You are on page 1of 7

4.

Magnetic Circuits

We will begin with a look at magnetic circuits, due to the fact that most major components of a power system involve the manipulation and utilization of magnetic fields along prescribed paths. For example, as we will see in the next section, a generators output voltage depends heavily upon the relative velocity between a conductor and a magnetic field. It also depends on the magnetic field strength. Because of this, we wish to be able to design generators such that we can dictate (to some degree) the strength and orientation of the magnetic field within the device. If the field is zero, no voltage is generated; however, if we can make the field strong enough, and orient it in the right way, we can greatly enhance the voltage supplied by a generator. This calls for the effective design of the magnetic circuit within the generator. The same reasoning holds for electric motors, transformers, and other devices such as relays and solenoids, the operation of which depends heavily on the effective design of the magnetic circuit. In the Fundamentals text we were introduced to Gauss Law for magnetic fields, which herein is restated in differential form as

B = 0

(4.1)

The property of no divergence of a magnetic field assumes, of course, the non-existence of magnetic monopoles. Until such time (if ever) that magnetic monopoles are discovered, this relationship tells us essentially that there are no sources or sinks of a magnetic flux line. This is in contrast to an electric field, in which a positive charge such as a proton can be seen as a source and a negative charge such as an electron can be seen as a sink. With magnetic fields, if one chooses a particular flux line and follows it from an arbitrary starting point, one will eventually return to that same starting point, assuming the field is static. That is, one would travel in a closed circuit. This may remind us of what we were taught of electrical current in, say, a DC circuit: any current leaving a DC source will eventually return to that source. Speaking of DC sources of current, we can draw an analogy with magnetic fields. If we consider a single-turn closed loop as shown in Figure 4-1 we can see that the magnetic field generated by the current in the loop emanates from the loop, eventually returning. This current-carrying loop, then, can be seen as a source of a magnetic field. If two turns were included, making the single loop a coil or winding instead, the magnetic field strength would double (assuming the external medium were a linear material such as air, and that the second turn was tightly packed with the first). Three turns would triple the field strength, and so on. As we learned earlier, the magnetic flux density B is proportional to the field strength H by the relation
B = H

(4.2)

where = r0 is the permeability of the medium. When the medium is air, r 1 and = 0 = 4 10-7 H/m. For the situation of Figure 4-1, the flux density will increase linearly with the field intensity.

Figure 4-1 Magnetic flux pattern produced by a current-carrying loop.

Note that in general this is not the case. For media such as iron or other ferrous materials, is actually a function of the field intensity, implying a nonlinear relationship between B and H . An example of this is shown in the DC magnetization curve of Figure 4-2(a). Only with relatively low field intensities can the material be considered approximately linear; that is, with an approximately constant permeability. The relative permeability of this sample material as a function of the field intensities are shown in Figure 4-2(b).
1.8 1.6 1.4

Induction, T

1.2 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40 50 60 70

Magnetizing Intensity, A-t/m

(a)
40000 35000 30000

Rel. Perm.

25000 20000 15000 10000 5000 0 0 10 20 30 40 50 60 70

Magnetizing Intensity, A-t/m

(b)

Figure 4-2 (a) DC Magnetization curve for 0.011 M4 electrical steel, and (b) Relative permeability for the same steel (Allegheny Ludlum Corp.)

Limiting the remainder of our discussion in this section to the special assumption of linear magnetic materials (or operation in the linear portion of the magnetization curve), let us take another look at the simple relation between B and H of Equation 4.2. Note the similarity of this expression to the relationship between current density J and electric field strength E , originally seen in Equation 3.1 of the Fundamentals text:

J = E

(4.3)

This expression was previously recognized as being another statement of the familiar Ohms Law, and essentially described a current density resulting from an electric field intensity applied in a medium with conductivity . Integrating the current density over the appropriate cross sectional area would yield the current passing through a conductor. For example, for the current density depicted in Figure 4-3(a),

I = J(x, y) dS u Z
S

(4.4)

In a similar fashion, a magnetic flux density arising from a magnetic field intensity (or magnetizing force) in a permeable material can be integrated over the appropriate cross sectional area to yield total flux passing through the material. This is illustrated in Figure 4-3(b) and the following relation:

= B(x, y) dS u Z
S

(4.5)

Using this analogy, it can be seen that magnetic permeability can be considered to be a magnetic conductivity. In fact, a magnetic circuit may be thought of as a closed path or periphery through which magnetic flux is established. Magnetic flux is therefore analogous to current in an electrical circuit, although flux is not the flow of particles like current is. In spite of the fact that flux literally means to flow, there is no flow; magnetic flux is simply established, and is visualized as a set of lines that are continuous. Again, these lines form closed loops.

(a)

(b) Figure 4-3 (a) Current flowing through a good electrical conductor. (b) Magnetic flux established in a permeable magnetic material.

Now, if electrical conductivity has a magnetic circuit analog in permeability , and electrical current I has an analog of magnetic flux , what can be said to be analogous to voltage and resistance in an electrical circuit? Well, we have already defined a current-carrying coil or winding as a source of magnetic flux. This is the analogy for voltage or electromotive force, and it is designated as magnetomotive force, or mmf. It is usually abbreviated F, and has the unit of Ampere-turns. That is, the strength of the mmf is literally the product of the current in the coil and the number of turns or loops. As stated previously, when the number of turns is, say, doubled, the magnetic field strength H is also doubled. In a linear material, this means that the flux is also doubled. This is the same thing that happens in a DC electrical circuit: doubling the source voltage will double the current in the circuit through the familiar property of linearity. One final electrical circuit quantity for which we will provide an analogy is resistance. In a magnetic circuit, the quantity which limits the establishment of flux is called reluctance, and we will abbreviate it as . Given a length l of magnetic material with a permeability and constant cross-sectional area A, as shown in Figure 4-4, reluctance is determined to be = l A (4.6)

One very interesting observation is that the DC resistance of an electrical conductor of length l and crosssectional area A is given as
RDC = l A

(4.7)

This similarity reinforces the notion that magnetic permeability can be considered a magnetic conductivity in the sense that a high permeability (low reluctance) would result in greater flux. In the case of an electrical circuit, a high conductivity (low resistance) would result in greater current. Use of the electrical circuit analogy is very beneficial in the analysis of magnetic circuits. Magnetic circuits are, again, paths through which magnetic flux is established. Use of a high-permeability material in a magnetic circuit acts to confine the flux primarily to a designated path, much the way a highly conductive material (like copper) confines electrical current to a particular path. Of particular interest and usefulness is the fact that magnetic circuits, once the relevant quantities of mmf and reluctance are determined, can be analyzed in the same manner as DC electrical circuits. For example, consider the magnetic circuit of Figure 4-4. A single 100-turn coil, carrying 1-A of electrical current, is wrapped about a leg of a core with relative magnetic permeability r = 1000. The core is cut in a certain location, yielding an air gap of length lgap = 1 mm. The thickness of the core is 1 cm, yielding a core cross sectional area of 1 cm 2. It is also seen that the mean length of the core lcore 2(2.5 cm) = 5 cm (neglecting the 1 mm due to the air gap). From the right hand rule, the direction of magnetic flux is clockwise in the core. The reluctance of the core can be determined easily as:
core = 5 cm 0.157 m = = 1.25 106 H -1 2 7 2 1000 0 (1 cm ) 1000(4 10 H/m)(0.0001 m )

Similary,
gap = 1 mm 0.001 m 10 = = H -1 = 7.96 106 H -1 2 7 2 7 0 (1 cm ) (4 10 H/m)(0.0001 m ) 4 10

Figure 4-4 A simple magnetic circuit.

Notice how the reluctance (magnetic circuit resistance) of the gap is greater than the reluctance of the rest of the circuit, even though the length of the gap is only about 0.6% of the length of the core. In practice, small air gaps can be used in certain applications, such as transformers, to help control the flux and flux density in the core material. Now we can draw the equivalent representation of the magnetic circuit as shown in Figure 4-5.

Figure 4-5 Equivalent circuit representation of the magnetic circuit of Figure 4-5.

The flux in the core and air gap can be found from Ohms Law: = 100 A t = 10.86 10 6 Wb 6 -1 (1.25 + 7.96) 10 H

Given the constant core cross-sectional area, and ignoring fringing of the field at the air gap, we determine the flux density in the circuit to be: B= 10.86 10-6 Wb = 0.1086 Wb/m 2 = 0.1086 T 2 0.0001 m

Since there is but one path through which the flux may be established, the flux (like current in a series electrical circuit) is constant throughout. Having made the assumption that fringing in the air gap may be neglected, flux density is also the same in this example. However, there will be some fringing effect in the air gap, which would effectively increase the cross-sectional area of the gap. This will mean a lower gap reluctance, and actually a slightly higher flux in the circuit than previously calculated. Now, recalling from Chapter 3 of the Fundamentals text, the self inductance L of a winding (a macroscopic circuit parameter) is defined as the flux linkage of the winding divided by the current I through the winding. That is,
L= I

(4.8)

where can be approximated (for a tightly packed winding about a highly permeable material), as:

(4.9)

That is, if the flux is considered precisely the same through each loop of the winding, then the flux linkage is simply the flux through a single loop multiplied by the number of loops. Note that in general this will not be the case, and flux linkage will actually be the sum of each loops individual flux value, that is:
= i
i =1 N

(4.10)

In our example, the flux linkage would be N = 100 (10.86 10-6 Wb) = 10.86 10-4 Wb , yielding an inductance of L= 10.86 10-4 Wb = 1.086 mH. 1A

As a point of interest, in this example it is seen that the energy stored in the inductors magnetic field is, by circuit theory,
2 WL = 1 LI2 = 1 1.086 103 H (1 A ) = 0.543 mJ. 2 2

Lets see if we can calculate the same value for inductor energy using EM field theory. First, we must recall that energy density in a magnetic field is given as: wL = 1 H 2 . 2 (4.11)

(As an aside, notice the very strong similarity between Eq. 4.11 and the equation for the energy in an inductor.) Now, the core of the inductor in our example will have a different energy density than the gap. We might initially expect the total energy of the magnetic field to consist to a greater extent of the core component. To see whether or not this is true, we will determine the core and gap energy densities and energies separately. For the core,

B B2 wL, core = 1 core H = 1 core core = 1 core 2 2 2 core core


2

( 0.1086 T ) 2 1 = 2 1000 4 10 7 H/m

= 4.69 J/m 3.

Furthermore, for the gap,


2 2 1 Bgap = 1 ( 0.1086 T ) = 4,690 J/m 3 . wL, gap = 2 2 4 10 7 H/m gap

Note the far greater energy density in the gap. However, to determine energies in the core and gap, we will need to introduce the respective volumes. In the core,
Vcore = A core lcore = 0.0001 m 2 ( 0.157 m ) = 1.571 105 m3 .

The volume of the gap simply:

Vgap = A gap lgap = 0.0001 m 2 ( 0.001 m ) = 1 107 m3 .


So, although the energy density in the gap is much greater than that of the core, the gap volume is far less. The total energy in the inductors magnetic field is therefore: WL = 4.69 J/m3 1.571 105 m3 + 4,690 J/m3 1 107 m3 = 0.074 mJ + 0.469 mJ = 0.543 mJ

)(

) (

)(

Even there was found to be much greater energy associated with the gap, contrary perhaps to initial expectation, the overall result is exactly the same as the previous result, and helps to illustrate the following point: circuit theory and EM theory, being the prevailing theories upon which the field of electrical engineering is built, offer two distinct methods of looking at the same thing. Circuit theory is a more macroscopic or high level method of examining electrical behavior of components and systems, and is less concerned with the more fundamental, microscopic viewpoint involving field quantities. Both theories have their place and, if properly applied, should not contradict. Part of being a good engineer is knowing when to employ which theory.

You might also like