 1) The magnitude of the gain is: |G(j[)| = 1/|j + 1| = 1/sqrt(2) ~= 0.707 This point is at [ = 1000, or f =159Hz. There are some interesting things to note about this frequency response. Consider the interactive graph below. On that graph you can see the low frequency asymptote, the high frequency asymptote and the point where the gain is .707 of the low frequency gain. Check the intersection of the two lines.

The intersection of the two lines occurs where [ = 1/X. For obvious reasons, this intersection is called the corner frequency.

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

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

y

Any transfer function is a ratio of polynomials - and those polynomials have real coefficients.

y

y

y

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

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

y

y

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

Decibels And More
When we introduced Bode' plots, we noted that the vertical scale of a Bode' plot is often in terms of decibels. It's time you got acquainted with decibels if you haven't heard of them before. Here's a start. Originally, decibels were used to measure power gains. If a system had an output power, Po, and an input power, Pi, then the ratio of output power to input power - the power gain - is: Po/Pi The decibel gain is proportional to the logarithm - to the base ten (10) - of the power gain The gain can be expressed as the logarithm - to the base ten (10) - of the power gain Gain = log10(Po/Pi) When expressed this way, the units are bels. A decibel is one tenth of a bel, so the gain expressed in decibels is:

Gaindb = 10 log10(Po/Pi) The unit bel is something of a story in itself. Alexander Graham Bell did a lot of work with the deaf, and he was recognized for his work with an honorary doctorate in 1880 by Gaulladet College in Washington, D.C.(and he also delivered the commencement address) He is more famous for his founding of the National Geographic Society, and other work he did. Alexander Graham Bell was also honored by having a unit named in his honor - the bel. Today, the decibel is a commonly used unit to measure sound intensity and it is well known that high decibel levels contribute to deafness - a very ironic closing of the circle. Today, power is not so much an issue. We're more interested in voltage gain of an amplifier. There's an interesting transition from power to voltage that will help us understand how gain - expressed in decibels - is viewed today. In an amplifier, if the amplifier has an input resistance R1, then the power input to the amplifier is given by: Power In = V12/R1 Similarly, the output power into a resistor Ro is given by: Power Out = Vo2/Ro Now, look at the ratio of output power to input power: Power Out/Power In = (Vo2/Ro)/(Vi2/Ri) Now, compute the decibel gain: Gaindb = 10 log10(Po/Pi) = 10 log10(Po/Pi) = 10 log10(Vo2/Ro)/(Vi2/Ri) = 10 log10(Vo2/Vi2) + 10 log10(Ri/Ro)

= 20 log10(Vo/Vi) + 10 log10(Ri/Ro) The final result has a term in it that depends upon the resistors. Gaindb = 20 log10(Vo/Vi) + 10 log10(Ri/Ro) Today, engineers are often more concerned with things like voltage gain. The resistances and power involved are not a concern at all when analyzing control systems, so the resistance term is ignored, and we take the gain, in db, of a system to be: Gaindb = 20 log10(Vo/Vi) We should realize that we can plot gain, in db, for a system as a function of frequency. The ratio of output voltage to input voltage is simply the ratio of output amplitude to input amplitude at some frequency - our old friend, frequency response. OK! You know about decibels. But there's some other things you need to know about Bode' plots. The vertical axis on a true Bode' plot is scaled in db. The horizontal axis is scaled using a logarithmic frequency scale. Here's some not-soobvious facts about the frequency scale. An increase of frequency by a factor of 10 is referred to as a decade. That's a fairly obvious reference. Ten years is a decade when speaking of time. Our currency is based on a decimal system because it's based on factors of 10. An increase of frequency by a factor of 2 is referred to as an octave. We're getting into the Latin and Greek roots here. Decade is based on a Latin root referring to the number 10. Octave is based on a classical root referring to the number two - or is it? Right or wrong? Wrong! Octave refers to eight, not two. The reason a doubling of frequency is called an octave is that the musical world defined the term far earlier than we ever thought of it. An octave is a doubling of frequency, but it's eight notes in the scale to go up an octave. Ok, now we're going to put this all together. Here's a Bode' plot for a first order system. It has a DC gain of 20db, and a corner frequency near f = 80 Hz. Now, look at the slope of the high frequency portion of the plot.

y y y

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

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

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

If we go back to the transfer function for a first order system, we can reexamine the high frequency behavior. Here's the transfer function. G(j[) = 1/(j[X+ 1)
y

If [ is large (and only if it is large!), then the imaginary term in the denominator dominates, and we have: G(j[) ~= 1/j[X |G(j[)| ~= 1/[X

log(|G(j[)|) = log(1/[X) = -log([X) = -log([) - log(X)

Express things in terms of decibels. log(|G(j[)|) = -log([) - log(X)

Gaindb = 20 log10(|G(j[)|) = -20 log([) - 20 log(X) Now, if we start with some frequency, [o, we can calculate the gain at the frequency. Gaindb([o) = -20 log([o) - 20 log(X) Now, take a frequency one decade higher, at 10[o. Gaindb(10[o) = -20 log(10[o) - 20 log(X) We can calculate the difference in the db gain at these two frequencies. Gaindb(10[o) - Gaindb([o) = [-20 log(10[o) - 20 log(X)] - [-20 log(10[o) - 20 log(X)] The difference is: Gaindb(10[o) - Gaindb([o) = -20 log(10[o) + 20 log([o) = -20 log(10) - 20 log([o) + 20 log([o) = -20 log(10) = -20 db - in one decade! Reflecting on the derivation above, we realize that this derivation says that the slope is -20 db/decade for the high frequency asymptote in the Bode' plot. It's also possible to express that another way. If we consider two frequencies that are an octave apart, we can see that the slope can also be said to be 6db/octave. The difference in the frequency response between the two frequencies is: Gaindb(2wo) - Gaindb(wo) = -20 log(2[o) + 20 log([o) = - 20 log(2) - 20 log([o) + 20 log([o) = -20 log(2) = -6.0206 db - in one decade - and it's usually just rounded to 6db/octave. It's time to leave this topic. However consider this. We've only looked at one first order system. Higher order systems - even second order systems - are bound to have some differences in their Bode' plot behavior. High frequency asymptotes

will drop off at different slopes, for example, although we'll find that they drop off at integral multiples of -20db/decade or -6 db/octave. There are lots of interesting things you need to know, and you can start looking at second order systems now.

Bode' Plots For 2nd Order Systems
We've looked at first order systems. Remember our general goal:
y

Given a Transfer Function: o Be able to plot the Bode' plot, manually or with a math analysis program. Know that the Bode' plot you generated "makes sense".

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

y

y

A little reflection will probably tell you some things. o For example, this system could have two complex roots. It's not obvious, but to have two complex roots, the only thing necessary is that the damping ratio, ^, be less than one.

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

^- the damping ratio = 0.1 [n- the undamped natural frequency = 1000. Gdc- the DC gain of the system = 1.0.

This system also has at least one unexpected feature - the "hump" in the frequency response between f = 100 and f = 200 - a resonant peak. It's important to understand how that peak in the frequency response comes about. Let's look at the transfer function of a second order system. Here's a general form for such a system. Examine how that system behaves for different frequencies.

y y y

Substitute s = j[, to get the frequency response. For small [, the gain is just Gdc. For large [, the gain is Gdc/[2. That means that the high frequency gain drops off at -40 db/decade.

y

There are intermediate frequencies where interesting things happen!

We will start by looking at the interesting things that happen at the intermediate frequencies. Here's the transfer function again, with s replaced now by j[.
y

We will examine what happens when [ = [n.

y

At the natural frequency, the (j[)2 term becomes -[n2, cancelling out the last term in the denominator, the [n2 term, since j2 = -1.

y

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

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

y

y

The only term that involves the damping ratio is the one left in the denominator when [ = [n. The damping ratio is in the denominator, so the smaller the damping ratio, the larger the frequency response is going to be. At [ = [n, the magnitude of the frequency response function is:

y

or G(j[n) =Gdc/j2^

The formula for the gain of the frequency response at [ = [n is interesting because:
y

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

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

For small [, the gain is just Gdc. For large [, the gain Gdc/[2. For small [, the gain is just Gdc, assuming Gdc = 10 (or 20 db) on the plot. For large [, the gain is Gdc/[2, - dropping off at -40 db/decade.

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

And, we can insert the point at the resonant frequency, using our formula. o G(j[n) =Gdc/j2^

For this example, we'll assume ^ = 0.1. Remember: Gdc= 10, and ^ = 0.1, so this works out to be a gain of 50 at the resonant peak, the equivalent of 34 db. Do we have a problem here?
y y y

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

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

Gdc= 10, ^ = 0.1, [n = 2 T 20, (since that natural frequency was 20 Hz.)

With these paramters, note the following in the plot.
y y y

y

The DC gain is 20 db which corresponds to a gain of 10. The resonant peak is pretty much right at 20 Hz as it should be. The resonant peak is about 13 or 14 db high. o A gain of 50 would be 14 db, do that also checks! The high frequency slope looks to be around -12 db/octave or -20 db/decade.

All of these observations confirm the calculations, and they really point out that it can be important to understand how the resonant peak depends upon the damping ratio. To make that correspondence between resonant peak and damping ratio as clear as possible, we have here an example of a frequency response for another system. We'll let you control the damping ratio, but we're going to set the DC gain and the natural frequency. Hopefully, you'll see how this peak depends upon the system's damping ratio. Use the right and left arrow controls to step the movie a single step forward or backward.
y y y

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

y y y

There is a resonant peak in the second order system response. The size of the resonant peak depends upon the damping ratio. For damping ratios less than about 0.5 the peak is relatively insignificant.

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

A damping ratio of 0.1 An undamped natural frequency of 159 Hz. (1000 rad/sec.)

Notice the following for this plot.
y y

The phase starts at zero degrees for low frequencies. The phase asymptotically approaches -180o for high frequencies.

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

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

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

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

We will start with a system that is not all that large - a second order system with two real poles. Just for discussion, we'll use the system with the transfer function shown below.

y

y y

If we wanted to sketch this Bode' plot we could start by looking at the DC gain. Remember that the DC gain is just G(jw) with [ = 0. Letting [ = 0 in G(j[), we get: o G(0) = 10. (20db) At low frequencies, the (.002s + 1) term in the denominator will still look pretty much like 1.0. However, as we go up in frequency, the (.01s + 1) term will have an effect. The (.01s + 1) term introduces a corner frequency which we discussed earlier in the section on Bode' plots for first order systems. The corner frequency is at: o f = 100/2T = 15.9Hz.

y

y y

y

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

y

y

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

y

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

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

Start by looking at the DC gain - as before.
y y

y y y y y

y

y

y

Remember that the DC gain is just G(jw) with w = 0. Letting w = 0 in G(j[), we get: o G(0) = 10. (20db) As we go up in frequency from DC, the (.01s + 1) term will have an effect. The (.01s + 1) term introduces a corner frequency - as before. The corner frequency is at f = 100/2p = 15.9Hz. Check the slope. It should be -20 db/decade. At slightly higher frequencies, the (.002s + 1) term will start to have an effect. The (.002s + 1) term will add another -20db/decade - or wait a minute - is that +20 db/decade? Because it is a zero, it is +20db/dec, and the corner frequency is at: o f = 500/2T = 79.5Hz. For frequencies above 79.5 Hz, the gain would be 10*.002/.01 = 2 or 6db.

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

y

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

Sponsor Documents

Or use your account on DocShare.tips

Hide

Forgot your password?

Or register your new account on DocShare.tips

Hide

Lost your password? Please enter your email address. You will receive a link to create a new password.

Back to log-in

Close