103
Fleming's left hand rule for motors Fleming's left hand rule Alternate representation of Fleming's LHR Fleming's left hand rule (for electric motors ) shows the direction of the thrust on a conductor carrying a current in a magnetic field . The left hand is held with the thumb , index finger and middle finger mutually at right angles . It can be recalled by remembering that "motors drive on the left, in Britain anyway." The First finger represents the direction of the magnetic Field. (north to south) The Second finger represents the direction of the Current (the direction of the current is the direction of conventional current; from positive to negative). The Thumb represents the direction of the Thrust or resultant Motion.

Basic Laws and Rules in Electrical Power

Embed Size (px)

Citation preview

Page 1: Basic Laws and Rules in Electrical Power

Fleming's left hand rule for motors

Fleming's left hand rule

Alternate representation of Fleming's LHR

Fleming's left hand rule (for electric motors) shows the direction of the thrust on a conductor

carrying a current in a magnetic field.

The left hand is held with the thumb, index finger and middle finger mutually at right angles. It

can be recalled by remembering that "motors drive on the left, in Britain anyway."

The First finger represents the direction of the magnetic Field. (north to south)

The Second finger represents the direction of the Current (the direction of the current is

the direction of conventional current; from positive to negative).

The Thumb represents the direction of the Thrust or resultant Motion.

Page 2: Basic Laws and Rules in Electrical Power

This can also be remembered using "FBI" and moving from thumb to second finger.

The thumb is the force F.

The first finger is the magnetic field B.

The second finger is that of current I.

There also exists Fleming's right hand rule (for generators). The appropriately-handed rule can be

recalled by remembering that the letter "g" is in "right" and "generator".

Both mnemonics are named after British professional engineer John Ambrose Fleming who

invented them.

Other mnemonics also exist that use a left hand rule or a right hand rule for predicting resulting

motion from a pre-existing current and field.

De Graaf's translation of Fleming's left-hand rule - which also uses thrust, field and current - and

the right-hand rule, is the FBI rule. The FBI rule changes Thrust into F (Lorentz force), B

(direction of the magnetic field) and I (current). The FBI rule is easily remembered by US

citizens because of the commonly known abbreviation for the Federal Bureau of Investigation.

This works for both dynamos and motors. If electrons flow forwards (away from us) in a wire

held horizontally between a south pole on its left and a north pole on its right, the electrons will

try to move to their left, through the lines of force ("forest") running across from the north pole.

This will pull the wire downwards.

If the wire itself is pulled downwards the electrons in it will be moving down through the lines of

force and will wish to move to their left, ie towards us (the opposite direction from the previous

one) which is exactly as it should be.

Hold a north pole at the left side of a cathode ray tube tv, in which electrons are rushing toward

the screen. The lines of force run, roughly, from left to right across the tube. The electrons will

turn to their left in this "forest", ie downwards, and therefore the image will move down. A south

pole will make the image rise.

Fleming's right hand rule

Page 3: Basic Laws and Rules in Electrical Power

Fleming's right hand rule

Fleming's right hand rule (for generators) shows the direction of induced current when a

conductor moves in a magnetic field.

The right hand is held with the thumb, first finger and second finger mutually perpendicular to

each other (at right angles), as shown in the diagram .

The Thumb represents the direction of Motion of the conductor.

The First finger represents the direction of the Field. (north to south)

The Second finger represents the direction of the induced or generated Current (the

direction of the induced current will be the direction of conventional current; from

positive to negative).

Faraday's law of induction

History

Electromagnetic induction was discovered independently by Michael Faraday and Joseph Henry

in 1831; however, Faraday was the first to publish the results of his experiments.[2][3]

Faraday's disk

Page 4: Basic Laws and Rules in Electrical Power

In Faraday's first experimental demonstration of electromagnetic induction (August 1831), he

wrapped two wires around opposite sides of an iron torus (an arrangement similar to a modern

transformer). Based on his assessment of recently-discovered properties of electromagnets, he

expected that when current started to flow in one wire, a sort of wave would travel through the

ring and cause some electrical effect on the opposite side. He plugged one wire into a

galvanometer, and watched it as he connected the other wire to a battery. Indeed, he saw a

transient current (which he called a "wave of electricity") when he connected the wire to the

battery, and another when he disconnected it.[4]

Within two months, Faraday had found several

other manifestations of electromagnetic induction. For example, he saw transient currents when

he quickly slid a bar magnet in and out of a coil of wires, and he generated a steady (DC) current

by rotating a copper disk near a bar magnet with a sliding electrical lead ("Faraday's disk").[5]

Faraday explained electromagnetic induction using a concept he called lines of force. However,

scientists at the time widely rejected his theoretical ideas, mainly because they were not

formulated mathematically.[6]

An exception was Maxwell, who used Faraday's ideas as the basis

of his quantitative electromagnetic theory.[6][7][8]

In Maxwell's papers, the time varying aspect of

electromagnetic induction is expressed as a differential equation which Oliver Heaviside referred

to as Faraday's law even though it is slightly different in form from the original version of

Faraday's law, and doesn't cater for motionally induced EMF. Heaviside's version is the form

recognized today in the group of equations known as Maxwell's equations.

Lenz's law, formulated by Heinrich Lenz in 1834, describes "flux through the circuit", and gives

the direction of the induced electromotive force and current resulting from electromagnetic

induction (elaborated upon in the examples below).

Faraday's experiment showing induction between coils of wire: The liquid battery (right)

provides a current which flows through the small coil (A), creating a magnetic field. When the

coils are stationary, no current is induced. But when the small coil is moved in or out of the large

coil (B), the magnetic flux through the large coil changes, inducing a current which is detected

by the galvanometer (G).[9]

Faraday's law as two different phenomena

Page 5: Basic Laws and Rules in Electrical Power

Some physicists have remarked that Faraday's law is a single equation describing two different

phenomena: The motional EMF generated by a magnetic force on a moving wire, and the

transformer EMF generated by an electric force due to a changing magnetic field. James Clerk

Maxwell drew attention to this fact in his 1861 paper On Physical Lines of Force. In the latter

half of part II of that paper, Maxwell gives a separate physical explanation for each of the two

phenomena. A reference to these two aspects of electromagnetic induction is made in some

modern textbooks.[10]

As Richard Feynman states:[11]

So the "flux rule" that the emf in a circuit is equal to the rate of change of the magnetic flux

through the circuit applies whether the flux changes because the field changes or because the

circuit moves (or both).... Yet in our explanation of the rule we have used two completely

distinct laws for the two cases – for "circuit moves" and for "field

changes".

We know of no other place in physics where such a simple and accurate general principle

requires for its real understanding an analysis in terms of two different phenomena.

– Richard P. Feynman, The Feynman Lectures on Physics

Reflection on this apparent dichotomy was one of the principal paths that led Einstein to develop

special relativity:

It is known that Maxwell’s electrodynamics—as usually understood at the present time—when

applied to moving bodies, leads to asymmetries which do not appear to be inherent in the

phenomena. Take, for example, the reciprocal electrodynamic action of a magnet and a

conductor. The observable phenomenon here depends only on the relative motion of the

conductor and the magnet, whereas the customary view draws a sharp distinction between the

two cases in which either the one or the other of these bodies is in motion. For if the magnet is in

motion and the conductor at rest, there arises in the neighbourhood of the magnet an electric field

with a certain definite energy, producing a current at the places where parts of the conductor are

situated. But if the magnet is stationary and the conductor in motion, no electric field arises in

the neighbourhood of the magnet. In the conductor, however, we find an electromotive force, to

which in itself there is no corresponding energy, but which gives rise—assuming equality of

relative motion in the two cases discussed—to electric currents of the same path and intensity as

those produced by the electric forces in the former case.

– Albert Einstein, On the Electrodynamics of Moving Bodies[12]

Flux through a surface and EMF around a loop

Page 6: Basic Laws and Rules in Electrical Power

The definition of surface integral relies on splitting the surface Σ into small surface elements.

Each element is associated with a vector dA of magnitude equal to the area of the element and

with direction normal to the element and pointing outward.

A vector field F(r, t) defined throughout space, and a surface Σ bounded by curve ∂Σ moving

with velocity v over which the field is integrated.

Faraday's law of induction makes use of the magnetic flux ΦB through a surface Σ, defined by an

integral over a surface:

where dA is an element of surface area of the moving surface Σ(t), B is the magnetic field, and

B·dA is a vector dot product. The surface is considered to have a "mouth" outlined by a closed

curve denoted ∂Σ(t). When the flux changes, Faraday's law of induction says that the work

done (per unit charge) moving a test charge around the closed curve ∂Σ(t), called the

electromotive force (EMF), is given by:

where is the magnitude of the electromotive force (EMF) in volts and ΦB is the magnetic flux

in webers. The direction of the electromotive force is given by Lenz's law.

For a tightly-wound coil of wire, composed of N identical loops, each with the same ΦB,

Faraday's law of induction states that

Page 7: Basic Laws and Rules in Electrical Power

where N is the number of turns of wire and ΦB is the magnetic flux in webers through a single

loop.

In choosing a path ∂Σ(t) to find EMF, the path must satisfy the basic requirements that (i) it is a

closed path, and (ii) the path must capture the relative motion of the parts of the circuit (the

origin of the t-dependence in ∂Σ(t) ). It is not a requirement that the path follow a line of current

flow, but of course the EMF that is found using the flux law will be the EMF around the chosen

path. If a current path is not followed, the EMF might not be the EMF driving the current.

Example: Spatially varying Magnetic field

Figure 3: Closed rectangular wire loop moving along x-axis at velocity v in magnetic field B that

varies with position x.

Consider the case in Figure 3 of a closed rectangular loop of wire in the xy-plane translated in the

x-direction at velocity v. Thus, the center of the loop at xC satisfies v = dxC / dt. The loop has

length ℓ in the y-direction and width w in the x-direction. A time-independent but spatially

varying magnetic field B(x) points in the z-direction. The magnetic field on the left side is B( xC

− w / 2), and on the right side is B( xC + w / 2). The electromotive force is to be found by using

either the Lorentz force law or equivalently by using Faraday's induction law above.

Lorentz force law method

A charge q in the wire on the left side of the loop experiences a Lorentz force q v × B k = −q v

B(xC − w / 2) j   ( j, k unit vectors in the y- and z-directions; see vector cross product), leading to

an EMF (work per unit charge) of v ℓ B(xC − w / 2) along the length of the left side of the loop.

On the right side of the loop the same argument shows the EMF to be v ℓ B(xC + w / 2). The two

EMF's oppose each other, both pushing positive charge toward the bottom of the loop. In the

Page 8: Basic Laws and Rules in Electrical Power

case where the B-field increases with increase in x, the force on the right side is largest, and the

current will be clockwise: using the right-hand rule, the B-field generated by the current opposes

the impressed field.[13]

The EMF driving the current must increase as we move counterclockwise

(opposite to the current). Adding the EMF's in a counterclockwise tour of the loop we find

Faraday's law method

At any position of the loop the magnetic flux through the loop is

The sign choice is decided by whether the normal to the surface points in the same direction as

B, or in the opposite direction. If we take the normal to the surface as pointing in the same

direction as the B-field of the induced current, this sign is negative. The time derivative of the

flux is then (using the chain rule of differentiation or the general form of Leibniz rule for

differentiation of an integral):

(where v = dxC / dt is the rate of motion of the loop in the x-direction ) leading to:

as before.

The equivalence of these two approaches is general and, depending on the example, one or the

other method may prove more practical.

Example: Moving loop in uniform Magnetic field

Page 9: Basic Laws and Rules in Electrical Power

Figure 4: Rectangular wire loop rotating at angular velocity ω in radially outward pointing

magnetic field B of fixed magnitude. Current is collected by brushes attached to top and bottom

discs, which have conducting rims.

Figure 4 shows a spindle formed of two discs with conducting rims and a conducting loop

attached vertically between these rims. The entire assembly spins in a magnetic field that points

radially outward, but is the same magnitude regardless of its direction. A radially oriented

collecting return loop picks up current from the conducting rims. At the location of the collecting

return loop, the radial B-field lies in the plane of the collecting loop, so the collecting loop

contributes no flux to the circuit. The electromotive force is to be found directly and by using

Faraday's law above.

Lorentz force law method

In this case the Lorentz force drives the current in the two vertical arms of the moving loop

downward, so current flows from the top disc to the bottom disc. In the conducting rims of the

discs, the Lorentz force is perpendicular to the rim, so no EMF is generated in the rims, nor in

the horizontal portions of the moving loop. Current is transmitted from the bottom rim to the top

rim through the external return loop, which is oriented so the B-field is in its plane. Thus, the

Lorentz force in the return loop is perpendicular to the loop, and no EMF is generated in this

return loop. Traversing the current path in the direction opposite to the current flow, work is

done against the Lorentz force only in the vertical arms of the moving loop, where

where v= velocity of moving charge [14]

Consequently, the EMF is

where v = velocity of conductor or magnet [14]

and l = vertical length of the loop. In this case the

velocity is related to the angular rate of rotation by v = r ω, with r = radius of cylinder. Notice

that the same work is done on any path that rotates with the loop and connects the upper and

lower rim.

Page 10: Basic Laws and Rules in Electrical Power

Faraday's law method

An intuitively appealing but mistaken approach to using the flux rule would say the flux through

the circuit was just ΦB = B w ℓ, where w = width of the moving loop. This number is time-

independent, so the approach predicts incorrectly that no EMF is generated. The flaw in this

argument is that it fails to consider the entire current path, which is a closed loop.

To use the flux rule, we have to look at the entire current path, which includes the path through

the rims in the top and bottom discs. We can choose an arbitrary closed path through the rims

and the rotating loop, and the flux law will find the EMF around the chosen path. Any path that

has a segment attached to the rotating loop captures the relative motion of the parts of the circuit.

As an example path, let's traverse the circuit in the direction of rotation in the top disc, and in the

direction opposite to the direction of rotation in the bottom disc (shown by arrows in Figure 4).

In this case, for the moving loop at an angle θ from the collecting loop, a portion of the cylinder

of area A = r ℓ θ is part of the circuit. This area is perpendicular to the B-field, and so contributes

to the flux an amount:

where the sign is negative because the right-hand rule suggests the B-field generated by the

current loop is opposite in direction to the applied B field. As this is the only time-dependent

portion of the flux, the flux law predicts an EMF of

in agreement with the Lorentz force law calculation.

Now let's try a different path. Follow a path traversing the rims via the opposite choice of

segments. Then the coupled flux would decrease as θ increased, but the right-hand rule would

suggest the current loop added to the applied B-field, so the EMF around this path is the same as

for the first path. Any mixture of return paths leads to the same result for EMF, so it is actually

immaterial which path is followed.

Direct evaluation of the change in flux

Page 11: Basic Laws and Rules in Electrical Power

Figure 5: A simplified version of Figure 4. The loop slides with velocity v in a stationary,

homogeneous B-field.

The use of a closed path to find EMF as done above appears to depend upon details of the path

geometry. In contrast, the Lorentz-law approach is independent of such restrictions. The

following discussion is intended to provide a better understanding of the equivalence of paths

and escape the particulars of path selection when using the flux law.

Figure 5 is an idealization of Figure 4 with the cylinder unwrapped onto a plane. The same path-

related analysis works, but a simplification is suggested. The time-independent aspects of the

circuit cannot affect the time-rate-of-change of flux. For example, at a constant velocity of

sliding the loop, the details of current flow through the loop are not time dependent. Instead of

concern over details of the closed loop selected to find the EMF, one can focus on the area of B-

field swept out by the moving loop. This suggestion amounts to finding the rate at which flux is

cut by the circuit.[15]

That notion provides direct evaluation of the rate of change of flux, without

concern over the time-independent details of various path choices around the circuit. Just as with

the Lorentz law approach, it is clear that any two paths attached to the sliding loop, but differing

in how they cross the loop, produce the same rate-of-change of flux.

In Figure 5 the area swept out in unit time is simply dA / dt = v ℓ, regardless of the details of the

selected closed path, so Faraday's law of induction provides the EMF as:[16]

This path independence of EMF shows that if the sliding loop is replaced by a solid conducting

plate, or even some complex warped surface, the analysis is the same: find the flux in the area

swept out by the moving portion of the circuit. In a similar way, if the sliding loop in the drum

generator of Figure 4 is replaced by a 360° solid conducting cylinder, the swept area calculation

is exactly the same as for the case with only a loop. That is, the EMF predicted by Faraday's law

is exactly the same for the case with a cylinder with solid conducting walls or, for that matter, a

cylinder with a cheese grater for walls. Notice, though, that the current that flows as a result of

this EMF will not be the same because the resistance of the circuit determines the current.

The Maxwell-Faraday equation

Page 12: Basic Laws and Rules in Electrical Power

Figure 6: An illustration of Kelvin-Stokes theorem with surface Σ its boundary ∂Σ and

orientation n set by the right-hand rule.

A changing magnetic field creates an electric field; this phenomenon is described by the

Maxwell-Faraday equation:[17]

where:

denotes curl

E is the electric field

B is the magnetic field

This equation appears in modern sets of Maxwell's equations and is often referred to as Faraday's

law. However, because it contains only partial time derivatives, its application is restricted to

situations where the test charge is stationary in a time varying magnetic field. It does not account

for electromagnetic induction in situations where a charged particle is moving in a magnetic

field.

It also can be written in an integral form by the Kelvin-Stokes theorem:[18]

where the movement of the derivative before the integration requires a time-independent surface

Σ (considered in this context to be part of the interpretation of the partial derivative), and as

indicated in Figure 6:

Page 13: Basic Laws and Rules in Electrical Power

Σ is a surface bounded by the closed contour ∂Σ; both Σ and ∂Σ are fixed, independent of

time

E is the electric field,

dℓ is an infinitesimal vector element of the contour ∂Σ,

B is the magnetic field.

dA is an infinitesimal vector element of surface Σ , whose magnitude is the area of an

infinitesimal patch of surface, and whose direction is orthogonal to that surface patch.

Both dℓ and dA have a sign ambiguity; to get the correct sign, the right-hand rule is used, as

explained in the article Kelvin-Stokes theorem. For a planar surface Σ, a positive path element dℓ

of curve ∂Σ is defined by the right-hand rule as one that points with the fingers of the right hand

when the thumb points in the direction of the normal n to the surface Σ.

The integral around ∂Σ is called a path integral or line integral. The surface integral at the right-

hand side of the Maxwell-Faraday equation is the explicit expression for the magnetic flux ΦB

through Σ. Notice that a nonzero path integral for E is different from the behavior of the electric

field generated by charges. A charge-generated E-field can be expressed as the gradient of a

scalar field that is a solution to Poisson's equation, and has a zero path integral. See gradient

theorem.

The integral equation is true for any path ∂Σ through space, and any surface Σ for which that

path is a boundary. Note, however, that ∂Σ and Σ are understood not to vary in time in this

formula. This integral form cannot treat motional EMF because Σ is time-independent. Notice as

well that this equation makes no reference to EMF ,  and indeed cannot do so without

introduction of the Lorentz force law to enable a calculation of work.

Figure 7: Area swept out by vector element dℓ of curve ∂Σ in time dt when moving with velocity

v.

Using the complete Lorentz force to calculate the EMF,

a statement of Faraday's law of induction more general than the integral form of the Maxwell-

Faraday equation is (see Lorentz force):

Page 14: Basic Laws and Rules in Electrical Power

where ∂Σ(t) is the moving closed path bounding the moving surface Σ(t), and v is the velocity of

movement. See Figure 2. Notice that the ordinary time derivative is used, not a partial time

derivative, implying the time variation of Σ(t) must be included in the differentiation. In the

integrand the element of the curve dℓ moves with velocity v.

Figure 7 provides an interpretation of the magnetic force contribution to the EMF on the left side

of the above equation. The area swept out by segment dℓ of curve ∂Σ in time dt when moving

with velocity v is (see geometric meaning of cross-product):

so the change in magnetic flux ΔΦB through the portion of the surface enclosed by ∂Σ in time dt

is:

and if we add these ΔΦB-contributions around the loop for all segments dℓ, we obtain the

magnetic force contribution to Faraday's law. That is, this term is related to motional EMF.

Example: viewpoint of a moving observer

Revisiting the example of Figure 3 in a moving frame of reference brings out the close

connection between E- and B-fields, and between motional and induced EMF's.[19]

Imagine an

observer of the loop moving with the loop. The observer calculates the EMF around the loop

using both the Lorentz force law and Faraday's law of induction. Because this observer moves

with the loop, the observer sees no movement of the loop, and zero v × B. However, because the

B-field varies with position x, the moving observer sees a time-varying magnetic field, namely:

where k is a unit vector pointing in the z-direction.[20]

Lorentz force law version

The Maxwell-Faraday equation says the moving observer sees an electric field Ey in the y-

direction given by:

Page 15: Basic Laws and Rules in Electrical Power

Here the chain rule is used:

Solving for Ey, to within a constant that contributes nothing to an integral around the loop,

Using the Lorentz force law, which has only an electric field component, the observer finds the

EMF around the loop at a time t to be:

which is exactly the same result found by the stationary observer, who sees the centroid xC has

advanced to a position xC + v t. However, the moving observer obtained the result under the

impression that the Lorentz force had only an electric component, while the stationary observer

thought the force had only a magnetic component.

Faraday's law of induction

Using Faraday's law of induction, the observer moving with xC sees a changing magnetic flux,

but the loop does not appear to move: the center of the loop xC is fixed because the moving

observer is moving with the loop. The flux is then:

where the minus sign comes from the normal to the surface pointing oppositely to the applied B-

field. The EMF from Faraday's law of induction is now:

Page 16: Basic Laws and Rules in Electrical Power

the same result. The time derivative passes through the integration because the limits of

integration have no time dependence. Again, the chain rule was used to convert the time

derivative to an x-derivative.

The stationary observer thought the EMF was a motional EMF, while the moving observer

thought it was an induced EMF.[21]

Electrical generator

Figure 8: Faraday's disc electric generator. The disc rotates with angular rate ω, sweeping the

conducting radius circularly in the static magnetic field B. The magnetic Lorentz force v × B

drives the current along the conducting radius to the conducting rim, and from there the circuit

completes through the lower brush and the axle supporting the disc. Thus, current is generated

from mechanical motion.

The EMF generated by Faraday's law of induction due to relative movement of a circuit and a

magnetic field is the phenomenon underlying electrical generators. When a permanent magnet is

moved relative to a conductor, or vice versa, an electromotive force is created. If the wire is

connected through an electrical load, current will flow, and thus electrical energy is generated,

converting the mechanical energy of motion to electrical energy. For example, the drum

generator is based upon Figure 4. A different implementation of this idea is the Faraday's disc,

shown in simplified form in Figure 8. Note that either the analysis of Figure 5, or direct

application of the Lorentz force law, shows that a solid conducting disc works the same way.

In the Faraday's disc example, the disc is rotated in a uniform magnetic field perpendicular to the

disc, causing a current to flow in the radial arm due to the Lorentz force. It is interesting to

understand how it arises that mechanical work is necessary to drive this current. When the

generated current flows through the conducting rim, a magnetic field is generated by this current

through Ampere's circuital law (labeled "induced B" in Figure 8). The rim thus becomes an

electromagnet that resists rotation of the disc (an example of Lenz's law). On the far side of the

figure, the return current flows from the rotating arm through the far side of the rim to the bottom

brush. The B-field induced by this return current opposes the applied B-field, tending to decrease

the flux through that side of the circuit, opposing the increase in flux due to rotation. On the near

side of the figure, the return current flows from the rotating arm through the near side of the rim

to the bottom brush. The induced B-field increases the flux on this side of the circuit, opposing

Page 17: Basic Laws and Rules in Electrical Power

the decrease in flux due to rotation. Thus, both sides of the circuit generate an emf opposing the

rotation. The energy required to keep the disc moving, despite this reactive force, is exactly equal

to the electrical energy generated (plus energy wasted due to friction, Joule heating, and other

inefficiencies). This behavior is common to all generators converting mechanical energy to

electrical energy.

Although Faraday's law always describes the working of electrical generators, the detailed

mechanism can differ in different cases. When the magnet is rotated around a stationary

conductor, the changing magnetic field creates an electric field, as described by the Maxwell-

Faraday equation, and that electric field pushes the charges through the wire. This case is called

an induced EMF. On the other hand, when the magnet is stationary and the conductor is rotated,

the moving charges experience a magnetic force (as described by the Lorentz force law), and this

magnetic force pushes the charges through the wire. This case is called motional EMF. (For

more information on motional EMF, induced EMF, Faraday's law, and the Lorentz force, see

above example, and see Griffiths.)[22]

Electrical motor

An electrical generator can be run "backwards" to become a motor. For example, with the

Faraday disc, suppose a DC current is driven through the conducting radial arm by a voltage.

Then by the Lorentz force law, this traveling charge experiences a force in the magnetic field B

that will turn the disc in a direction given by Fleming's left hand rule. In the absence of

irreversible effects, like friction or Joule heating, the disc turns at the rate necessary to make d

ΦB / dt equal to the voltage driving the current.

Electrical transformer

The EMF predicted by Faraday's law is also responsible for electrical transformers. When the

electric current in a loop of wire changes, the changing current creates a changing magnetic field.

A second wire in reach of this magnetic field will experience this change in magnetic field as a

change in its coupled magnetic flux, a d ΦB / d t. Therefore, an electromotive force is set up in

the second loop called the induced EMF or transformer EMF. If the two ends of this loop are

connected through an electrical load, current will flow.

Magnetic flow meter

Faraday's law is used for measuring the flow of electrically conductive liquids and slurries. Such

instruments are called magnetic flow meters. The induced voltage ℇ generated in the magnetic

field B due to a conductive liquid moving at velocity v is thus given by:

,

where ℓ is the distance between electrodes in the magnetic flow meter.

Page 18: Basic Laws and Rules in Electrical Power

Parasitic induction and waste heating

All metal objects moving in relation to a static magnetic field will experience inductive power

flow, as do all stationary metal objects in relation to a moving magnetic field. These power flows

are occasionally undesirable, resulting in flowing electric current at very low voltage and heating

of the metal.

There are a number of methods employed to control these undesirable inductive effects.

Electromagnets in electric motors, generators, and transformers do not use solid metal,

but instead use thin sheets of metal plate, called laminations. These thin plates reduce the

parasitic eddy currents, as described below.

Inductive coils in electronics typically use magnetic cores to minimize parasitic current

flow. They are a mixture of metal powder plus a resin binder that can hold any shape.

The binder prevents parasitic current flow through the powdered metal.

Electromagnet laminations

Eddy currents occur when a solid metallic mass is rotated in a magnetic field, because the outer

portion of the metal cuts more lines of force than the inner portion, hence the induced

electromotive force not being uniform, tends to set up currents between the points of greatest and

least potential. Eddy currents consume a considerable amount of energy and often cause a

harmful rise in temperature.[23]

Only five laminations or plates are shown in this example, so as to show the subdivision of the

eddy currents. In practical use, the number of laminations or punchings ranges from 40 to 66 per

inch, and brings the eddy current loss down to about one percent. While the plates can be

separated by insulation, the voltage is so low that the natural rust/oxide coating of the plates is

enough to prevent current flow across the laminations.[24]

Page 19: Basic Laws and Rules in Electrical Power

This is a rotor approximately 20mm in diameter from a DC motor used in a CD player. Note the

laminations of the electromagnet pole pieces, used to limit parasitic inductive losses.

Parasitic induction within inductors

In this illustration, a solid copper bar inductor on a rotating armature is just passing under the tip

of the pole piece N of the field magnet. Note the uneven distribution of the lines of force across

the bar inductor. The magnetic field is more concentrated and thus stronger on the left edge of

the copper bar (a,b) while the field is weaker on the right edge (c,d). Since the two edges of the

bar move with the same velocity, this difference in field strength across the bar creates whirls or

current eddies within the copper bar.[25]

This is one reason high voltage devices tend to be more efficient than low voltage devices. High

voltage devices use many turns of small-gauge wire in motors, generators, and transformers.

These many small turns of inductor wire in the electromagnet break up the eddy flows that can

form within the large, thick inductors of low voltage, high current devices.

Maxwell's equations

Maxwell's equations are a set of four partial differential equations that, together with the

Lorentz force law, form the foundation of classical electrodynamics, classical optics, and electric

Page 20: Basic Laws and Rules in Electrical Power

circuits. These in turn underlie the present radio-, television-, phone-, and information-

technologies.

Maxwell's equations have two major variants. The "microscopic" set of Maxwell's equations uses

total charge and total current including the difficult to calculate atomic level charges and currents

in materials. The "macroscopic" set of Maxwell's equations defines two new auxiliary fields that

can sidestep having to know these 'atomic' sized charges and currents.

Maxwell's equations are named after the Scottish physicist and mathematician James Clerk

Maxwell, since they are all found in a four-part paper, On Physical Lines of Force, which he

published between 1861 and 1862. The mathematical form of the Lorentz force law also

appeared in this paper.

It is often useful to write Maxwell's equations in other forms which are often called Maxwell's

equations as well. A relativistic formulation in terms of covariant field tensors is used in special

relativity. While, in quantum mechanics, a version based on the electric and magnetic potentials

is preferred.

Conceptual description

Conceptually, Maxwell's equations describe how electric charges and electric currents act as

sources for the electric and magnetic fields. Further, it describes how a time varying electric field

generates a time varying magnetic field and vice versa. (See below for a mathematical

description of these laws.) Of the four equations, two of them, Gauss's law and Gauss's law for

magnetism, describe how the fields emanate from charges. (For the magnetic field there is no

magnetic charge and therefore magnetic fields lines neither begin nor end anywhere.) The other

two equations describe how the fields 'circulate' around their respective sources; the magnetic

field 'circulates' around electric currents and time varying electric field in Ampère's law with

Maxwell's correction, while the electric field 'circulates' around time varying magnetic fields in

Faraday's law.

Gauss's law

Gauss's law describes the relationship between an electric field and the generating electric

charges: The electric field points away from positive charges and towards negative charges. In

the field line description, electric field lines begin only at positive electric charges and end only

at negative electric charges. 'Counting' the number of field lines in a closed surface, therefore,

yields the total charge enclosed by that surface. More technically, it relates the electric flux

through any hypothetical closed "Gaussian surface" to the electric charge within the surface.

Page 21: Basic Laws and Rules in Electrical Power

Gauss's law for magnetism: magnetic field lines never begin nor end but form loops or extend to

infinity as shown here with the magnetic field due to a ring of current.

Gauss's law for magnetism

Gauss's law for magnetism states that there are no "magnetic charges" (also called magnetic

monopoles), analogous to electric charges.[1]

Instead, the magnetic field due to materials is

generated by a configuration called a dipole. Magnetic dipoles are best represented as loops of

current but resembles a positive and negative 'magnetic charges' inseparably bound together and

having no net 'magnetic charge'. In terms of field lines, this equation states that magnetic field

lines neither begin nor end but make loops or extend to infinity and back. In other words, any

magnetic field line that enters a given volume must somewhere exit that volume. Equivalent

technical statements are that the total magnetic flux through any Gaussian surface is zero, or that

the magnetic field is a solenoidal vector field.

Faraday's law

In a geomagnetic storm, a surge in the flux of charged particles temporarily alters Earth's

magnetic field, which induces electric fields in Earth's atmosphere, thus causing surges in our

electrical power grids.

Faraday's law describes how a time varying magnetic field creates ("induces") an electric

field.[1]

This aspect of electromagnetic induction is the operating principle behind many electric

generators: for example a rotating bar magnet creates a changing magnetic field, which in turn

generates an electric field in a nearby wire. (Note: there are two closely related equations which

are called Faraday's law. The form used in Maxwell's equations is always valid but more

restrictive than that originally formulated by Michael Faraday.)

Page 22: Basic Laws and Rules in Electrical Power

Ampère's law with Maxwell's correction

An Wang's magnetic core memory (1954) is an application of Ampere's law. Each core stores

one bit of data.

Ampère's law with Maxwell's correction states that magnetic fields can be generated in two

ways: by electrical current (this was the original "Ampère's law") and by changing electric fields

(this was "Maxwell's correction").

Maxwell's correction to Ampère's law is particularly important: It means that a changing

magnetic field creates an electric field, and a changing electric field creates a magnetic field.[1][2]

Therefore, these equations allow self-sustaining "electromagnetic waves" to travel through empty

space (see electromagnetic wave equation).

The speed calculated for electromagnetic waves, which could be predicted from experiments on

charges and currents,[note 1]

exactly matches the speed of light; indeed, light is one form of

electromagnetic radiation (as are X-rays, radio waves, and others). Maxwell understood the

connection between electromagnetic waves and light in 1861, thereby unifying the previously-

separate fields of electromagnetism and optics.

Units and summary of equations

Maxwell's equations vary with the unit system used. Though the general form remains the same,

various definitions get changed and different constants appear at different places. The equations

in this section are given in SI units. Other units commonly used are Gaussian units (based on the

cgs system[3]

), Lorentz-Heaviside units (used mainly in particle physics) and Planck units (used

in theoretical physics). See below for CGS-Gaussian units.

For a description of the difference between the microscopic and macroscopic variants of

Maxwell's equations see the relevant sections below.

In the equations given below, symbols in bold represent vector quantities, and symbols in italics

represent scalar quantities. The definitions of terms used in the two tables of equations are given

in another table immediately following.

Table of 'microscopic' equations

Page 23: Basic Laws and Rules in Electrical Power

Formulation in terms of total charge and current[note 2]

Name Differential form Integral form

Gauss's law

Gauss's law for

magnetism

Maxwell–Faraday

equation

(Faraday's law of

induction)

Ampère's circuital law

(with Maxwell's

correction)

Table of 'macroscopic' equations

Formulation in terms of free charge and current

Name Differential form Integral form

Gauss's law

Gauss's law for magnetism

Maxwell–Faraday equation

(Faraday's law of induction)

Ampère's circuital law

(with Maxwell's correction)

Table of terms used in Maxwell's equations

The following table provides the meaning of each symbol and the SI unit of measure:

Definitions and units

Symbol Meaning (first term is the most common) SI Unit of Measure

electric field

also called the electric field intensity

volt per meter or,

equivalently,

Page 24: Basic Laws and Rules in Electrical Power

newton per coulomb

magnetic field

also called the magnetic induction

also called the magnetic field density

also called the magnetic flux density

tesla, or equivalently,

weber per square meter,

volt-second per square

meter

electric displacement field

also called the electric induction

also called the electric flux density

coulombs per square

meter or equivalently,

newton per volt-meter

magnetizing field

also called auxiliary magnetic field

also called magnetic field intensity

also called magnetic field

ampere per meter

the divergence operator per meter (factor

contributed by applying

either operator) the curl operator

partial derivative with respect to time

per second (factor

contributed by applying

the operator)

differential vector element of surface area A,

with infinitesimally small magnitude and

direction normal to surface S

square meters

differential vector element of path length

tangential to the path/curve meters

permittivity of free space, also called the

electric constant, a universal constant farads per meter

permeability of free space, also called the

magnetic constant, a universal constant

henries per meter, or

newtons per ampere

squared

free charge density (not including bound

charge)

coulombs per cubic

meter

total charge density (including both free and

bound charge)

coulombs per cubic

meter

free current density (not including bound

current)

amperes per square

meter

total current density (including both free and

bound current)

amperes per square

meter

net free electric charge within the three-

dimensional volume V (not including bound

charge)

coulombs

net electric charge within the three-

dimensional volume V (including both free

and bound charge)

coulombs

Page 25: Basic Laws and Rules in Electrical Power

line integral of the electric field along the

boundary ∂S of a surface S (∂S is always a

closed curve).

joules per coulomb

line integral of the magnetic field over the

closed boundary ∂S of the surface S tesla-meters

the electric flux (surface integral of the

electric field) through the (closed) surface

(the boundary of the volume V)

joule-meter per coulomb

the magnetic flux (surface integral of the

magnetic B-field) through the (closed)

surface (the boundary of the volume V)

tesla meters-squared or

webers

magnetic flux through any surface S, not

necessarily closed

webers or equivalently,

volt-seconds

electric flux through any surface S, not

necessarily closed

joule-meters per

coulomb

flux of electric displacement field through

any surface S, not necessarily closed coulombs

net free electrical current passing through the

surface S (not including bound current) amperes

net electrical current passing through the

surface S (including both free and bound

current)

amperes

Proof that the two general formulations are equivalent

The two alternate general formulations of Maxwell's equations given above are mathematically

equivalent and related by the following relations:

Page 26: Basic Laws and Rules in Electrical Power

where P and M are polarization and magnetization, and ρb and Jb are bound charge and current,

respectively. Substituting these equations into the 'macroscopic' Maxwell's equations gives

identically the microscopic equations.

Maxwell's 'microscopic' equations

The microscopic variant of Maxwell's equation expresses the electric E field and the magnetic B

field in terms of the total charge and total current present including the charges and currents at

the atomic level. It is sometimes called the general form of Maxwell's equations or "Maxwell's

equations in a vacuum". Both variants of Maxwell's equations are equally general, though, as

they are mathematically equivalent. The microscopic equations are most useful in waveguides

for example, when there are no dielectric or magnetic materials nearby.

Formulation in terms of total charge and current[note 3]

Name Differential form Integral form

Gauss's law

Gauss's law for

magnetism

Maxwell–Faraday

equation

(Faraday's law of

induction)

Ampère's circuital law

(with Maxwell's

correction)

With neither charges nor currents

In a region with no charges (ρ = 0) and no currents (J = 0), such as in a vacuum, Maxwell's

equations reduce to:

These equations lead directly to E and B satisfying the wave equation for which the solutions are

linear combinations of plane waves traveling at the speed of light,

Page 27: Basic Laws and Rules in Electrical Power

In addition, E and B are mutually perpendicular to each other and the direction of motion and are

in phase with each other. A sinusoidal plane wave is one special solution of these equations.

In fact, Maxwell's equations explain how these waves can physically propagate through space.

The changing magnetic field creates a changing electric field through Faraday's law. In turn, that

electric field creates a changing magnetic field through Maxwell's correction to Ampère's law.

This perpetual cycle allows these waves, now known as electromagnetic radiation, to move

through space at velocity c.

Maxwell's 'macroscopic' equations

Unlike the 'microscopic' equations, "Maxwell's macroscopic equations", also known as

Maxwell's equations in matter, factor out the bound charge and current to obtain equations that

depend only on the free charges and currents. These equations are more similar to those that

Maxwell himself introduced. The cost of this factorization is that additional fields need to be

defined: the displacement field D which is defined in terms of the electric field E and the

polarization P of the material, and the magnetic-H field, which is defined in terms of the

magnetic-B field and the magnetization M of the material.

Bound charge and current

Left: A schematic view of how an assembly of microscopic dipoles produces opposite surface

charges as shown at top and bottom. Right: How an assembly of microscopic current loops add

together to produce a macroscopically circulating current loop. Inside the boundaries, the

individual contributions tend to cancel, but at the boundaries no cancellation occurs.

When an electric field is applied to a dielectric material its molecules respond by forming

microscopic electric dipoles—their atomic nuclei move a tiny distance in the direction of the

field, while their electrons move a tiny distance in the opposite direction. This produces a

macroscopic bound charge in the material even though all of the charges involved are bound to

individual molecules. For example, if every molecule responds the same, similar to that shown in

the figure, these tiny movements of charge combine to produce a layer of positive bound charge

Page 28: Basic Laws and Rules in Electrical Power

on one side of the material and a layer of negative charge on the other side. The bound charge is

most conveniently described in terms of a polarization, P, in the material. If P is uniform, a

macroscopic separation of charge is produced only at the surfaces where P enter and leave the

material. For non-uniform P, a charge is also produced in the bulk.[5]

Somewhat similarly, in all materials the constituent atoms exhibit magnetic moments that are

intrinsically linked to the angular momentum of the atoms' components, most notably their

electrons. The connection to angular momentum suggests the picture of an assembly of

microscopic current loops. Outside the material, an assembly of such microscopic current loops

is not different from a macroscopic current circulating around the material's surface, despite the

fact that no individual magnetic moment is traveling a large distance. These bound currents can

be described using the magnetization M.[6]

The very complicated and granular bound charges and bound currents, therefore can be

represented on the macroscopic scale in terms of P and M which average these charges and

currents on a sufficiently large scale so as not to see the granularity of individual atoms, but also

sufficiently small that they vary with location in the material. As such, the Maxwell's

macroscopic equations ignores many details on a fine scale that may be unimportant to

understanding matters on a grosser scale by calculating fields that are averaged over some

suitably sized volume.

Equations

Formulation in terms of free charge and current

Name Differential form Integral form

Gauss's law

Gauss's law for magnetism

Maxwell–Faraday equation

(Faraday's law of induction)

Ampère's circuital law

(with Maxwell's correction)

Constitutive relations

In order to apply 'Maxwell's macroscopic equations', it is necessary to specify the relations

between displacement field D and E, and the magnetic H-field H and B. These equations specify

Page 29: Basic Laws and Rules in Electrical Power

the response of bound charge and current to the applied fields and are called constitutive

relations.

Determining the constitutive relationship between the auxiliary fields D and H and the E and B

fields starts with the definition of the auxiliary fields themselves:

where P is the polarization field and M is the magnetization field which are defined in terms of

microscopic bound charges and bound current respectively. Before getting to how to calculate M

and P it is useful to examine some special cases, though.

Without magnetic or dielectric materials

In the absence of magnetic or dielectric materials, the constitutive relations are simple:

where ε0 and μ0 are two universal constants, called the permittivity of free space and permeability

of free space, respectively. Substituting these back into Maxwell's macroscopic equations lead

directly to Maxwell's microscopic equations, except that the currents and charges are replaced

with free currents and free charges. This is expected since there are no bound charges nor

currents.

Isotropic Linear materials

In an (isotropic[7]

) linear material, where P is proportional to E and M is proportional to B the

constitutive relations are also straightforward. In terms of the polarizaton P and the

magnetization M they are:

where χe and χm are the electric and magnetic susceptibilities of a given material respectively. In

terms of D and H the constitutive relations are:

where ε and μ are constants (which depend on the material), called the permittivity and

permeability, respectively, of the material. These are related to the susceptibilities by:

Page 30: Basic Laws and Rules in Electrical Power

Substituting in the constitutive relations above into Maxwell's equations in linear, dispersionless,

time-invariant materials (differential form only) are:

These are formally identical to the general formulation in terms of E and B (given above),

except that the permittivity of free space was replaced with the permittivity of the material, the

permeability of free space was replaced with the permeability of the material, and only free

charges and currents are included (instead of all charges and currents). Unless that material is

homogeneous in space, ε and μ cannot be factored out of the derivative expressions on the left

sides.

General case

For real-world materials, the constitutive relations are not linear, except approximately.

Calculating the constitutive relations from first principles involves determining how P and M are

created from a given E and B.[note 4]

These relations may be empirical (based directly upon

measurements), or theoretical (based upon statistical mechanics, transport theory or other tools of

condensed matter physics). The detail employed may be macroscopic or microscopic, depending

upon the level necessary to the problem under scrutiny.

In general, though the constitutive relations can usually still be written:

but ε and μ are not, in general, simple constants, but rather functions. Examples are:

Dispersion and absorption where ε and μ are functions of frequency. (Causality does not

permit materials to be nondispersive; see, for example, Kramers–Kronig relations).

Neither do the fields need to be in phase which leads to ε and μ being complex. This also

leads to absorption.

Bi-(an)isotropy where H and D depend on both B and E:[8]

Nonlinearity where ε and μ are functions of E and B.

Anisotropy (such as birefringence or dichroism) which occurs when ε and μ are second-

rank tensors,

Page 31: Basic Laws and Rules in Electrical Power

Dependence of P and M on E and B at other locations and times. This could be due to

spatial inhomogeneity; for example in a domained structure, heterostructure or a liquid

crystal, or most commonly in the situation where there are simply multiple materials

occupying different regions of space). Or it could be due to a time varying medium or

due to hysteresis. In such cases P and M can be calculated as:[9][10]

in which the permittivity and permeability functions are replaced by integrals over the

more general electric and magnetic susceptibilities.[11]

In practice, some materials properties have a negligible impact in particular circumstances,

permitting neglect of small effects. For example: optical nonlinearities can be neglected for low

field strengths; material dispersion is unimportant when frequency is limited to a narrow

bandwidth; material absorption can be neglected for wavelengths for which a material is

transparent; and metals with finite conductivity often are approximated at microwave or longer

wavelengths as perfect metals with infinite conductivity (forming hard barriers with zero skin

depth of field penetration).

It may be noted that man-made materials can be designed to have customized permittivity and

permeability, such as metamaterials and photonic crystals.

Calculation of constitutive relations

In general, the constitutive equations are theoretically determined by calculating how a molecule

responds to the local fields through the Lorentz force. Other forces may need to be modeled as

well such as lattice vibrations in crystals or bond forces. Including all of the forces leads to

changes in the molecule which are used to calculate P and M as a function of the local fields.

The local fields differ from the applied fields due to the fields produced by the polarization and

magnetization of nearby material; an effect which also needs to be modeled. Further, real

materials are not continuous media; the local fields of real materials vary wildly on the atomic

scale. The fields need to be averaged over a suitable volume to form a continuum approximation.

These continuum approximations often require some type of quantum mechanical analysis such

as quantum field theory as applied to condensed matter physics. See, for example, density

functional theory, Green–Kubo relations and Green's function. Various approximate transport

equations have evolved, for example, the Boltzmann equation or the Fokker–Planck equation or

the Navier–Stokes equations. Some examples where these equations are applied are

magnetohydrodynamics, fluid dynamics, electrohydrodynamics, superconductivity, plasma

modeling. An entire physical apparatus for dealing with these matters has developed. A different

set of homogenization methods (evolving from a tradition in treating materials such as

conglomerates and laminates) are based upon approximation of an inhomogeneous material by a

Page 32: Basic Laws and Rules in Electrical Power

homogeneous effective medium[12][13]

(valid for excitations with wavelengths much larger than

the scale of the inhomogeneity).[14][15][16][17]

The theoretical modeling of the continuum-approximation properties of many real materials

often rely upon measurement as well,[18]

for example, ellipsometry measurements.

History

Relation between electricity, magnetism, and the speed of light

The relation between electricity, magnetism, and the speed of light can be summarized by the

modern equation:

The left-hand side is the speed of light, and the right-hand side is a quantity related to the

equations governing electricity and magnetism. Although the right-hand side has units of

velocity, it can be inferred from measurements of electric and magnetic forces, which involve no

physical velocities. Therefore, establishing this relationship provided convincing evidence that

light is an electromagnetic phenomenon.

The discovery of this relationship started in 1855, when Wilhelm Eduard Weber and Rudolf

Kohlrausch determined that there was a quantity related to electricity and magnetism, "the ratio

of the absolute electromagnetic unit of charge to the absolute electrostatic unit of charge" (in

modern language, the value ), and determined that it should have units of velocity.

They then measured this ratio by an experiment which involved charging and discharging a

Leyden jar and measuring the magnetic force from the discharge current, and found a value

3.107×108

m/s,[19]

remarkably close to the speed of light, which had recently been measured at 3.14×108

m/s by Hippolyte Fizeau in 1848 and at 2.98×108

m/s by Léon Foucault in 1850.[19]

However, Weber and Kohlrausch did not make the connection

to the speed of light.[19]

Towards the end of 1861 while working on part III of his paper On

Physical Lines of Force, Maxwell travelled from Scotland to London and looked up Weber and

Kohlrausch's results. He converted them into a format which was compatible with his own

writings, and in doing so he established the connection to the speed of light and concluded that

light is a form of electromagnetic radiation.[20]

The term Maxwell's equations

The four modern Maxwell's equations can be found individually throughout his 1861 paper,

derived theoretically using a molecular vortex model of Michael Faraday's "lines of force" and in

conjunction with the experimental result of Weber and Kohlrausch. But it wasn't until 1884 that

Oliver Heaviside,[21]

concurrently with similar work by Willard Gibbs and Heinrich Hertz,[22]

grouped the four together into a distinct set. This group of four equations was known variously

Page 33: Basic Laws and Rules in Electrical Power

as the Hertz-Heaviside equations and the Maxwell-Hertz equations,[21]

and are sometimes still

known as the Maxwell–Heaviside equations.[23]

Maxwell's contribution to science in producing these equations lies in the correction he made to

Ampère's circuital law in his 1861 paper On Physical Lines of Force. He added the displacement

current term to Ampère's circuital law and this enabled him to derive the electromagnetic wave

equation in his later 1865 paper A Dynamical Theory of the Electromagnetic Field and

demonstrate the fact that light is an electromagnetic wave. This fact was then later confirmed

experimentally by Heinrich Hertz in 1887. The physicist Richard Feynman predicted that, "The

American Civil War will pale into provincial insignificance in comparison with this important

scientific event of the same decade."[24]

The concept of fields was introduced by, among others, Faraday. Albert Einstein wrote:

The precise formulation of the time-space laws was the work of Maxwell. Imagine his feelings

when the differential equations he had formulated proved to him that electromagnetic fields

spread in the form of polarised waves, and at the speed of light! To few men in the world has

such an experience been vouchsafed ... it took physicists some decades to grasp the full

significance of Maxwell's discovery, so bold was the leap that his genius forced upon the

conceptions of his fellow-workers

—(Science, May 24, 1940)

Heaviside worked to eliminate the potentials (electric potential and magnetic potential) that

Maxwell had used as the central concepts in his equations;[21]

this effort was somewhat

controversial,[25]

though it was understood by 1884 that the potentials must propagate at the

speed of light like the fields, unlike the concept of instantaneous action-at-a-distance like the

then conception of gravitational potential.[22]

Modern analysis of, for example, radio antennas,

makes full use of Maxwell's vector and scalar potentials to separate the variables, a common

technique used in formulating the solutions of differential equations. However the potentials can

be introduced by algebraic manipulation of the four fundamental equations.

On Physical Lines of Force

The four modern day Maxwell's equations appeared throughout Maxwell's 1861 paper On

Physical Lines of Force:

1. Equation (56) in Maxwell's 1861 paper is ∇ ⋅ B = 0.

2. Equation (112) is Ampère's circuital law with Maxwell's displacement current added. It is

the addition of displacement current that is the most significant aspect of Maxwell's work

in electromagnetism, as it enabled him to later derive the electromagnetic wave equation

in his 1865 paper A Dynamical Theory of the Electromagnetic Field, and hence show that

light is an electromagnetic wave. It is therefore this aspect of Maxwell's work which

gives the equations their full significance. (Interestingly, Kirchhoff derived the

telegrapher's equations in 1857 without using displacement current. But he did use

Poisson's equation and the equation of continuity which are the mathematical ingredients

of the displacement current. Nevertheless, Kirchhoff believed his equations to be

Page 34: Basic Laws and Rules in Electrical Power

applicable only inside an electric wire and so he is not credited with having discovered

that light is an electromagnetic wave).

3. Equation (115) is Gauss's law.

4. Equation (54) is an equation that Oliver Heaviside referred to as 'Faraday's law'. This

equation caters for the time varying aspect of electromagnetic induction, but not for the

motionally induced aspect, whereas Faraday's original flux law caters for both

aspects.[26][27]

Maxwell deals with the motionally dependent aspect of electromagnetic

induction, v × B, at equation (77). Equation (77) which is the same as equation (D) in the

original eight Maxwell's equations listed below, corresponds to all intents and purposes to

the modern day force law F = q( E + v × B ) which sits adjacent to Maxwell's equations

and bears the name Lorentz force, even though Maxwell derived it when Lorentz was still

a young boy.

The difference between the B and the H vectors can be traced back to Maxwell's 1855 paper

entitled On Faraday's Lines of Force which was read to the Cambridge Philosophical Society.

The paper presented a simplified model of Faraday's work, and how the two phenomena were

related. He reduced all of the current knowledge into a linked set of differential equations.

Figure of Maxwell's molecular vortex model. For a uniform magnetic field, the field lines point

outward from the display screen, as can be observed from the black dots in the middle of the

hexagons. The vortex of each hexagonal molecule rotates counter-clockwise. The small green

circles are clockwise rotating particles sandwiching between the molecular vortices.

It is later clarified in his concept of a sea of molecular vortices that appears in his 1861 paper On

Physical Lines of Force. Within that context, H represented pure vorticity (spin), whereas B was

a weighted vorticity that was weighted for the density of the vortex sea. Maxwell considered

magnetic permeability µ to be a measure of the density of the vortex sea. Hence the relationship,

1. Magnetic induction current causes a magnetic current density

Page 35: Basic Laws and Rules in Electrical Power

was essentially a rotational analogy to the linear electric current relationship,

2. Electric convection current

where ρ is electric charge density. B was seen as a kind of magnetic current of vortices aligned in

their axial planes, with H being the circumferential velocity of the vortices. With µ representing

vortex density, it follows that the product of µ with vorticity H leads to the magnetic field

denoted as B.

The electric current equation can be viewed as a convective current of electric charge that

involves linear motion. By analogy, the magnetic equation is an inductive current involving spin.

There is no linear motion in the inductive current along the direction of the B vector. The

magnetic inductive current represents lines of force. In particular, it represents lines of inverse

square law force.

The extension of the above considerations confirms that where B is to H, and where J is to ρ,

then it necessarily follows from Gauss's law and from the equation of continuity of charge that E

is to D. i.e. B parallels with E, whereas H parallels with D.

A Dynamical Theory of the Electromagnetic Field

In 1864 Maxwell published A Dynamical Theory of the Electromagnetic Field in which he

showed that light was an electromagnetic phenomenon. Confusion over the term "Maxwell's

equations" is exacerbated because it is also sometimes used for a set of eight equations that

appeared in Part III of Maxwell's 1864 paper A Dynamical Theory of the Electromagnetic Field,

entitled "General Equations of the Electromagnetic Field,"[28]

a confusion compounded by the

writing of six of those eight equations as three separate equations (one for each of the Cartesian

axes), resulting in twenty equations and twenty unknowns. (As noted above, this terminology is

not common: Modern references to the term "Maxwell's equations" refer to the Heaviside

restatements.)

The eight original Maxwell's equations can be written in modern vector notation as follows:

(A) The law of total currents

(B) The equation of magnetic force

(C) Ampère's circuital law

(D) Electromotive force created by convection, induction, and by static electricity. (This is in

effect the Lorentz force)

Page 36: Basic Laws and Rules in Electrical Power

(E) The electric elasticity equation

(F) Ohm's law

(G) Gauss's law

(H) Equation of continuity

or

Notation

H is the magnetizing field, which Maxwell called the magnetic intensity.

J is the current density (withJtot being the total current including displacement

current).[note 5]

D is the displacement field (called the electric displacement by Maxwell).

ρ is the free charge density (called the quantity of free electricity by Maxwell).

A is the magnetic potential (called the angular impulse by Maxwell).

E is called the electromotive force by Maxwell. The term electromotive force is

nowadays used for voltage, but it is clear from the context that Maxwell's meaning

corresponded more to the modern term electric field.

υ is the electric potential (which Maxwell also called electric potential).

σ is the electrical conductivity (Maxwell called the inverse of conductivity the specific

resistance, what is now called the resistivity).

It is interesting to note the μv × H term that appears in equation D. Equation D is therefore

effectively the Lorentz force, similarly to equation (77) of his 1861 paper (see above).

When Maxwell derives the electromagnetic wave equation in his 1865 paper, he uses equation D

to cater for electromagnetic induction rather than Faraday's law of induction which is used in

modern textbooks. (Faraday's law itself does not appear among his equations.) However,

Maxwell drops the μv × H term from equation D when he is deriving the electromagnetic wave

equation, as he considers the situation only from the rest frame.

A Treatise on Electricity and Magnetism

In A Treatise on Electricity and Magnetism, an 1873 treatise on electromagnetism written by

James Clerk Maxwell, eleven general equations of the electromagnetic field are listed and these

include the eight that are listed in the 1865 paper.[29]

Page 37: Basic Laws and Rules in Electrical Power

Maxwell's equations and relativity

Maxwell's original equations are based on the idea that light travels through a sea of molecular

vortices known as the 'luminiferous aether', and that the speed of light has to be respective to the

reference frame of this aether. Measurements designed to measure the speed of the Earth through

the aether conflicted, though.[30]

A more theoretical approach was suggested by Hendrik Lorentz along with George FitzGerald

and Joseph Larmor. Both Larmor (1897) and Lorentz (1899, 1904) derived the Lorentz

transformation (so named by Henri Poincaré) as one under which Maxwell's equations were

invariant. Poincaré (1900) analyzed the coordination of moving clocks by exchanging light

signals. He also established mathematically the group property of the Lorentz transformation

(Poincaré 1905).

Einstein dismissed the aether as unnecessary and concluded that Maxwell's equations predict the

existence of a fixed speed of light, independent of the speed of the observer, and as such he used

Maxwell's equations as the starting point for his special theory of relativity. In doing so, he

established the Lorentz transformation as being valid for all matter and not just Maxwell's

equations. Maxwell's equations played a key role in Einstein's famous paper on special relativity;

for example, in the opening paragraph of the paper, he motivated his theory by noting that a

description of a conductor moving with respect to a magnet must generate a consistent set of

fields irrespective of whether the force is calculated in the rest frame of the magnet or that of the

conductor.[31]

General relativity has also had a close relationship with Maxwell's equations. For example,

Theodor Kaluza and Oskar Klein showed in the 1920s that Maxwell's equations can be derived

by extending general relativity into five dimensions. This strategy of using higher dimensions to

unify different forces remains an active area of research in particle physics.

Modified to include magnetic monopoles

Maxwell's equations of electromagnetism relate the electric and magnetic fields to the motions of

electric charges. The standard form of the equations provide for an electric charge, but posit no

magnetic charge. There is no known magnetic analog of an electron, however recently scientists

have described behavior in a crystalline state of matter known as spin-ice which have

macroscopic behavior like magnetic monopoles.[32][33]

(in accordance with the fact that magnetic

charge has never been seen and may not exist). Except for this, the equations are symmetric

under interchange of electric and magnetic field. In fact, symmetric equations can be written

when all charges are zero, and this is how the wave equation is derived (see immediately above).

Fully symmetric equations can also be written if one allows for the possibility of magnetic

charges.[34]

With the inclusion of a variable for these magnetic charges, say ρm, there will also be

a "magnetic current" variable in the equations, Jm. The extended Maxwell's equations (in cgs-

Gaussian units) are as follows:

Page 38: Basic Laws and Rules in Electrical Power

Name Without magnetic

monopoles

With magnetic monopoles

(hypothetical)

Gauss's law:

Gauss's law for

magnetism:

Maxwell–Faraday

equation

(Faraday's law of

induction):

Ampère's law

(with Maxwell's

extension):

If magnetic charges do not exist, or if they exist but where they are not present in a region, then

the new variables are zero, and the symmetric equations reduce to the conventional equations of

electromagnetism such as ∇ · B = 0.

Boundary conditions using Maxwell's equations

Like all sets of differential equations, Maxwell's equations cannot be uniquely solved without a

suitable set of boundary conditions[35][36][37]

and initial conditions.[38]

For example, consider a region with no charges and no currents. One particular solution that

satisfies all of Maxwell's equations in that region is that both E and B = 0 everywhere in the

region. This solution is obviously false if there is a charge just outside of the region. In this

particular example, all of the electric and magnetic fields in the interior are due to the charges

outside of the volume. Different charges outside of the volume produce different fields on the

surface of that volume and therefore have a different boundary conditions. In general, knowing

the appropriate boundary conditions for a given region along with the currents and charges in

that region allows one to solve for all the fields everywhere within that region. An example of

this type is a an electromagnetic scattering problem, where an electromagnetic wave originating

outside the scattering region is scattered by a target, and the scattered electromagnetic wave is

analyzed for the information it contains about the target by virtue of the interaction with the

target during scattering.[39]

In some cases, like waveguides or cavity resonators, the solution region is largely isolated from

the universe, for example, by metallic walls, and boundary conditions at the walls define the

fields with influence of the outside world confined to the input/output ends of the structure.[40]

In

other cases, the universe at large sometimes is approximated by an artificial absorbing

boundary,[41][42][43]

or, for example for radiating antennas or communication satellites, these

boundary conditions can take the form of asymptotic limits imposed upon the solution.[44]

In

addition, for example in an optical fiber or thin-film optics, the solution region often is broken up

into subregions with their own simplified properties, and the solutions in each subregion must be

joined to each other across the subregion interfaces using boundary conditions.[45][46][47]

A

particular example of this use of boundary conditions is the replacement of a material with a

Page 39: Basic Laws and Rules in Electrical Power

volume polarization by a charged surface layer, or of a material with a volume magnetization by

a surface current, as described in the section Bound charge and current.

Following are some links of a general nature concerning boundary value problems: Examples of

boundary value problems, Sturm–Liouville theory, Dirichlet boundary condition, Neumann

boundary condition, mixed boundary condition, Cauchy boundary condition, Sommerfeld

radiation condition. Needless to say, one must choose the boundary conditions appropriate to the

problem being solved. See also Kempel[48]

and the book by Friedman.[49]

Gaussian units

Gaussian units is a popular electromagnetism variant of the centimetre gram second system of

units (cgs). In gaussian units, Maxwell's equations are:[50]

where c is the speed of light in a vacuum. The microscopic equations are:

The relation between electric displacement field, electric field and polarization density is:

And likewise the relation between magnetic induction, magnetic field and total magnetization is:

In the linear approximation, the electric susceptibility and magnetic susceptibility are defined so

that:

,

Page 40: Basic Laws and Rules in Electrical Power

(Note: although the susceptibilities are dimensionless numbers in both cgs and SI, they differ in

value by a factor of 4π.) The permittivity and permeability are:

,

so that

,

In vacuum, ε = μ = 1, therefore D = E, and B = H.

The force exerted upon a charged particle by the electric field and magnetic field is given by the

Lorentz force equation:

where q is the charge on the particle and v is the particle velocity. This is slightly different from

the SI-unit expression above. For example, the magnetic field B has the same units as the electric

field E.

Some equations in the article are given in Gaussian units but not SI or vice-versa. Fortunately,

there are general rules to convert from one to the other; see the article Gaussian units for details.

Alternative formulations of Maxwell's equations

Special relativity motivated a compact mathematical formulation of Maxwell's equations, in

terms of covariant tensors. Quantum mechanics also motivated other formulations.

For example, consider a conductor moving in the field of a magnet.[51]

In the frame of the

magnet, that conductor experiences a magnetic force. But in the frame of a conductor moving

relative to the magnet, the conductor experiences a force due to an electric field. The following

formulation shows how Maxwell's equations take the same form in any inertial coordinate

system.

Covariant formulation of Maxwell's equations

In special relativity, in order to more clearly express the fact that Maxwell's ('microscopic')

equations take the same form in any inertial coordinate system, Maxwell's equations are written

in terms of four-vectors and tensors in the "manifestly covariant" form. The purely spatial

components of the following are in SI units.

One ingredient in this formulation is the electromagnetic tensor, a rank-2 covariant

antisymmetric tensor combining the electric and magnetic fields:

Page 41: Basic Laws and Rules in Electrical Power

and the result of raising its indices

The other ingredient is the four-current:

where ρ is the charge density and J is the current density.

With these ingredients, Maxwell's equations can be written:

and

The first tensor equation is an expression of the two inhomogeneous Maxwell's equations,

Gauss's law and Ampere's law with Maxwell's correction. The second equation is an expression

of the two homogeneous equations, Faraday's law of induction and Gauss's law for magnetism.

The second equation is equivalent to

where is the contravariant version of the Levi-Civita symbol, and

is the 4-gradient. In the tensor equations above, repeated indices are summed over according to

Einstein summation convention. We have displayed the results in several common notations.

Page 42: Basic Laws and Rules in Electrical Power

Upper and lower components of a vector, vα and v

α respectively, are interchanged with the

fundamental tensor g, e.g., g = η = diag(−1, +1, +1, +1).

Alternative covariant presentations of Maxwell's equations also exist, for example in terms of the

four-potential; see Covariant formulation of classical electromagnetism for details.

Potential formulation

In advanced classical mechanics and in quantum mechanics (where it is necessary) it is

sometimes useful to express Maxwell's equations in a 'potential formulation' involving the

electric potential (also called scalar potential), υ, and the magnetic potential, A, (also called

vector potential). These are defined such that:

With these definitions, the two homogeneous Maxwell's equations (Faraday's Law and Gauss's

law for magnetism) are automatically satisfied and the other two (inhomogeneous) equations

give the following equations (for "Maxwell's microscopic equations"):

These equations, taken together, are as powerful and complete as Maxwell's equations.

Moreover, if we work only with the potentials and ignore the fields, the problem has been

reduced somewhat, as the electric and magnetic fields each have three components which need to

be solved for (six components altogether), while the electric and magnetic potentials have only

four components altogether.

Many different choices of A and υ are consistent with a given E and B, making these choices

physically equivalent – a flexibility known as gauge freedom. Suitable choice of A and υ can

simplify these equations, or can adapt them to suit a particular situation.

Four-potential

In the Lorentz gauge, the two equations that represent the potentials can be reduced to one

manifestly Lorentz invariant equation, using four-vectors: the four-current defined by

Page 43: Basic Laws and Rules in Electrical Power

formed from the current density j and charge density ρ, and the electromagnetic four-potential

defined by

formed from the vector potential A and the scalar potential . The resulting single equation, due

to Arnold Sommerfeld, a generalization of an equation due to Bernhard Riemann and known as

the Riemann–Sommerfeld equation[52]

or the covariant form of the Maxwell equations,[53]

is:

,

where is the d'Alembertian operator, or four-Laplacian, ,

sometimes written , or , where is the four-gradient.

Differential formulations

In free space, where ε = ε0 and μ = μ0 are constant everywhere, Maxwell's equations simplify

considerably once the language of differential geometry and differential forms is used. In what

follows, cgs-Gaussian units, not SI units are used. (To convert to SI, see here.) The electric and

magnetic fields are now jointly described by a 2-form F in a 4-dimensional spacetime manifold.

Maxwell's equations then reduce to the Bianchi identity

where d denotes the exterior derivative — a natural coordinate and metric independent

differential operator acting on forms — and the source equation

where the (dual) Hodge star operator * is a linear transformation from the space of 2-forms to the

space of (4-2)-forms defined by the metric in Minkowski space (in four dimensions even by any

metric conformal to this metric), and the fields are in natural units where 1/4πε0 = 1. Here, the 3-

form J is called the electric current form or current 3-form satisfying the continuity equation

The current 3-form can be integrated over a 3-dimensional space-time region. The physical

interpretation of this integral is the charge in that region if it is spacelike, or the amount of charge

that flows through a surface in a certain amount of time if that region is a spacelike surface cross

a timelike interval. As the exterior derivative is defined on any manifold, the differential form

version of the Bianchi identity makes sense for any 4-dimensional manifold, whereas the source

equation is defined if the manifold is oriented and has a Lorentz metric. In particular the

differential form version of the Maxwell equations are a convenient and intuitive formulation of

the Maxwell equations in general relativity.

Page 44: Basic Laws and Rules in Electrical Power

In a linear, macroscopic theory, the influence of matter on the electromagnetic field is described

through more general linear transformation in the space of 2-forms. We call

the constitutive transformation. The role of this transformation is comparable to the Hodge

duality transformation. The Maxwell equations in the presence of matter then become:

where the current 3-form J still satisfies the continuity equation dJ = 0.

When the fields are expressed as linear combinations (of exterior products) of basis forms θp,

the constitutive relation takes the form

where the field coefficient functions are antisymmetric in the indices and the constitutive

coefficients are antisymmetric in the corresponding pairs. In particular, the Hodge duality

transformation leading to the vacuum equations discussed above are obtained by taking

which up to scaling is the only invariant tensor of this type that can be defined with the metric.

In this formulation, electromagnetism generalises immediately to any 4-dimensional oriented

manifold or with small adaptations any manifold, requiring not even a metric. Thus the

expression of Maxwell's equations in terms of differential forms leads to a further notational and

conceptual simplification. Whereas Maxwell's Equations could be written as two tensor

equations instead of eight scalar equations, from which the propagation of electromagnetic

disturbances and the continuity equation could be derived with a little effort, using differential

forms leads to an even simpler derivation of these results.

Conceptual insight from this formulation

On the conceptual side, from the point of view of physics, this shows that the second and third

Maxwell equations should be grouped together, be called the homogeneous ones, and be seen as

geometric identities expressing nothing else than: the field F derives from a more "fundamental"

potential A. While the first and last one should be seen as the dynamical equations of motion,

Page 45: Basic Laws and Rules in Electrical Power

obtained via the Lagrangian principle of least action, from the "interaction term" A J (introduced

through gauge covariant derivatives), coupling the field to matter.

Often, the time derivative in the third law motivates calling this equation "dynamical", which is

somewhat misleading; in the sense of the preceding analysis, this is rather an artifact of breaking

relativistic covariance by choosing a preferred time direction. To have physical degrees of

freedom propagated by these field equations, one must include a kinetic term F *F for A; and

take into account the non-physical degrees of freedom which can be removed by gauge

transformation A → A' = A − dα. See also gauge fixing and Faddeev–Popov ghosts.

Geometric Algebra (GA) formulation

In geometric algebra, Maxwell's equations are reduced to a single equation,

[54]

where F and J are multivectors

and

with the unit pseudoscalar I2 = −1

The GA spatial gradient operator ∇ acts on a vector field, such that

In spacetime algebra using the same geometric product the equation is simply

the spacetime derivative of the electromagnetic field is its source. Here the (non-bold) spacetime

gradient

is a four vector, as is the current density

For a demonstration that the equations given reproduce Maxwell's equations see the main article.

Page 46: Basic Laws and Rules in Electrical Power

Classical electrodynamics as the curvature of a line bundle

An elegant and intuitive way to formulate Maxwell's equations is to use complex line bundles or

principal bundles with fibre U(1). The connection on the line bundle has a curvature

which is a two-form that automatically satisfies and can be interpreted as a

field-strength. If the line bundle is trivial with flat reference connection d we can write

and F = dA with A the 1-form composed of the electric potential and the magnetic

vector potential.

In quantum mechanics, the connection itself is used to define the dynamics of the system. This

formulation allows a natural description of the Aharonov-Bohm effect. In this experiment, a

static magnetic field runs through a long magnetic wire (e.g., an iron wire magnetized

longitudinally). Outside of this wire the magnetic induction is zero, in contrast to the vector

potential, which essentially depends on the magnetic flux through the cross-section of the wire

and does not vanish outside. Since there is no electric field either, the Maxwell tensor F = 0

throughout the space-time region outside the tube, during the experiment. This means by

definition that the connection is flat there.

However, as mentioned, the connection depends on the magnetic field through the tube since the

holonomy along a non-contractible curve encircling the tube is the magnetic flux through the

tube in the proper units. This can be detected quantum-mechanically with a double-slit electron

diffraction experiment on an electron wave traveling around the tube. The holonomy corresponds

to an extra phase shift, which leads to a shift in the diffraction pattern.[55][56]

Curved spacetime

Traditional formulation

Matter and energy generate curvature of spacetime. This is the subject of general relativity.

Curvature of spacetime affects electrodynamics. An electromagnetic field having energy and

momentum also generates curvature in spacetime. Maxwell's equations in curved spacetime can

be obtained by replacing the derivatives in the equations in flat spacetime with covariant

derivatives. (Whether this is the appropriate generalization requires separate investigation.) The

sourced and source-free equations become (cgs-Gaussian units):

and

Here,

Page 47: Basic Laws and Rules in Electrical Power

is a Christoffel symbol that characterizes the curvature of spacetime and Dγ is the covariant

derivative.

Formulation in terms of differential forms

The formulation of the Maxwell equations in terms of differential forms can be used without

change in general relativity. The equivalence of the more traditional general relativistic

formulation using the covariant derivative with the differential form formulation can be seen as

follows. Choose local coordinates xα which gives a basis of 1-forms dx

α in every point of the

open set where the coordinates are defined. Using this basis and cgs-Gaussian units we define

The antisymmetric infinitesimal field tensor Fαβ, corresponding to the field 2-form F

The current-vector infinitesimal 3-form J

Here g is as usual the determinant of the metric tensor gαβ. A small computation that uses the

symmetry of the Christoffel symbols (i.e., the torsion-freeness of the Levi Civita connection) and

the covariant constantness of the Hodge star operator then shows that in this coordinate

neighborhood we have:

the Bianchi identity

the source equation

the continuity equation

Notes

Page 48: Basic Laws and Rules in Electrical Power

1. ^ The quantity we would now call , with units of velocity, was directly measured before

Maxwell's equations, in an 1855 experiment by Wilhelm Eduard Weber and Rudolf Kohlrausch.

They charged a leyden jar (a kind of capacitor), and measured the electrostatic force associated

with the potential; then, they discharged it while measuring the magnetic force from the current in

the discharge-wire. Their result was 3.107×108

m/s, remarkably close to the speed of light. See The story of electrical and magnetic

measurements: from 500 B.C. to the 1940s, by Joseph F. Keithley, p115

2. ^ In some books *e.g., in [4]

), the term effective charge is used instead of total charge, while free

charge is simply called charge.

3. ^ In some books *e.g., in [4]

), the term effective charge is used instead of total charge, while free

charge is simply called charge.

4. ^ The free charges and currents respond to the fields through the Lorentz force law and this

response is calculated at a fundamental level using mechanics. The response of bound charges

and currents is dealt with using grosser methods subsumed under the notions of magnetization

and polarization. Depending upon the problem, one may choose to have no free charges

whatsoever.

5. ^ Here it is noted that a quite different quantity, the magnetic polarization, μ0M by decision of an

international IUPAP commission has been given the same name J. So for the electric current

density, a name with small letters, j would be better. But even then the mathematicians would still

use the large-letter-name J for the corresponding current-twoform (see below).

References

1. ^ a b c J.D. Jackson, "Maxwell's Equations" video glossary entry

2. ^ Principles of physics: a calculus-based text, by R.A. Serway, J.W. Jewett, page 809.

3. ^ David J Griffiths (1999). Introduction to electrodynamics (Third ed.). Prentice Hall. pp. 559–

562. ISBN 013805326X. http://worldcat.org/isbn/013805326X.

4. ^ a b U. Krey and A. Owen's Basic Theoretical Physics (Springer 2007)

5. ^ See David J. Griffiths (1999). Introduction to Electrodynamics (third ed.). Prentice Hall. for a

good description of how P relates to the bound charge.

6. ^ See David J. Griffiths (1999). Introduction to Electrodynamics (third ed.). Prentice Hall. for a

good description of how M relates to the bound current.

7. ^ The generalization to non-isotropic materials is straight forward; simply replace the constants

with tensor quantities.

8. ^ In general materials are bianisotropic. TG Mackay and A Lakhtakia (2010). Electromagnetic

Anisotropy and Bianisotropy: A Field Guide. World Scientific.

http://www.worldscibooks.com/physics/7515.html.

9. ^ Halevi, Peter (1992). Spatial dispersion in solids and plasmas. Amsterdam: North-Holland.

ISBN 978-0444874054.

10. ^ Jackson, John David (1999). Classical Electrodynamics (3rd ed.). New York: Wiley. ISBN 0-

471-30932-X.

11. ^ Note that the 'magnetic susceptibility' term used her is in terms of B and is different from the

standard definition in terms of H.

12. ^ Aspnes, D.E., "Local-field effects and effective-medium theory: A microscopic perspective,"

Am. J. Phys. 50, p. 704-709 (1982).

13. ^ Habib Ammari & Hyeonbae Kang (2006). Inverse problems, multi-scale analysis and effective

medium theory : workshop in Seoul, Inverse problems, multi-scale analysis, and homogenization,

June 22–24, 2005, Seoul National University, Seoul, Korea. Providence RI: American

Mathematical Society. p. 282. ISBN 0821839683.

Page 49: Basic Laws and Rules in Electrical Power

http://books.google.com/?id=dK7JwVPbUkMC&printsec=frontcover&dq=%22effective+mediu

m%22.

14. ^ O. C. Zienkiewicz, Robert Leroy Taylor, J. Z. Zhu, Perumal Nithiarasu (2005). The Finite

Element Method (Sixth ed.). Oxford UK: Butterworth-Heinemann. p. 550 ff. ISBN 0750663219.

http://books.google.com/?id=rvbSmooh8Y4C&printsec=frontcover&dq=finite+element+inauthor

:Zienkiewicz.

15. ^ N. Bakhvalov and G. Panasenko, Homogenization: Averaging Processes in Periodic Media

(Kluwer: Dordrecht, 1989); V. V. Jikov, S. M. Kozlov and O. A. Oleinik, Homogenization of

Differential Operators and Integral Functionals (Springer: Berlin, 1994).

16. ^ Vitaliy Lomakin, Steinberg BZ, Heyman E, & Felsen LB (2003). "Multiresolution

Homogenization of Field and Network Formulations for Multiscale Laminate Dielectric Slabs".

IEEE Transactions on Antennas and Propagation 51 (10): 2761 ff.

doi:10.1109/TAP.2003.816356. http://www.ece.ucsd.edu/~vitaliy/A8.pdf.

17. ^ AC Gilbert (Ronald R Coifman, Editor) (2000-05). Topics in Analysis and Its Applications:

Selected Theses. Singapore: World Scientific Publishing Company. p. 155. ISBN 9810240945.

http://books.google.com/?id=d4MOYN5DjNUC&printsec=frontcover&dq=homogenization+date

:2000-2009.

18. ^ Edward D. Palik & Ghosh G (1998). Handbook of Optical Constants of Solids. London UK:

Academic Press. p. 1114. ISBN 0125444222.

http://books.google.com/?id=AkakoCPhDFUC&dq=optical+constants+inauthor:Palik.

19. ^ a b c The story of electrical and magnetic measurements: from 500 B.C. to the 1940s, by Joseph

F. Keithley, p115

20. ^ "The Dictionary of Scientific Biography", by Charles Coulston Gillispie

21. ^ a b c but are now universally known as Maxwell's equations. However, in 1940 Einstein referred

to the equations as Maxwell's equations in "The Fundamentals of Theoretical Physics" published

in the Washington periodical Science, May 24, 1940. Paul J. Nahin (2002-10-09). Oliver

Heaviside: the life, work, and times of an electrical genius of the Victorian age. JHU Press.

pp. 108–112. ISBN 9780801869099.

http://books.google.com/?id=e9wEntQmA0IC&pg=PA111&dq=nahin+hertz-heaviside+maxwell-

hertz.

22. ^ a b Jed Z. Buchwald (1994). The creation of scientific effects: Heinrich Hertz and electric

waves. University of Chicago Press. p. 194. ISBN 9780226078885.

http://books.google.com/?id=2bDEvvGT1EYC&pg=PA194&dq=maxwell+faraday+time-

derivative+vector-potential.

23. ^ Myron Evans (2001-10-05). Modern nonlinear optics. John Wiley and Sons. p. 240.

ISBN 9780471389316. http://books.google.com/?id=9p0kK6IG94gC&pg=PA240&dq=maxwell-

heaviside+equations.

24. ^ Crease, Robert. The Great Equations: Breakthroughs in Science from Pythagoras to

Heisenberg, page 133 (2008).

25. ^ Oliver J. Lodge (November 1888). "Sketch of the Electrical Papers in Section A, at the Recent

Bath Meeting of the British Association". Electrical Engineer 7: 535.

26. ^ J. R. Lalanne, F. Carmona, and L. Servant (1999-11). Optical spectroscopies of electronic

absorption. World Scientific. p. 8. ISBN 9789810238612. http://books.google.com/?id=7rWD-

TdxKkMC&pg=PA8&dq=maxwell-faraday+derivative.

27. ^ Roger F. Harrington (2003-10-17). Introduction to Electromagnetic Engineering. Courier

Dover Publications. pp. 49–56. ISBN 9780486432410.

http://books.google.com/?id=ZlC2EV8zvX8C&pg=PR7&dq=maxwell-faraday-equation+law-of-

induction.

28. ^ page 480.

29. ^ http://www.mathematik.tu-darmstadt.de/~bruhn/Original-MAXWELL.htm

Page 50: Basic Laws and Rules in Electrical Power

30. ^ Experiments like the Michelson-Morley experiment in 1887 showed that the 'aether' moved at

the same speed as Earth. While other experiments, such as measurements of the aberration of

light from stars, showed that the ether is moving relative to earth.

31. ^ "On the Electrodynamics of Moving Bodies". Fourmilab.ch.

http://www.fourmilab.ch/etexts/einstein/specrel/www/. Retrieved 2008-10-19.

32. ^ http://www.sciencemag.org/cgi/content/abstract/1178868

33. ^ http://www.nature.com/nature/journal/v461/n7266/full/nature08500.html

34. ^ "IEEEGHN: Maxwell's Equations". Ieeeghn.org.

http://www.ieeeghn.org/wiki/index.php/Maxwell%27s_Equations. Retrieved 2008-10-19.

35. ^ Peter Monk; ), Peter Monk (Ph.D (2003). Finite Element Methods for Maxwell's Equations.

Oxford UK: Oxford University Press. p. 1 ff. ISBN 0198508883.

http://books.google.com/?id=zI7Y1jT9pCwC&pg=PA1&dq=electromagnetism+%22boundary+c

onditions%22.

36. ^ Thomas B. A. Senior & John Leonidas Volakis (1995-03-01). Approximate Boundary

Conditions in Electromagnetics. London UK: Institution of Electrical Engineers. p. 261 ff.

ISBN 0852968493.

http://books.google.com/?id=eOofBpuyuOkC&pg=PA261&dq=electromagnetism+%22boundary

+conditions%22.

37. ^ T Hagstrom (Björn Engquist & Gregory A. Kriegsmann, Eds.) (1997). Computational Wave

Propagation. Berlin: Springer. p. 1 ff. ISBN 0387948740.

http://books.google.com/?id=EdZefkIOR5cC&pg=PA1&dq=electromagnetism+%22boundary+c

onditions%22.

38. ^ Henning F. Harmuth & Malek G. M. Hussain (1994). Propagation of Electromagnetic Signals.

Singapore: World Scientific. p. 17. ISBN 9810216890.

http://books.google.com/?id=6_CZBHzfhpMC&pg=PA45&dq=electromagnetism+%22initial+co

nditions%22.

39. ^ Fioralba Cakoni; Colton, David L (2006). "The inverse scattering problem for an imperfect

conductor". Qualitative methods in inverse scattering theory. Springer Science & Business. p. 61.

ISBN 3540288449. http://books.google.com/?id=7DqqOjPJetYC&pg=PR6#PPA61,M1.,

Khosrow Chadan et al. (1997). An introduction to inverse scattering and inverse spectral

problems. Society for Industrial and Applied Mathematics. p. 45. ISBN 0898713870.

http://books.google.com/?id=y2rywYxsDEAC&pg=PA45.

40. ^ S. F. Mahmoud (1991). Electromagnetic Waveguides: Theory and Applications applications.

London UK: Institution of Electrical Engineers. Chapter 2. ISBN 0863412327.

http://books.google.com/?id=toehQ7vLwAMC&pg=PA2&dq=Maxwell%27s+equations+wavegu

ides.

41. ^ Jean-Michel Lourtioz (2005-05-23). Photonic Crystals: Towards Nanoscale Photonic Devices.

Berlin: Springer. p. 84. ISBN 354024431X.

http://books.google.com/?id=vSszZ2WuG_IC&pg=PA84&dq=electromagnetism+boundary++-

element.

42. ^ S. G. Johnson, Notes on Perfectly Matched Layers, online MIT course notes (Aug. 2007).

43. ^ Taflove A & Hagness S C (2005). Computational Electrodynamics: The Finite-difference

Time-domain Method. Boston MA: Artech House. Chapters 6 & 7. ISBN 1580538320.

http://www.amazon.com/gp/reader/1580538320/ref=sib_dp_pop_toc?ie=UTF8&p=S008#reader-

link.

44. ^ David M Cook (2002). The Theory of the Electromagnetic Field. Mineola NY: Courier Dover

Publications. p. 335 ff. ISBN 0486425673. http://books.google.com/?id=bI-

ZmZWeyhkC&pg=RA1-PA335&dq=electromagnetism+infinity+boundary+conditions.

45. ^ Korada Umashankar (1989-09). Introduction to Engineering Electromagnetic Fields.

Singapore: World Scientific. p. §10.7; pp. 359ff. ISBN 9971509210.

Page 51: Basic Laws and Rules in Electrical Power

http://books.google.com/?id=qp5qHvB_mhcC&pg=PA359&dq=electromagnetism+%22boundar

y+conditions%22.

46. ^ Joseph V. Stewart (2001). Intermediate Electromagnetic Theory. Singapore: World Scientific.

Chapter III, pp. 111 ff Chapter V, Chapter VI. ISBN 9810244703.

http://books.google.com/?id=mwLI4nQ0thQC&printsec=frontcover&dq=intitle:Intermediate+inti

tle:electromagnetic+intitle:theory.

47. ^ Tai L. Chow (2006). Electromagnetic theory. Sudbury MA: Jones and Bartlett. p. 333ff and

Chapter 3: pp. 89ff. ISBN 0-7637-3827-1.

http://books.google.com/?id=dpnpMhw1zo8C&pg=PA153&dq=isbn=0763738271.

48. ^ John Leonidas Volakis, Arindam Chatterjee & Leo C. Kempel (1998). Finite element method

for electromagnetics : antennas, microwave circuits, and scattering applications. New York:

Wiley IEEE. p. 79 ff. ISBN 0780334256.

http://books.google.com/?id=55q7HqnMZCsC&pg=PA79&dq=electromagnetism+%22boundary

+conditions%22.

49. ^ Bernard Friedman (1990). Principles and Techniques of Applied Mathematics. Mineola NY:

Dover Publications. ISBN 0486664449. http://www.amazon.com/Principles-Techniques-Applied-

Mathematics-

Friedman/dp/0486664449/ref=sr_1_1?ie=UTF8&s=books&qisbn=1207010487&sr=1-1.

50. ^ Littlejohn, Robert (Fall 2007). "Gaussian, SI and Other Systems of Units in Electromagnetic

Theory" (PDF). Physics 221A, University of California, Berkeley lecture notes.

http://bohr.physics.berkeley.edu/classes/221/0708/notes/emunits.pdf. Retrieved 2008-05-06.

51. ^ Albert Einstein (1905) On the electrodynamics of moving bodies

52. ^ Carver A. Mead (2002-08-07). Collective Electrodynamics: Quantum Foundations of

Electromagnetism. MIT Press. pp. 37–38. ISBN 9780262632607.

http://books.google.com/?id=GkDR4e2lo2MC&pg=PA37&dq=Riemann+Summerfeld.

53. ^ Frederic V. Hartemann (2002). High-field electrodynamics. CRC Press. p. 102.

ISBN 9780849323782.

http://books.google.com/?id=tIkflVrfkG0C&pg=PA102&dq=d%27Alembertian+covariant-

form+maxwell-lorentz.

54. ^ Oersted Medal Lecture David Hestenes (Am. J. Phys. 71 (2), February 2003, pp. 104--121)

Online:http://geocalc.clas.asu.edu/html/Oersted-ReformingTheLanguage.html p26

55. ^ M. Murray (5 September 2008). "Line Bundles. Honours 1996". University of Adelaide.

http://www.maths.adelaide.edu.au/michael.murray/line_bundles.pdf. Retrieved 2010-11-19.

56. ^ R. Bott (1985). "On some recent interactions between mathematics and physics". Canadian

Mathematical Bulletin 28 (2): 129–164.

Right-hand rule

The left-handed orientation is shown on the left, and the right-handed on the right.

Page 52: Basic Laws and Rules in Electrical Power

Use of right hand.

In mathematics and physics, the right-hand rule is a common mnemonic for understanding

notation conventions for vectors in 3 dimensions. It was invented for use in electromagnetism by

British physicist John Ambrose Fleming in the late 19th century.[1][2]

When choosing three vectors that must be at right angles to each other, there are two distinct

solutions, so when expressing this idea in mathematics, one must remove the ambiguity of which

solution is meant.

There are variations on the mnemonic depending on context, but all variations are related to the

one idea of choosing a convention.

Direction associated with an ordered pair of directions

One form of the right-hand rule is used in situations in which an ordered operation must be

performed on two vectors a and b that has a result which is a vector c perpendicular to both a and

b. The most common example is the vector cross product. The right-hand rule imposes the

following procedure for choosing one of the two directions.

With the thumb, index, and middle fingers at right angles to each other (with the index

finger pointed straight), the middle finger points in the direction of c when the thumb

represents a and the index finger represents b.

Other (equivalent) finger assignments are possible. For example, the first (index) finger can

represent a, the first vector in the product; the second (middle) finger, b, the second vector; and

the thumb, c, the product.[3]

Direction associated with a rotation

Page 53: Basic Laws and Rules in Electrical Power

Prediction of direction of field (B), given that the current I flows in the direction of the thumb.

A different form of the right-hand rule, sometimes called the right-hand grip rule, is used in

situations where a vector must be assigned to the rotation of a body, a magnetic field or a fluid.

Alternatively, when a rotation is specified by a vector, and it is necessary to understand the way

in which the rotation occurs, the right-hand grip rule is applicable.

This version of the rule is used in two complementary applications of Ampère's circuital law:

1. An electric current passes through a solenoid, resulting in a magnetic field. When you

wrap your right hand around the solenoid with your fingers in the direction of the

conventional current, your thumb points in the direction of the magnetic north pole.

2. An electric current passes through a straight wire. Here, the thumb points in the direction

of the conventional current (from positive to negative), and the fingers point in the

direction of the magnetic lines of flux.

The principle is also used to determine the direction of the torque vector. If you grip the

imaginary axis of rotation of the rotational force so that your fingers point in the direction of the

force, then the extended thumb points in the direction of the torque vector.

The right hand grip rule is a convention derived from the right-hand rule convention for vectors.

When applying the rule to current in a straight wire for example, the direction of the magnetic

field (counterclockwise instead of clockwise when viewed from the tip of the thumb) is a result

of this convention and not an underlying physical phenomenon

Applications

The first form of the rule is used to determine the direction of the cross product of two vectors.

This leads to widespread use in physics, wherever the cross product occurs. A list of physical

quantities whose directions are related by the right-hand rule is given below. (Some of these are

related only indirectly to cross products, and use the second form.)

The angular velocity of a rotating object and the rotational velocity of any point on the

object

A torque, the force that causes it, and the position of the point of application of the force

Page 54: Basic Laws and Rules in Electrical Power

A magnetic field, the position of the point where it is determined, and the electric current

(or change in electric flux) that causes it

A magnetic field in a coil of wire and the electric current in the wire

The force of a magnetic field on a charged particle, the magnetic field itself, and the

velocity of the object

The vorticity at any point in the field of flow of a fluid

The induced current from motion in a magnetic field (known as Fleming's right hand

rule)

Symmetry

Vector Right-hand Right-hand Right-hand Left-hand Left-hand Left-Hand

a, x or I Thumb Fingers or

palm First or Index Thumb

Fingers or

palm First or index

b, y or

B First or index Thumb

Fingers or

palm

Fingers or

palm First or index Thumb

c, z or F Fingers or

palm First or index Thumb First or index Thumb

Fingers or

palm

Ohm's law

V, I, and R, the parameters of Ohm's law.

Ohm's law states that the current through a conductor between two points is directly

proportional to the potential difference or voltage across the two points, and inversely

proportional to the resistance between them.[1]

The mathematical equation that describes this relationship is:[2]

where I is the current through the resistance in units of amperes, V is the potential difference

measured across the resistance in units of volts, and R is the resistance of the conductor in units

Page 55: Basic Laws and Rules in Electrical Power

of ohms. More specifically, Ohm's law states that the R in this relation is constant, independent

of the current.[3]

The law was named after the German physicist Georg Ohm, who, in a treatise published in 1827,

described measurements of applied voltage and current through simple electrical circuits

containing various lengths of wire. He presented a slightly more complex equation than the one

above (see History section below) to explain his experimental results. The above equation is the

modern form of Ohm's law.

In physics, the term Ohm's law is also used to refer to various generalizations of the law

originally formulated by Ohm. The simplest example of this is:

where J is the current density at a given location in a resistive material, E is the electric field at

that location, and σ is a material dependent parameter called the conductivity. This reformulation

of Ohm's law is due to Gustav Kirchhoff.[4]

Microscopic origins of Ohm's law

Drude Model electrons (shown here in blue) constantly bounce between heavier, stationary

crystal ions (shown in red).

The dependence of the current density on the applied electric field is essentially quantum

mechanical in nature; (see Classical and quantum conductivity.) A qualitative description leading

to Ohm's law can be based upon classical mechanics using the Drude model developed by Paul

Drude in 1900.[5][6]

The Drude model treats electrons (or other charge carriers) like pinballs bouncing between the

ions that make up the structure of the material. Electrons will be accelerated in the opposite

direction to the electric field by the average electric field at their location. With each collision,

though, the electron is deflected in a random direction with a velocity that is much larger than the

velocity gained by the electric field. The net result is that electrons take a tortuous path due to the

collisions, but generally drift in a direction opposing the electric field.

Page 56: Basic Laws and Rules in Electrical Power

The drift velocity then determines the electric current density and its relationship to E and is

independent of the collisions. Drude calculated the average drift velocity from p = −eEτ where p

is the average momentum, −e is the charge of the electron and τ is the average time between the

collisions. Since both the momentum and the current density are proportional to the drift

velocity, the current density becomes proportional to the applied electric field; this leads to

Ohm's law.

Hydraulic analogy

A hydraulic analogy is sometimes used to describe Ohm's Law. Water pressure, measured by

pascals (or PSI), is the analog of voltage because establishing a water pressure difference

between two points along a (horizontal) pipe causes water to flow. Water flow rate, as in liters

per second, is the analog of current, as in coulombs per second. Finally, flow restrictors — such

as apertures placed in pipes between points where the water pressure is measured — are the

analog of resistors. We say that the rate of water flow through an aperture restrictor is

proportional to the difference in water pressure across the restrictor. Similarly, the rate of flow of

electrical charge, that is, the electric current, through an electrical resistor is proportional to the

difference in voltage measured across the resistor.

Flow and pressure variables can be calculated in fluid flow network with the use of the hydraulic

ohm analogy.[7][8]

The method can be applied to both steady and transient flow situations. In the

linear laminar flow region, Poiseuille's law describes the hydraulic resistance of a pipe, but in the

turbulent flow region the pressure–flow relations become nonlinear.

Circuit analysis

Ohm's law triangle

In circuit analysis, three equivalent expressions of Ohm's law are used interchangeably:

Each equation is quoted by some sources as the defining relationship of Ohm's law,[2][9][10]

or all

three are quoted,[11]

or derived from a proportional form,[12]

or even just the two that do not

correspond to Ohm's original statement may sometimes be given.[13][14]

The interchangeability of the equation may be represented by a triangle, where V (voltage) is

placed on the top section, the I (current) is placed to the left section, and the R (resistance) is

Page 57: Basic Laws and Rules in Electrical Power

placed to the right. The line that divides the left and right sections indicate multiplication, and the

divider between the top and bottom sections indicates division (hence the division bar).

Resistive circuits

Resistors are circuit elements that impede the passage of electric charge in agreement with Ohm's

law, and are designed to have a specific resistance value R. In a schematic diagram the resistor is

shown as a zig-zag symbol. An element (resistor or conductor) that behaves according to Ohm's

law over some operating range is referred to as an ohmic device (or an ohmic resistor) because

Ohm's law and a single value for the resistance suffice to describe the behavior of the device

over that range.

Ohm's law holds for circuits containing only resistive elements (no capacitances or inductances)

for all forms of driving voltage or current, regardless of whether the driving voltage or current is

constant (DC) or time-varying such as AC. At any instant of time Ohm's law is valid for such

circuits.

Resistors which are in series or in parallel may be grouped together into a single "equivalent

resistance" in order to apply Ohm's law in analyzing the circuit. This application of Ohm's law is

illustrated with examples in "How To Analyze Resistive Circuits Using Ohm's Law" on

wikiHow.

Reactive circuits with time-varying signals

When reactive elements such as capacitors, inductors, or transmission lines are involved in a

circuit to which AC or time-varying voltage or current is applied, the relationship between

voltage and current becomes the solution to a differential equation, so Ohm's law (as defined

above) does not directly apply since that form contains only resistances having value R, not

complex impedances which may contain capacitance ("C") or inductance ("L").

Equations for time-invariant AC circuits take the same form as Ohm's law, however, the

variables are generalized to complex numbers and the current and voltage waveforms are

complex exponentials.[15]

In this approach, a voltage or current waveform takes the form Aest, where t is time, s is a

complex parameter, and A is a complex scalar. In any linear time-invariant system, all of the

currents and voltages can be expressed with the same s parameter as the input to the system,

allowing the time-varying complex exponential term to be canceled out and the system described

algebraically in terms of the complex scalars in the current and voltage waveforms.

The complex generalization of resistance is impedance, usually denoted Z; it can be shown that

for an inductor,

and for a capacitor,

Page 58: Basic Laws and Rules in Electrical Power

We can now write,

where V and I are the complex scalars in the voltage and current respectively and Z is the

complex impedance.

This form of Ohm's law, with Z taking the place of R, generalizes the simpler form. When Z is

complex, only the real part is responsible for dissipating heat.

In the general AC circuit, Z varies strongly with the frequency parameter s, and so also will the

relationship between voltage and current.

For the common case of a steady sinusoid, the s parameter is taken to be jω, corresponding to a

complex sinusoid Aejωt. The real parts of such complex current and voltage waveforms describe

the actual sinusoidal currents and voltages in a circuit, which can be in different phases due to

the different complex scalars.

Linear approximations

Ohm's law is one of the basic equations used in the analysis of electrical circuits. It applies to

both metal conductors and circuit components (resistors) specifically made for this behaviour.

Both are ubiquitous in electrical engineering. Materials and components that obey Ohm's law are

described as "ohmic" [16]

which means they produce the same value for resistance (R = V/I)

regardless of the value of V or I which is applied and whether the applied voltage or current is

DC (direct current) of either positive or negative polarity or AC (alternating current).

In a true ohmic device, the same value of resistance will be calculated from R = V/I regardless of

the value of the applied voltage V. That is, the ratio of V/I is constant, and when current is

plotted as a function of voltage the curve is linear (a straight line). If voltage is forced to some

value V, then that voltage V divided by measured current I will equal R. Or if the current is

forced to some value I, then the measured voltage V divided by that current I is also R. Since the

plot of I versus V is a straight line, then it is also true that for any set of two different voltages V1

and V2 applied across a given device of resistance R, producing currents I1 = V1/R and I2 = V2/R,

that the ratio (V1-V2)/(I1-I2) is also a constant equal to R. The operator "delta" (Δ) is used to

represent a difference in a quantity, so we can write ΔV = V1-V2 and ΔI = I1-I2. Summarizing, for

any truly ohmic device having resistance R, V/I = ΔV/ΔI = R for any applied voltage or current

or for the difference between any set of applied voltages or currents.

Page 59: Basic Laws and Rules in Electrical Power

Plot of I–V curve of an ideal p-n junction diode at 1μA reverse leakage current. Failure of the

device to follow Ohm's law is clearly shown since the curve is not a straight line.

There are, however, components of electrical circuits which do not obey Ohm's law; that is, their

relationship between current and voltage (their I–V curve) is nonlinear. An example is the p-n

junction diode (curve at right). As seen in the figure, the current does not increase linearly with

applied voltage for a diode. One can determine a value of current (I) for a given value of applied

voltage (V) from the curve, but not from Ohm's law, since the value of "resistance" is not

constant as a function of applied voltage. Further, the current only increases significantly if the

applied voltage is positive, not negative. The ratio V/I for some point along the nonlinear curve is

sometimes called the static, or chordal, or DC, resistance[17][18]

, but as seen in the figure the

value of total V over total I varies depending on the particular point along the nonlinear curve

which is chosen. This means the "DC resistance" V/I at some point on the curve is not the same

as what would be determined by applying an AC signal having peak amplitude ΔV volts or ΔI

amps centered at that same point along the curve and measuring ΔV/ΔI. However, in some diode

applications, the AC signal applied to the device is small and it is possible to analyze the circuit

in terms of the dynamic, small-signal, or incremental resistance, defined as the one over the

slope of the V–I curve at the average value (DC operating point) of the voltage (that is, one over

the derivative of current with respect to voltage). For sufficiently small signals, the dynamic

resistance allows the Ohm's law small signal resistance to be calculated as approximately one

over the slope of a line drawn tangentially to the V-I curve at the DC operating point.[19]

Temperature effects

Ohm's law has sometimes been stated as, "for a conductor in a given state, the electromotive

force is proportional to the current produced." That is, that the resistance, the ratio of the applied

electromotive force (or voltage) to the current, "does not vary with the current strength ." The

qualifier "in a given state" is usually interpreted as meaning "at a constant temperature," since

the resistivity of materials is usually temperature dependent. Because the conduction of current is

related to Joule heating of the conducting body, according to Joule's first law, the temperature of

a conducting body may change when it carries a current. The dependence of resistance on

temperature therefore makes resistance depend upon the current in a typical experimental setup,

making the law in this form difficult to directly verify. Maxwell and others worked out several

methods to test the law experimentally in 1876, controlling for heating effects.[20]

Relation to heat conductions

Page 60: Basic Laws and Rules in Electrical Power

Ohm's principle predicts the flow of electrical charge (i.e. current) in electrical conductors when

subjected to the influence of voltage differences; Jean-Baptiste-Joseph Fourier's principle

predicts the flow of heat in heat conductors when subjected to the influence of temperature

differences.

The same equation describes both phenomena, the equation's variables taking on different

meanings in the two cases. Specifically, solving a heat conduction (Fourier) problem with

temperature (the driving "force") and flux of heat (the rate of flow of the driven "quantity", i.e.

heat energy) variables also solves an analogous electrical conduction (Ohm) problem having

electric potential (the driving "force") and electric current (the rate of flow of the driven

"quantity", i.e. charge) variables.

The basis of Fourier's work was his clear conception and definition of thermal conductivity. He

assumed that, all else being the same, the flux of heat is strictly proportional to the gradient of

temperature. Although undoubtedly true for small temperature gradients, strictly proportional

behavior will be lost when real materials (e.g. ones having a thermal conductivity that is a

function of temperature) are subjected to large temperature gradients.

A similar assumption is made in the statement of Ohm's law: other things being alike, the

strength of the current at each point is proportional to the gradient of electric potential. The

accuracy of the assumption that flow is proportional to the gradient is more readily tested, using

modern measurement methods, for the electrical case than for the heat case.

Other versions of Ohm's law

Ohm's law, in the form above, is an extremely useful equation in the field of electrical/electronic

engineering because it describes how voltage, current and resistance are interrelated on a

"macroscopic" level, that is, commonly, as circuit elements in an electrical circuit. Physicists

who study the electrical properties of matter at the microscopic level use a closely related and

more general vector equation, sometimes also referred to as Ohm's law, having variables that are

closely related to the V, I, and R scalar variables of Ohm's law, but are each functions of position

within the conductor. Physicists often use this continuum form of Ohm's Law:[21]

where "E" is the electric field vector with units of volts per meter (analogous to "V" of Ohm's

law which has units of volts), "J" is the current density vector with units of amperes per unit area

(analogous to "I" of Ohm's law which has units of amperes), and "ρ" (Greek "rho") is the

resistivity with units of ohm·meters (analogous to "R" of Ohm's law which has units of ohms).

The above equation is sometimes written[22]

as J = σE where "σ" is the conductivity which is the

reciprocal of ρ.

Page 61: Basic Laws and Rules in Electrical Power

Current flowing through a uniform cylindrical conductor (such as a round wire) with a uniform

field applied.

The potential difference between two points is defined as:[23]

with the element of path along the integration of electric field vector E. If the applied E field

is uniform and oriented along the length of the conductor as shown in the figure, then defining

the voltage V in the usual convention of being opposite in direction to the field (see figure), and

with the understanding that the voltage V is measured differentially across the length of the

conductor allowing us to drop the Δ symbol, the above vector equation reduces to the scalar

equation:

Since the E field is uniform in the direction of wire length, for a conductor having uniformly

consistent resistivity ρ, the current density J will also be uniform in any cross-sectional area and

oriented in the direction of wire length, so we may write:[24]

Substituting the above 2 results (for E and J respectively) into the continuum form shown at the

beginning of this section:

The electrical resistance of a uniform conductor is given in terms of resistivity by:[24]

Page 62: Basic Laws and Rules in Electrical Power

where l is the length of the conductor in SI units of meters, a is the cross-sectional area (for a

round wire a = πr2 if r is radius) in units of meters squared, and ρ is the resistivity in units of

ohm·meters.

After substitution of R from the above equation into the equation preceding it, the continuum

form of Ohm's law for a uniform field (and uniform current density) oriented along the length of

the conductor reduces to the more familiar form:

A perfect crystal lattice, with low enough thermal motion and no deviations from periodic

structure, would have no resistivity,[25]

but a real metal has crystallographic defects, impurities,

multiple isotopes, and thermal motion of the atoms. Electrons scatter from all of these, resulting

in resistance to their flow.

The more complex generalized forms of Ohm's law are important to condensed matter physics,

which studies the properties of matter and, in particular, its electronic structure. In broad terms,

they fall under the topic of constitutive equations and the theory of transport coefficients.

Magnetic effects

The continuum form of the equation is only valid in the reference frame of the conducting

material. If the material is moving at velocity v relative to a magnetic field B, a term must be

added as follows:

See Lorentz force for more on this and Hall effect for some other implications of a magnetic

field. This equation is not a modification to Ohm's law. Rather, it is analogous in circuit analysis

terms to taking into account inductance as well as resistance.

History

In January 1781, before Georg Ohm's work, Henry Cavendish experimented with Leyden jars

and glass tubes of varying diameter and length filled with salt solution. He measured the current

by noting how strong a shock he felt as he completed the circuit with his body. Cavendish wrote

that the "velocity" (current) varied directly as the "degree of electrification" (voltage). He did not

communicate his results to other scientists at the time,[26]

and his results were unknown until

Maxwell published them in 1879.[27]

Ohm did his work on resistance in the years 1825 and 1826, and published his results in 1827 as

the book Die galvanische Kette, mathematisch bearbeitet (The galvanic Circuit investigated

mathematically).[28]

He drew considerable inspiration from Fourier's work on heat conduction in

the theoretical explanation of his work. For experiments, he initially used voltaic piles, but later

used a thermocouple as this provided a more stable voltage source in terms of internal resistance

and constant potential difference. He used a galvanometer to measure current, and knew that the

Page 63: Basic Laws and Rules in Electrical Power

voltage between the thermocouple terminals was proportional to the junction temperature. He

then added test wires of varying length, diameter, and material to complete the circuit. He found

that his data could be modeled through the equation

where x was the reading from the galvanometer, l was the length of the test conductor, a

depended only on the thermocouple junction temperature, and b was a constant of the entire

setup. From this, Ohm determined his law of proportionality and published his results.

Ohm's law was probably the most important of the early quantitative descriptions of the physics

of electricity. We consider it almost obvious today. When Ohm first published his work, this was

not the case; critics reacted to his treatment of the subject with hostility. They called his work a

"web of naked fancies"[29]

and the German Minister of Education proclaimed that "a professor

who preached such heresies was unworthy to teach science."[30]

The prevailing scientific

philosophy in Germany at the time asserted that experiments need not be performed to develop

an understanding of nature because nature is so well ordered, and that scientific truths may be

deduced through reasoning alone. Also, Ohm's brother Martin, a mathematician, was battling the

German educational system. These factors hindered the acceptance of Ohm's work, and his work

did not become widely accepted until the 1840s. Fortunately, Ohm received recognition for his

contributions to science well before he died.

In the 1850s, Ohm's law was known as such, and was widely considered proved, and alternatives

such as "Barlow's law" discredited, in terms of real applications to telegraph system design, as

discussed by Samuel F. B. Morse in 1855.[31]

While the old term for electrical conductance, the mho (the inverse of the resistance unit ohm), is

still used, a new name, the siemens, was adopted in 1971, honoring Ernst Werner von Siemens.

The siemens is preferred in formal papers.

In the 1920s, it was discovered that the current through an ideal resistor actually has statistical

fluctuations, which depend on temperature, even when voltage and resistance are exactly

constant; this fluctuation, now known as Johnson–Nyquist noise, is due to the discrete nature of

charge. This thermal effect implies that measurements of current and voltage that are taken over

sufficiently short periods of time will yield ratios of V/I that fluctuate from the value of R

implied by the time average or ensemble average of the measured current; Ohm's law remains

correct for the average current, in the case of ordinary resistive materials.

Ohm's work long preceded Maxwell's equations and any understanding of frequency-dependent

effects in AC circuits. Modern developments in electromagnetic theory and circuit theory do not

contradict Ohm's law when they are evaluated within the appropriate limits.

References

Page 64: Basic Laws and Rules in Electrical Power

1. ^ Consoliver, Earl L., and Mitchell, Grover I. (1920). Automotive ignition systems. McGraw-Hill.

p. 4.

http://books.google.com/?id=_dYNAAAAYAAJ&pg=PA4&dq=ohm%27s+law+current+proport

ional+voltage+resistance.

2. ^ a b Robert A. Millikan and E. S. Bishop (1917). Elements of Electricity. American Technical

Society. p. 54.

http://books.google.com/?id=dZM3AAAAMAAJ&pg=PA54&dq=%22Ohm%27s+law%22++cur

rent+directly+proportional.

3. ^ Oliver Heaviside (1894). Electrical papers. 1. Macmillan and Co. p. 283. ISBN 0821828401.

http://books.google.com/?id=lKV-

AAAAMAAJ&pg=PA284&dq=ohm%27s+law+constant+ratio&q=ohm's%20law%20constant%

20ratio.

4. ^ Olivier Darrigol, Electrodynamics from Ampère to Einstein, p.70, Oxford University Press,

2000 ISBN 0198505949.

5. ^ Drude, Paul (1900). "Zur Elektronentheorie der metalle". Annalen der Physik 306 (3): 566.

doi:10.1002/andp.19003060312. http://www3.interscience.wiley.com/cgi-

bin/fulltext/112485959/PDFSTART.

6. ^ Drude, Paul (1900). "Zur Elektronentheorie der Metalle; II. Teil. Galvanomagnetische und

thermomagnetische Effecte". Annalen der Physik 308 (11): 369. doi:10.1002/andp.19003081102.

http://www3.interscience.wiley.com/cgi-bin/fulltext/112485893/PDFSTART.

7. ^ A. Akers, M. Gassman, & R. Smith (2006). Hydraulic Power System Analysis. New York:

Taylor & Francis. Chapter 13. ISBN 0-8247-9956-9.

http://books.google.com/?id=Uo9gpXeUoKAC&pg=PA299&dq=ohm+intitle:Hydraulic+intitle:P

ower+intitle:System+intitle:Analysis.

8. ^ A. Esposito, "A Simplified Method for Analyzing Circuits by Analogy", Machine Design,

October 1969, pp. 173–177.

9. ^ James William Nilsson and Susan A. Riedel (2008). Electric circuits. Prentice Hall. p. 29.

ISBN 9780131989252.

http://books.google.com/?id=sxmM8RFL99wC&pg=PA29&dq=%22Ohm%27s+law+expresses+

the+voltage%22++%22V+%3D+iR%22.

10. ^ Alvin M. Halpern and Erich Erlbach (1998). Schaum's outline of theory and problems of

beginning physics II. McGraw-Hill Professional. p. 140. ISBN 9780070257078.

http://books.google.com/?id=vN2chIay624C&pg=PA140&dq=%22Ohm%27s+law+that+R%3D

+V/I+is+a+constant%22.

11. ^ Dale R. Patrick and Stephen W. Fardo (1999). Understanding DC circuits. Newnes. p. 96.

ISBN 9780750671101.

http://books.google.com/?id=wyC5SFtZskMC&pg=PA96&dq=%22Ohm%27s+law%22+%22R+

%3D%22+%22V+%3D%22+%22I+%3D%22.

12. ^ Thomas O'Conor Sloane (1909). Elementary electrical calculations. D. Van Nostrand Co.

p. 41.

http://books.google.com/?id=e89IAAAAIAAJ&pg=PA41&dq=R%3D+%22Ohm%27s+law%22

+proportional.

13. ^ Linnaeus Cumming (1902). Electricity treated experimentally for the use of schools and

students. Longman's Green and Co. p. 220.

http://books.google.com/?id=3iUJAAAAIAAJ&pg=PA220&dq=V%3DIR+%22Ohm%27s+law

%22.

14. ^ Benjamin Stein (1997). Building technology (2nd ed.). John Wiley and Sons. p. 169.

ISBN 9780471593195.

http://books.google.com/?id=J_RSbj_KzAQC&pg=PA169&dq=%22Ohm%27s+law+that+V%3

D%22.

Page 65: Basic Laws and Rules in Electrical Power

15. ^ Rajendra Prasad (2006). Fundamentals of Electrical Engineering. Prentice-Hall of India.

ISBN 9788120327290.

http://books.google.com/?id=nsmcbzOJU3kC&pg=PA140&dq=ohm%27s-

law+complex+exponentials.

16. ^ Hughes, E, Electrical Technology, pp10, Longmans, 1969.

17. ^ Forbes T. Brown (2006). Engineering System Dynamics. CRC Press. p. 43.

ISBN 9780849396489.

http://books.google.com/?id=UzqX4j9VZWcC&pg=PA43&dq=%22chordal+resistance%22.

18. ^ Kenneth L. Kaiser (2004). Electromagnetic Compatibility Handbook. CRC Press. pp. 13–52.

ISBN 9780849320873.

http://books.google.com/?id=nZzOAsroBIEC&pg=PT1031&dq=%22static+resistance%22+%22

dynamic+resistance%22+nonlinear.

19. ^ Horowitz, Paul; Winfield Hill (1989). The Art of Electronics (2nd ed.). Cambridge University

Press. p. 13. ISBN 0-521-37095-7.

http://books.google.com/?id=bkOMDgwFA28C&pg=PA13&dq=small-

signal+%22dynamic+resistance%22.

20. ^ Normal Lockyer, ed (September 21, 1876). "Reports". Nature (Macmillan Journals Ltd) 14:

452. http://books.google.com/?id=-8gKAAAAYAAJ&pg=PA452&dq=ohm%27s-

law+temperature&q=ohm's-law%20temperature.

21. ^ Lerner, Lawrence S. (1977). Physics for scientists and engineers. Jones & Bartlett. p. 736.

ISBN 9780763704605. http://books.google.com/?id=Nv5GAyAdijoC&pg=PA736.

22. ^ Seymour J, Physical Electronics, Pitman, 1972, pp 53–54

23. ^ Lerner L, Physics for scientists and engineers, Jones & Bartlett, 1997, pp. 685–686

24. ^ a b Lerner L, Physics for scientists and engineers, Jones & Bartlett, 1997, pp. 732–733

25. ^ Seymour J, Physical Electronics, pp 48–49, Pitman, 1972

26. ^ "Electricity". Encyclopedia Britannica. 1911. http://www.1911encyclopedia.org/Electricity.

27. ^ Sanford P. Bordeau (1982) Volts to Hertz...the Rise of Electricity. Burgess Publishing

Company, Minneapolis, MN. pp.86–107, ISBN 0-8087-4908-0

28. ^ G. S. Ohm (1827). Die galvanische Kette, mathematisch bearbeitet. Berlin: T. H. Riemann.

http://www.ohm-hochschule.de/bib/textarchiv/Ohm.Die_galvanische_Kette.pdf.

29. ^ Davies, B, "A web of naked fancies?", Physics Education 15 57–61, Institute of Physics, Issue

1, Jan 1980 [1]

30. ^ Hart, IB, Makers of Science, London, Oxford University Press, 1923. p. 243. [2]

31. ^ Taliaferro Preston (1855). Shaffner's Telegraph Companion: Devoted to the Science and Art of

the Morse Telegraph. Vol.2. Pudney & Russell.

http://books.google.com/?id=TDEOAAAAYAAJ&pg=RA1-PA43&dq=ohm%27s-law+date:0-

1860.

Kirchhoff's circuit laws

Kirchhoff's circuit laws are two equalities that deal with the conservation of charge and energy

in electrical circuits, and were first described in 1845 by Gustav Kirchhoff. Widely used in

electrical engineering, they are also called Kirchhoff's rules or simply Kirchhoff's laws (see also

Kirchhoff's laws for other meanings of that term).

Both circuit rules can be directly derived from Maxwell's equations, but Kirchhoff preceded

Maxwell and instead generalized work by Georg Ohm.

Page 66: Basic Laws and Rules in Electrical Power

Kirchhoff's current law (KCL)

The current entering any junction is equal to the current leaving that junction. i1 + i4 = i2 + i3

This law is also called Kirchhoff's point rule, Kirchhoff's junction rule (or nodal rule), and

Kirchhoff's first rule.

The principle of conservation of electric charge implies that:

At any node (junction) in an electrical circuit, the sum of currents flowing into that node

is equal to the sum of currents flowing out of that node.

or

The algebraic sum of currents in a network of conductors meeting at a point is zero.

(Assuming that current entering the junction is taken as positive and current leaving the

junction is taken as negative).

Recalling that current is a signed (positive or negative) quantity reflecting direction towards or

away from a node, this principle can be stated as:

n is the total number of branches with currents flowing towards or away from the node.

This formula is also valid for complex currents:

The law is based on the conservation of charge whereby the charge (measured in coulombs) is

the product of the current (in amperes) and the time (which is measured in seconds).

Changing charge density

Page 67: Basic Laws and Rules in Electrical Power

Physically speaking, the restriction regarding the "capacitor plate" means that Kirchhoff's current

law is only valid if the charge density remains constant in the point that it is applied to. This is

normally not a problem because of the strength of electrostatic forces: the charge buildup would

cause repulsive forces to disperse the charges.

However, a charge build-up can occur in a capacitor, where the charge is typically spread over

wide parallel plates, with a physical break in the circuit that prevents the positive and negative

charge accumulations over the two plates from coming together and cancelling. In this case, the

sum of the currents flowing into one plate of the capacitor is not zero, but rather is equal to the

rate of charge accumulation. However, if the displacement current dD/dt is included, Kirchhoff's

current law once again holds. (This is really only required if one wants to apply the current law

to a point on a capacitor plate. In circuit analyses, however, the capacitor as a whole is typically

treated as a unit, in which case the ordinary current law holds since exactly the current that enters

the capacitor on the one side leaves it on the other side.)

More technically, Kirchhoff's current law can be found by taking the divergence of Ampère's law

with Maxwell's correction and combining with Gauss's law, yielding:

This is simply the charge conservation equation (in integral form, it says that the current flowing

out of a closed surface is equal to the rate of loss of charge within the enclosed volume

(Divergence theorem). Kirchhoff's current law is equivalent to the statement that the divergence

of the current is zero, true for time-invariant ρ, or always true if the displacement current is

included with J.

Uses

A matrix version of Kirchhoff's current law is the basis of most circuit simulation software, such

as SPICE.

Kirchhoff's voltage law (KVL)

The sum of all the voltages around the loop is equal to zero. v1 + v2 + v3 - v4 = 0

Page 68: Basic Laws and Rules in Electrical Power

This law is also called Kirchhoff's second law, Kirchhoff's loop (or mesh) rule, and

Kirchhoff's second rule.

The principle of conservation of energy implies that

The directed sum of the electrical potential differences (voltage) around any closed

circuit is zero.

or

More simply, the sum of the emfs in any closed loop is equivalent to the sum of the

potential drops in that loop.

or

The algebraic sum of the products of the resistances of the conductors and the currents in

them in a closed loop is equal to the total emf available in that loop.

Similarly to KCL, it can be stated as:

Here, n is the total number of voltages measured. The voltages may also be complex:

This law is based on the conservation of "energy given/taken by potential field" (not including

energy taken by dissipation). Given a voltage potential, a charge which has completed a closed

loop doesn't gain or lose energy as it has gone back to initial potential level.

This law holds true even when resistance (which causes dissipation of energy) is present in a

circuit. The validity of this law in this case can be understood if one realizes that a charge in fact

doesn't go back to its starting point, due to dissipation of energy. A charge will just terminate at

the negative terminal, instead of positive terminal. This means all the energy given by the

potential difference has been fully consumed by resistance which in turn loses the energy as heat

dissipation.

To summarize, Kirchhoff's voltage law has nothing to do with gain or loss of energy by

electronic components (resistors, capacitors, etc). It is a law referring to the potential field

generated by voltage sources. In this potential field, regardless of what electronic components

are present, the gain or loss in "energy given by the potential field" must be zero when a charge

completes a closed loop.

Electric field and electric potential

Page 69: Basic Laws and Rules in Electrical Power

Kirchhoff's voltage law could be viewed as a consequence of the principle of conservation of

energy. Otherwise, it would be possible to build a perpetual motion machine that passed a

current in a circle around the circuit.

Considering that electric potential is defined as a line integral over an electric field, Kirchhoff's

voltage law can be expressed equivalently as

which states that the line integral of the electric field around closed loop C is zero.

In order to return to the more special form, this integral can be "cut in pieces" in order to get the

voltage at specific components.

Limitations

This is a simplification of Faraday's law of induction for the special case where there is no

fluctuating magnetic field linking the closed loop. Therefore, it practically suffices for explaining

circuits containing only resistors and capacitors.

In the presence of a changing magnetic field the electric field is not conservative and it cannot

therefore define a pure scalar potential—the line integral of the electric field around the circuit is

not zero. This is because energy is being transferred from the magnetic field to the current (or

vice versa). In order to "fix" Kirchhoff's voltage law for circuits containing inductors, an

effective potential drop, or electromotive force (emf), is associated with each inductance of the

circuit, exactly equal to the amount by which the line integral of the electric field is not zero by

Faraday's law of induction.

References

Paul, Clayton R. (2001). Fundamentals of Electric Circuit Analysis. John Wiley & Sons.

ISBN 0-471-37195-5.

Serway, Raymond A.; Jewett, John W. (2004). Physics for Scientists and Engineers (6th

ed.). Brooks/Cole. ISBN 0-534-40842-7.

Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity, Magnetism, Light,

and Elementary Modern Physics (5th ed.). W. H. Freeman. ISBN 0-7167-0810-8.

External links

http://www.sasked.gov.sk.ca/docs/physics/u3c13phy.html

MIT video lecture on the KVL and KCL methods

Page 70: Basic Laws and Rules in Electrical Power

TRANSFORMER FUNDAMENTALS

A transformer is a device that transfers electrical energy from one circuit to another through

inductively coupled conductors—the transformer's coils. A varying current in the first or primary

winding creates a varying magnetic flux in the transformer's core and thus a varying magnetic

field through the secondary winding. This varying magnetic field induces a varying

electromotive force (EMF) or "voltage" in the secondary winding. This effect is called mutual

induction.

If a load is connected to the secondary, an electric current will flow in the secondary winding and

electrical energy will be transferred from the primary circuit through the transformer to the load.

In an ideal transformer, the induced voltage in the secondary winding (Vs) is in proportion to the

primary voltage (Vp), and is given by the ratio of the number of turns in the secondary (Ns) to the

number of turns in the primary (Np) as follows:

By appropriate selection of the ratio of turns, a transformer thus allows an alternating current

(AC) voltage to be "stepped up" by making Ns greater than Np, or "stepped down" by making Ns

less than Np.

Basic principles a simple single phase (1υ) Transformer consists of two electrical conductors

called the primary coil and the secondary coil. The primary is fed with a varying (alternating or

pulsed direct current) electric current which creates a varying magnetic field around the

conductor. According to the principle of mutual inductance, the secondary, which is placed in

this varying magnetic field, will develop a potential difference called an electromotive force or

EMF. If the ends of the secondary are connected together to form an electrical circuit, this EMF

will cause a current to flow in the secondary. Thus, some of the electrical power fed into the

primary is delivered to the secondary.

In practical Transformers, the primary and secondary conductors are coils of wire because a coil

creates a denser magnetic field (higher magnetic flux) than a straight conductor.

A Transformer winding should never be energised from a constant DC voltage source, as this

would cause a large direct current to flow. In such a situation, in an ideal Transformer, the

current would rise indefinitely as a linear function of time. In practice, the series resistance of the

winding limits the amount of current that can flow, until the Transformer either reaches thermal

equilibrium or is destroyed.

Transformers alone cannot do the following: Convert DC to AC or vice versa Change the voltage

Page 71: Basic Laws and Rules in Electrical Power

or current of DC Change the frequency (the "cycles") of AC. However, transformers are

components of the systems that perform all these functions.

Electrical laws consider the following two laws: According to the law of conservation of energy,

the power delivered by a Transformer cannot exceed the power fed into it. The power dissipated

in a load at any instant is equal to the product of the voltage across it and the current passing

through it (see also Ohm's law). It follows from the above two laws that a Transformer is not a

power amplifier. If the Transformer is used to change transfer power from one voltage level to

another, the magnitudes of the currents in the two windings must also be different, in inverse

ratio to the voltages. Thus if current were to be brought down by the Transformer, voltage would

go up. If voltage were to be brought down by the Transformer, current would go up. For

example, suppose 50 watts is fed into a transformer with a ratio of 25:2.

P = I*E (Power = Current * Electromotiveforce) 50 W = 25 A * 2 V in the primary circuit

Now with Transformer change: 50 W = 2 A * 25 V in the secondary circuit.

The high-current low-voltage windings have fewer turns of wire. The high-voltage, low-current

windings have more turns of wire.

The electromotive force (EMF) developed in the secondary is proportional to the ratio of the

number of turns in the secondary coil to the number of turns in the primary coil. Neglecting all

leakage flux, an ideal transformer follows the equation: Vp/Vs = Np/Ns

Where Vp is the voltage in the primary coil, Vs is the voltage in the secondary coil, Np is the

number of turns of wire on the primary coil, and Ns is the number of turns of wire on the

secondary coil. This leads to the most common use of the transformer: to convert power at one

voltage to power at a different voltage.

Again, neglecting leakage flux, the relationship between voltage, number of turns, magnetic flux

intensity and core area is given by: E = 4.44 * F * n * a * b Where E is the sinusoidal root mean

square (RMS) voltage of the winding, F is the frequency in hertz, n is the number of turns of

wire, a is the area of the core (square units) and b is magnetic flux density in webers per square

unit. The value 4.44 collects a number of constants required by the system of units.

History

Discovery

Page 72: Basic Laws and Rules in Electrical Power

Faraday's experiment with induction between coils of wire

[1]

The phenomenon of electromagnetic induction was discovered independently by Michael

Faraday and Joseph Henry in 1831. However, Faraday was the first to publish the results of his

experiments and thus receive credit for the discovery.[2]

The relationship between electromotive

force (EMF) or "voltage" and magnetic flux was formalized in an equation now referred to as

"Faraday's law of induction":

.

where is the magnitude of the EMF in volts and ΦB is the magnetic flux through the circuit (in

webers).[3]

Faraday performed the first experiments on induction between coils of wire, including winding a

pair of coils around an iron ring, thus creating the first toroidal closed-core transformer.[4]

Induction coils

The first type of transformer to see wide use was the induction coil, invented by Rev. Nicholas

Callan of Maynooth College, Ireland in 1836. He was one of the first researchers to realize that

the more turns the secondary winding has in relation to the primary winding, the larger is the

increase in EMF. Induction coils evolved from scientists' and inventors' efforts to get higher

voltages from batteries. Since batteries produce direct current (DC) rather than alternating

current (AC), induction coils relied upon vibrating electrical contacts that regularly interrupted

the current in the primary to create the flux changes necessary for induction. Between the 1830s

and the 1870s, efforts to build better induction coils, mostly by trial and error, slowly revealed

the basic principles of transformers.

In 1876, Russian engineer Pavel Yablochkov invented a lighting system based on a set of

induction coils where the primary windings were connected to a source of alternating current and

the secondary windings could be connected to several "electric candles" (arc lamps) of his own

design.[5][6]

The coils Yablochkov employed functioned essentially as transformers.[5]

Page 73: Basic Laws and Rules in Electrical Power

In 1878, the Ganz Company in Hungary began manufacturing equipment for electric lighting

and, by 1883, had installed over fifty systems in Austria-Hungary. Their systems used alternating

current exclusively and included those comprising both arc and incandescent lamps, along with

generators and other equipment.[7]

Lucien Gaulard and John Dixon Gibbs first exhibited a device with an open iron core called a

"secondary generator" in London in 1882, then sold the idea to the Westinghouse company in the

United States.[8]

They also exhibited the invention in Turin, Italy in 1884, where it was adopted

for an electric lighting system.[9]

However, the efficiency of their open-core bipolar apparatus

remained very low.[10]

Induction coils with open magnetic circuits are inefficient for transfer of power to loads. Until

about 1880, the paradigm for AC power transmission from a high voltage supply to a low

voltage load was a series circuit. Open-core transformers with a ratio near 1:1 were connected

with their primaries in series to allow use of a high voltage for transmission while presenting a

low voltage to the lamps. The inherent flaw in this method was that turning off a single lamp

affected the voltage supplied to all others on the same circuit. Many adjustable transformer

designs were introduced to compensate for this problematic characteristic of the series circuit,

including those employing methods of adjusting the core or bypassing the magnetic flux around

part of a coil.[11]

Efficient, practical transformer designs did not appear until the 1880s, but within a decade the

transformer would be instrumental in the "War of Currents", and in seeing AC distribution

systems triumph over their DC counterparts, a position in which they have remained dominant

ever since.[12]

Closed-core lighting transformers

Drawing of Ganz Company's 1885 prototype. Capacity: 1400 VA, frequency: 40 Hz, voltage

ratio: 120/72 V

Page 74: Basic Laws and Rules in Electrical Power

Prototypes of the world's first high-efficiency transformers. They were built by the Z.B.D. team

on 16th September 1884.[13]

In the autumn of 1884[14]

, Ganz Company engineers Károly Zipernowsky, Ottó Bláthy and

Miksa Déri had determined that open-core devices were impracticable, as they were incapable of

reliably regulating voltage. In their joint patent application for the "Z.B.D." transformers, they

described two designs with closed magnetic circuits: the "closed-core" and "shell-core"

transformers. In the closed-core, the primary and secondary windings were wound around a

closed iron ring; in the shell-core, the windings were passed through the iron core. In both

designs, the magnetic flux linking the primary and secondary windings traveled almost entirely

within the iron core, with no intentional path through air. The new Z.B.D. transformers reached

98 percent efficiency, which was 3.4 times higher than the open core bipolar devices of Gaulard

and Gibs.[15]

When employed in parallel connected electric distribution systems, closed-core

transformers finally made it technically and economically feasible to provide electric power for

lighting in homes, businesses and public spaces.[16][17]

Bláthy had suggested the use of closed-

cores, Zipernowsky the use of shunt connections, and Déri had performed the experiments;[18]

Bláthy also discovered the transformer formula, Vs/Vp = Ns/Np.[citation needed]

The vast majority of

transformers in use today rely on the basic principles discovered by the three engineers. They

also reportedly popularized the word "transformer" to describe a device for altering the EMF of

an electric current,[16][19]

although the term had already been in use by 1882.[20][21]

In 1886, the

Ganz Company installed the world's first power station that used AC generators to power a

parallel-connected common electrical network, the steam-powered Rome-Cerchi power plant.[22]

Stanley's 1886 design for adjustable gap open-core induction coils

[23]

Although George Westinghouse had bought Gaulard and Gibbs' patents in 1885, the Edison

Electric Light Company held an option on the U.S. rights for the Z.B.D. transformers, requiring

Westinghouse to pursue alternative designs on the same principles. He assigned to William

Stanley the task of developing a device for commercial use in United States.[24]

Stanley's first

patented design was for induction coils with single cores of soft iron and adjustable gaps to

Page 75: Basic Laws and Rules in Electrical Power

regulate the EMF present in the secondary winding. (See drawing at left.)[23]

This design was

first used commercially in the U.S. in 1886.[12]

But Westinghouse soon had his team working on

a design whose core comprised a stack of thin "E-shaped" iron plates, separated individually or

in pairs by thin sheets of paper or other insulating material. Prewound copper coils could then be

slid into place, and straight iron plates laid in to create a closed magnetic circuit. Westinghouse

applied for a patent for the new design in December 1886; it was granted in July 1887.[18][25]

Other early transformers

In 1889, Russian-born engineer Mikhail Dolivo-Dobrovolsky developed the first three-phase

transformer at the Allgemeine Elektricitäts-Gesellschaft ("General Electricity Company") in

Germany.[26]

In 1891, Nikola Tesla invented the Tesla coil, an air-cored, dual-tuned resonant transformer for

generating very high voltages at high frequency.[27][28]

Audio frequency transformers ("repeating coils") were used by early experimenters in the

development of the telephone.[citation needed]

Basic principles

The transformer is based on two principles: first, that an electric current can produce a magnetic

field (electromagnetism), and, second that a changing magnetic field within a coil of wire

induces a voltage across the ends of the coil (electromagnetic induction). Changing the current in

the primary coil changes the magnetic flux that is developed. The changing magnetic flux

induces a voltage in the secondary coil.

An ideal transformer

An ideal transformer is shown in the adjacent figure. Current passing through the primary coil

creates a magnetic field. The primary and secondary coils are wrapped around a core of very

Page 76: Basic Laws and Rules in Electrical Power

high magnetic permeability, such as iron, so that most of the magnetic flux passes through both

the primary and secondary coils.

Induction law

The voltage induced across the secondary coil may be calculated from Faraday's law of

induction, which states that:

where Vs is the instantaneous voltage, Ns is the number of turns in the secondary coil and Φ is the

magnetic flux through one turn of the coil. If the turns of the coil are oriented perpendicular to

the magnetic field lines, the flux is the product of the magnetic flux density B and the area A

through which it cuts. The area is constant, being equal to the cross-sectional area of the

transformer core, whereas the magnetic field varies with time according to the excitation of the

primary. Since the same magnetic flux passes through both the primary and secondary coils in an

ideal transformer,[29]

the instantaneous voltage across the primary winding equals

Taking the ratio of the two equations for Vs and Vp gives the basic equation[30]

for stepping up or

stepping down the voltage

Np/Ns is known as the turns ratio, and is the primary functional characteristic of any transformer.

In the case of step-up transformers, this may sometimes be stated as the reciprocal, Ns/Np. Turns

ratio is commonly expressed as an irreducible fraction or ratio: for example, a transformer with

primary and secondary windings of, respectively, 100 and 150 turns is said to have a turns ratio

of 2:3 rather than 0.667 or 100:150.

Ideal power equation

Page 77: Basic Laws and Rules in Electrical Power

The ideal transformer as a circuit element

If the secondary coil is attached to a load that allows current to flow, electrical power is

transmitted from the primary circuit to the secondary circuit. Ideally, the transformer is perfectly

efficient; all the incoming energy is transformed from the primary circuit to the magnetic field

and into the secondary circuit. If this condition is met, the incoming electric power must equal

the outgoing power:

giving the ideal transformer equation

Transformers normally have high efficiency, so this formula is a reasonable approximation.

If the voltage is increased, then the current is decreased by the same factor. The impedance in

one circuit is transformed by the square of the turns ratio.[29]

For example, if an impedance Zs is

attached across the terminals of the secondary coil, it appears to the primary circuit to have an

impedance of (Np/Ns)2Zs. This relationship is reciprocal, so that the impedance Zp of the primary

circuit appears to the secondary to be (Ns/Np)2Zp.

Detailed operation

The simplified description above neglects several practical factors, in particular the primary

current required to establish a magnetic field in the core, and the contribution to the field due to

current in the secondary circuit.

Models of an ideal transformer typically assume a core of negligible reluctance with two

windings of zero resistance.[31]

When a voltage is applied to the primary winding, a small current

flows, driving flux around the magnetic circuit of the core.[31]

The current required to create the

flux is termed the magnetizing current; since the ideal core has been assumed to have near-zero

Page 78: Basic Laws and Rules in Electrical Power

reluctance, the magnetizing current is negligible, although still required to create the magnetic

field.

The changing magnetic field induces an electromotive force (EMF) across each winding.[32]

Since the ideal windings have no impedance, they have no associated voltage drop, and so the

voltages VP and VS measured at the terminals of the transformer, are equal to the corresponding

EMFs. The primary EMF, acting as it does in opposition to the primary voltage, is sometimes

termed the "back EMF".[33]

This is due to Lenz's law which states that the induction of EMF

would always be such that it will oppose development of any such change in magnetic field.

Practical considerations

Leakage flux

Leakage flux of a transformer

The ideal transformer model assumes that all flux generated by the primary winding links all the

turns of every winding, including itself. In practice, some flux traverses paths that take it outside

the windings.[34]

Such flux is termed leakage flux, and results in leakage inductance in series with

the mutually coupled transformer windings.[33]

Leakage results in energy being alternately stored

in and discharged from the magnetic fields with each cycle of the power supply. It is not directly

a power loss (see "Stray losses" below), but results in inferior voltage regulation, causing the

secondary voltage to fail to be directly proportional to the primary, particularly under heavy

load.[34]

Transformers are therefore normally designed to have very low leakage inductance.

However, in some applications, leakage can be a desirable property, and long magnetic paths, air

gaps, or magnetic bypass shunts may be deliberately introduced to a transformer's design to limit

the short-circuit current it will supply.[33]

Leaky transformers may be used to supply loads that

exhibit negative resistance, such as electric arcs, mercury vapor lamps, and neon signs; or for

safely handling loads that become periodically short-circuited such as electric arc welders.[35]

Air gaps are also used to keep a transformer from saturating, especially audio-frequency

transformers in circuits that have a direct current flowing through the windings.

Page 79: Basic Laws and Rules in Electrical Power

Leakage inductance is also helpful when transformers are operated in parallel. It can be shown

that if the "per-unit" inductance of two transformers is the same (a typical value is 5% for low

power transformers), they will automatically split power "correctly" (e.g. 500 kVA unit in

parallel with 1,000 kVA unit, the larger one will carry twice the current).

Effect of frequency

Transformer universal EMF equation

If the flux in the core is purely sinusoidal, the relationship for either winding between its rms

voltage Erms of the winding , and the supply frequency f, number of turns N, core cross-sectional

area a and peak magnetic flux density B is given by the universal EMF equation:[31]

If the flux does not contain even harmonics the following equation can be used for half-cycle

average voltage Eavg of any waveshape:

The time-derivative term in Faraday's Law shows that the flux in the core is the integral with

respect to time of the applied voltage.[36]

Hypothetically an ideal transformer would work with

direct-current excitation, with the core flux increasing linearly with time.[37]

In practice, the flux

would rise to the point where magnetic saturation of the core occurs, causing a huge increase in

the magnetizing current and overheating the transformer. All practical transformers must

therefore operate with alternating (or pulsed) current.[37]

The EMF of a transformer at a given flux density increases with frequency.[31]

By operating at

higher frequencies, transformers can be physically more compact because a given core is able to

transfer more power without reaching saturation and fewer turns are needed to achieve the same

impedance. However, properties such as core loss and conductor skin effect also increase with

frequency. Aircraft and military equipment employ 400 Hz power supplies which reduce core

and winding weight.[38]

Conversely, frequencies used for some railway electrification systems

were much lower (e.g. 16.7 Hz and 25 Hz) than normal utility frequencies (50 – 60 Hz) for

historical reasons concerned mainly with the limitations of early electric traction motors. As

such, the transformers used to step down the high over-head line voltages (e.g. 15 kV) are much

heavier for the same power rating than those designed only for the higher frequencies.

Operation of a transformer at its designed voltage but at a higher frequency than intended will

lead to reduced magnetizing current; at lower frequency, the magnetizing current will increase.

Operation of a transformer at other than its design frequency may require assessment of voltages,

losses, and cooling to establish if safe operation is practical. For example, transformers may need

to be equipped with "volts per hertz" over-excitation relays to protect the transformer from

overvoltage at higher than rated frequency.

Page 80: Basic Laws and Rules in Electrical Power

One example of state-of-the-art design is those transformers used for electric multiple unit high

speed trains, particularly those required to operate across the borders of countries using different

standards of electrification. The position of such transformers is restricted to being hung below

the passenger compartment. They have to function at different frequencies (down to 16.7 Hz)

and voltages (up to 25 kV) whilst handling the enhanced power requirements needed for

operating the trains at high speed.

Knowledge of natural frequencies of transformer windings is of importance for the determination

of the transient response of the windings to impulse and switching surge voltages.

Energy losses

An ideal transformer would have no energy losses, and would be 100% efficient. In practical

transformers energy is dissipated in the windings, core, and surrounding structures. Larger

transformers are generally more efficient, and those rated for electricity distribution usually

perform better than 98%.[39]

Experimental transformers using superconducting windings achieve efficiencies of 99.85%.[40]

The increase in efficiency from about 98 to 99.85% can save considerable energy, and hence

money, in a large heavily-loaded transformer; the trade-off is in the additional initial and running

cost of the superconducting design.

Losses in transformers (excluding associated circuitry) vary with load current, and may be

expressed as "no-load" or "full-load" loss. Winding resistance dominates load losses, whereas

hysteresis and eddy currents losses contribute to over 99% of the no-load loss. The no-load loss

can be significant, so that even an idle transformer constitutes a drain on the electrical supply and

a running cost; designing transformers for lower loss requires a larger core, good-quality silicon

steel, or even amorphous steel, for the core, and thicker wire, increasing initial cost, so that there

is a trade-off between initial cost and running cost. (Also see energy efficient transformer).[41]

Transformer losses are divided into losses in the windings, termed copper loss, and those in the

magnetic circuit, termed iron loss. Losses in the transformer arise from:

Winding resistance Current flowing through the windings causes resistive heating of the conductors. At

higher frequencies, skin effect and proximity effect create additional winding resistance

and losses.

Hysteresis losses Each time the magnetic field is reversed, a small amount of energy is lost due to

hysteresis within the core. For a given core material, the loss is proportional to the

frequency, and is a function of the peak flux density to which it is subjected.[41]

Eddy currents Ferromagnetic materials are also good conductors, and a core made from such a material

also constitutes a single short-circuited turn throughout its entire length. Eddy currents

therefore circulate within the core in a plane normal to the flux, and are responsible for

resistive heating of the core material. The eddy current loss is a complex function of the

Page 81: Basic Laws and Rules in Electrical Power

square of supply frequency and inverse square of the material thickness.[41]

Eddy current

losses can be reduced by making the core of a stack of plates electrically insulated from

each other, rather than a solid block; all transformers operating at low frequencies use

laminated or similar cores.

Magnetostriction Magnetic flux in a ferromagnetic material, such as the core, causes it to physically

expand and contract slightly with each cycle of the magnetic field, an effect known as

magnetostriction. This produces the buzzing sound commonly associated with

transformers,[30]

and can cause losses due to frictional heating.

Mechanical losses In addition to magnetostriction, the alternating magnetic field causes fluctuating forces

between the primary and secondary windings. These incite vibrations within nearby

metalwork, adding to the buzzing noise, and consuming a small amount of power.[42]

Stray losses Leakage inductance is by itself largely lossless, since energy supplied to its magnetic

fields is returned to the supply with the next half-cycle. However, any leakage flux that

intercepts nearby conductive materials such as the transformer's support structure will

give rise to eddy currents and be converted to heat.[43]

There are also radiative losses due

to the oscillating magnetic field, but these are usually small.

Dot convention

It is common in transformer schematic symbols for there to be a dot at the end of each coil

within a transformer, particularly for transformers with multiple primary and secondary

windings. The dots indicate the direction of each winding relative to the others. Voltages at the

dot end of each winding are in phase; current flowing into the dot end of a primary coil will

result in current flowing out of the dot end of a secondary coil.

Equivalent circuit

Refer to the diagram below

The physical limitations of the practical transformer may be brought together as an equivalent

circuit model (shown below) built around an ideal lossless transformer.[44]

Power loss in the

windings is current-dependent and is represented as in-series resistances Rp and Rs. Flux leakage

results in a fraction of the applied voltage dropped without contributing to the mutual coupling,

and thus can be modeled as reactances of each leakage inductance Xp and Xs in series with the

perfectly coupled region.

Iron losses are caused mostly by hysteresis and eddy current effects in the core, and are

proportional to the square of the core flux for operation at a given frequency.[45]

Since the core

flux is proportional to the applied voltage, the iron loss can be represented by a resistance RC in

parallel with the ideal transformer.

A core with finite permeability requires a magnetizing current Im to maintain the mutual flux in

the core. The magnetizing current is in phase with the flux; saturation effects cause the

Page 82: Basic Laws and Rules in Electrical Power

relationship between the two to be non-linear, but for simplicity this effect tends to be ignored in

most circuit equivalents.[45]

With a sinusoidal supply, the core flux lags the induced EMF by 90°

and this effect can be modeled as a magnetizing reactance (reactance of an effective inductance)

Xm in parallel with the core loss component. Rc and Xm are sometimes together termed the

magnetizing branch of the model. If the secondary winding is made open-circuit, the current I0

taken by the magnetizing branch represents the transformer's no-load current.[44]

The secondary impedance Rs and Xs is frequently moved (or "referred") to the primary side after

multiplying the components by the impedance scaling factor (Np/Ns)2.

Transformer equivalent circuit, with secondary impedances referred to the primary side

The resulting model is sometimes termed the "exact equivalent circuit", though it retains a

number of approximations, such as an assumption of linearity.[44]

Analysis may be simplified by

moving the magnetizing branch to the left of the primary impedance, an implicit assumption that

the magnetizing current is low, and then summing primary and referred secondary impedances,

resulting in so-called equivalent impedance.

The parameters of equivalent circuit of a transformer can be calculated from the results of two

transformer tests: open-circuit test and short-circuit test.

Types

A wide variety of transformer designs are used for different applications, though they share

several common features. Important common transformer types include:

Autotransformer

Page 83: Basic Laws and Rules in Electrical Power

An autotransformer with a sliding brush contact

An autotransformer has a single winding with two end terminals, and one or more terminals at

intermediate tap points. The primary voltage is applied across two of the terminals, and the

secondary voltage taken from two terminals, almost always having one terminal in common with

the primary voltage. The primary and secondary circuits therefore have a number of windings

turns in common.[46]

Since the volts-per-turn is the same in both windings, each develops a

voltage in proportion to its number of turns. In an autotransformer part of the current flows

directly from the input to the output, and only part is transferred inductively, allowing a smaller,

lighter, cheaper core to be used as well as requiring only a single winding[47]

. However, a

transformer with separate windings isolates the primary from the secondary, which is safer when

using mains voltages.

An adjustable autotransformer is made by exposing part of the winding coils and making the

secondary connection through a sliding brush, giving a variable turns ratio.[48]

Such a device is

often referred to as a variac.

Autotransformers are often used to step up or down between voltages in the 110-117-120 volt

range and voltages in the 220-230-240 volt range, e.g., to output either 110 or 120V (with taps)

from 230V input, allowing equipment from a 100 or 120V region to be used in a 230V region.

Polyphase transformers

Page 84: Basic Laws and Rules in Electrical Power

Three-phase step-down transformer mounted between two utility poles

For three-phase supplies, a bank of three individual single-phase transformers can be used, or all

three phases can be incorporated as a single three-phase transformer. In this case, the magnetic

circuits are connected together, the core thus containing a three-phase flow of flux.[49]

A number

of winding configurations are possible, giving rise to different attributes and phase shifts.[50]

One

particular polyphase configuration is the zigzag transformer, used for grounding and in the

suppression of harmonic currents.[51]

Leakage transformers

Leakage transformer

A leakage transformer, also called a stray-field transformer, has a significantly higher leakage

inductance than other transformers, sometimes increased by a magnetic bypass or shunt in its

core between primary and secondary, which is sometimes adjustable with a set screw. This

provides a transformer with an inherent current limitation due to the loose coupling between its

Page 85: Basic Laws and Rules in Electrical Power

primary and the secondary windings. The output and input currents are low enough to prevent

thermal overload under all load conditions—even if the secondary is shorted.

Leakage transformers are used for arc welding and high voltage discharge lamps (neon lights and

cold cathode fluorescent lamps, which are series-connected up to 7.5 kV AC). It acts then both as

a voltage transformer and as a magnetic ballast.

Other applications are short-circuit-proof extra-low voltage transformers for toys or doorbell

installations.

Resonant transformers

A resonant transformer is a kind of leakage transformer. It uses the leakage inductance of its

secondary windings in combination with external capacitors, to create one or more resonant

circuits. Resonant transformers such as the Tesla coil can generate very high voltages, and are

able to provide much higher current than electrostatic high-voltage generation machines such as

the Van de Graaff generator.[52]

One of the applications of the resonant transformer is for the

CCFL inverter. Another application of the resonant transformer is to couple between stages of a

superheterodyne receiver, where the selectivity of the receiver is provided by tuned transformers

in the intermediate-frequency amplifiers.[53]

Audio transformers

Audio transformers are those specifically designed for use in audio circuits. They can be used to

block radio frequency interference or the DC component of an audio signal, to split or combine

audio signals, or to provide impedance matching between high and low impedance circuits, such

as between a high impedance tube (valve) amplifier output and a low impedance loudspeaker, or

between a high impedance instrument output and the low impedance input of a mixing console.

Such transformers were originally designed to connect different telephone systems to one

another while keeping their respective power supplies isolated, and are still commonly used to

interconnect professional audio systems or system components.

Being magnetic devices, audio transformers are susceptible to external magnetic fields such as

those generated by AC current-carrying conductors. "Hum" is a term commonly used to describe

unwanted signals originating from the "mains" power supply (typically 50 or 60 Hz). Audio

transformers used for low-level signals, such as those from microphones, often include shielding

to protect against extraneous magnetically coupled signals.

Instrument transformers

Instrument transformers are used for measuring voltage and current in electrical power systems,

and for power system protection and control. Where a voltage or current is too large to be

conveniently used by an instrument, it can be scaled down to a standardized, low value.

Instrument transformers isolate measurement, protection and control circuitry from the high

currents or voltages present on the circuits being measured or controlled.

Page 86: Basic Laws and Rules in Electrical Power

Current transformers, designed for placing around conductors

A current transformer is a transformer designed to provide a current in its secondary coil

proportional to the current flowing in its primary coil.[54]

Voltage transformers (VTs), also referred to as "potential transformers" (PTs), are designed to

have an accurately known transformation ratio in both magnitude and phase, over a range of

measuring circuit impedances. A voltage transformer is intended to present a negligible load to

the supply being measured. The low secondary voltage allows protective relay equipment and

measuring instruments to be operated at a lower voltages.[55]

Both current and voltage instrument transformers are designed to have predictable characteristics

on overloads. Proper operation of over-current protective relays requires that current

transformers provide a predictable transformation ratio even during a short-circuit.

Classification

Transformers can be classified in many different ways; an incomplete list is:

By power capacity: from a fraction of a volt-ampere (VA) to over a thousand MVA;

By frequency range: power-, audio-, or radio frequency;

By voltage class: from a few volts to hundreds of kilovolts;

By cooling type: air-cooled, oil-filled, fan-cooled, or water-cooled;

By application: such as power supply, impedance matching, output voltage and current

stabilizer, or circuit isolation;

By purpose: distribution, rectifier, arc furnace, amplifier output, etc.;

By winding turns ratio: step-up, step-down, isolating with equal or near-equal ratio,

variable, multiple windings.

Construction

Cores

Page 87: Basic Laws and Rules in Electrical Power

Laminated core transformer showing edge of laminations at top of photo

Laminated steel cores

Transformers for use at power or audio frequencies typically have cores made of high

permeability silicon steel.[56]

The steel has a permeability many times that of free space, and the

core thus serves to greatly reduce the magnetizing current, and confine the flux to a path which

closely couples the windings.[57]

Early transformer developers soon realized that cores

constructed from solid iron resulted in prohibitive eddy-current losses, and their designs

mitigated this effect with cores consisting of bundles of insulated iron wires.[8]

Later designs

constructed the core by stacking layers of thin steel laminations, a principle that has remained in

use. Each lamination is insulated from its neighbors by a thin non-conducting layer of

insulation.[49]

The universal transformer equation indicates a minimum cross-sectional area for

the core to avoid saturation.

The effect of laminations is to confine eddy currents to highly elliptical paths that enclose little

flux, and so reduce their magnitude. Thinner laminations reduce losses,[56]

but are more laborious

and expensive to construct.[58]

Thin laminations are generally used on high frequency

transformers, with some types of very thin steel laminations able to operate up to 10 kHz.

One common design of laminated core is made from interleaved stacks of E-shaped steel sheets

capped with I-shaped pieces, leading to its name of "E-I transformer".[58]

Such a design tends to

exhibit more losses, but is very economical to manufacture. The cut-core or C-core type is made

by winding a steel strip around a rectangular form and then bonding the layers together. It is then

cut in two, forming two C shapes, and the core assembled by binding the two C halves together

with a steel strap.[58]

They have the advantage that the flux is always oriented parallel to the

metal grains, reducing reluctance.

A steel core's remanence means that it retains a static magnetic field when power is removed.

When power is then reapplied, the residual field will cause a high inrush current until the effect

of the remaining magnetism is reduced, usually after a few cycles of the applied alternating

current.[59]

Overcurrent protection devices such as fuses must be selected to allow this harmless

inrush to pass. On transformers connected to long, overhead power transmission lines, induced

currents due to geomagnetic disturbances during solar storms can cause saturation of the core

and operation of transformer protection devices.[60]

Page 88: Basic Laws and Rules in Electrical Power

Distribution transformers can achieve low no-load losses by using cores made with low-loss

high-permeability silicon steel or amorphous (non-crystalline) metal alloy. The higher initial cost

of the core material is offset over the life of the transformer by its lower losses at light load.[61]

Solid cores

Powdered iron cores are used in circuits (such as switch-mode power supplies) that operate

above main frequencies and up to a few tens of kilohertz. These materials combine high

magnetic permeability with high bulk electrical resistivity. For frequencies extending beyond the

VHF band, cores made from non-conductive magnetic ceramic materials called ferrites are

common.[58]

Some radio-frequency transformers also have movable cores (sometimes called

'slugs') which allow adjustment of the coupling coefficient (and bandwidth) of tuned radio-

frequency circuits.

Toroidal cores

Small toroidal core transformer

Toroidal transformers are built around a ring-shaped core, which, depending on operating

frequency, is made from a long strip of silicon steel or permalloy wound into a coil, powdered

iron, or ferrite.[62]

A strip construction ensures that the grain boundaries are optimally aligned,

improving the transformer's efficiency by reducing the core's reluctance. The closed ring shape

eliminates air gaps inherent in the construction of an E-I core.[35]

The cross-section of the ring is

usually square or rectangular, but more expensive cores with circular cross-sections are also

available. The primary and secondary coils are often wound concentrically to cover the entire

surface of the core. This minimizes the length of wire needed, and also provides screening to

minimize the core's magnetic field from generating electromagnetic interference.

Toroidal transformers are more efficient than the cheaper laminated E-I types for a similar power

level. Other advantages compared to E-I types, include smaller size (about half), lower weight

(about half), less mechanical hum (making them superior in audio amplifiers), lower exterior

magnetic field (about one tenth), low off-load losses (making them more efficient in standby

circuits), single-bolt mounting, and greater choice of shapes. The main disadvantages are higher

cost and limited power capacity (see "Classification" above). Because of the lack of a residual

gap in the magnetic path, toroidal transformers also tend to exhibit higher inrush current,

compared to laminated E-I types.

Page 89: Basic Laws and Rules in Electrical Power

Ferrite toroidal cores are used at higher frequencies, typically between a few tens of kilohertz to

hundreds of megahertz, to reduce losses, physical size, and weight of switch-mode power

supplies. A drawback of toroidal transformer construction is the higher labor cost of winding.

This is because it is necessary to pass the entire length of a coil winding through the core

aperture each time a single turn is added to the coil. As a consequence, toroidal transformers are

uncommon above ratings of a few kVA. Small distribution transformers may achieve some of

the benefits of a toroidal core by splitting it and forcing it open, then inserting a bobbin

containing primary and secondary windings.

Air cores

A physical core is not an absolute requisite and a functioning transformer can be produced

simply by placing the windings near each other, an arrangement termed an "air-core"

transformer. The air which comprises the magnetic circuit is essentially lossless, and so an air-

core transformer eliminates loss due to hysteresis in the core material.[33]

The leakage inductance

is inevitably high, resulting in very poor regulation, and so such designs are unsuitable for use in

power distribution.[33]

They have however very high bandwidth, and are frequently employed in

radio-frequency applications,[63]

for which a satisfactory coupling coefficient is maintained by

carefully overlapping the primary and secondary windings. They're also used for resonant

transformers such as Tesla coils where they can achieve reasonably low loss in spite of the high

leakage inductance.

Windings

Windings are usually arranged concentrically to minimize flux leakage.

Page 90: Basic Laws and Rules in Electrical Power

Cut view through transformer windings. White: insulator. Green spiral: Grain oriented silicon

steel. Black: Primary winding made of oxygen-free copper. Red: Secondary winding. Top left:

Toroidal transformer. Right: C-core, but E-core would be similar. The black windings are made

of film. Top: Equally low capacitance between all ends of both windings. Since most cores are at

least moderately conductive they also need insulation. Bottom: Lowest capacitance for one end

of the secondary winding needed for low-power high-voltage transformers. Bottom left:

Reduction of leakage inductance would lead to increase of capacitance.

The conducting material used for the windings depends upon the application, but in all cases the

individual turns must be electrically insulated from each other to ensure that the current travels

throughout every turn.[36]

For small power and signal transformers, in which currents are low and

the potential difference between adjacent turns is small, the coils are often wound from

enamelled magnet wire, such as Formvar wire. Larger power transformers operating at high

voltages may be wound with copper rectangular strip conductors insulated by oil-impregnated

paper and blocks of pressboard.[64]

High-frequency transformers operating in the tens to hundreds of kilohertz often have windings

made of braided Litz wire to minimize the skin-effect and proximity effect losses.[36]

Large

power transformers use multiple-stranded conductors as well, since even at low power

frequencies non-uniform distribution of current would otherwise exist in high-current

windings.[64]

Each strand is individually insulated, and the strands are arranged so that at certain

points in the winding, or throughout the whole winding, each portion occupies different relative

positions in the complete conductor. The transposition equalizes the current flowing in each

strand of the conductor, and reduces eddy current losses in the winding itself. The stranded

conductor is also more flexible than a solid conductor of similar size, aiding manufacture.[64]

For signal transformers, the windings may be arranged in a way to minimize leakage inductance

and stray capacitance to improve high-frequency response. This can be done by splitting up each

coil into sections, and those sections placed in layers between the sections of the other winding.

This is known as a stacked type or interleaved winding.

Both the primary and secondary windings on power transformers may have external connections,

called taps, to intermediate points on the winding to allow selection of the voltage ratio. In power

distribution transformers the taps may be connected to an automatic on-load tap changer for

Page 91: Basic Laws and Rules in Electrical Power

voltage regulation of distribution circuits. Audio-frequency transformers, used for the

distribution of audio to public address loudspeakers, have taps to allow adjustment of impedance

to each speaker. A center-tapped transformer is often used in the output stage of an audio power

amplifier in a push-pull circuit. Modulation transformers in AM transmitters are very similar.

Certain transformers have the windings protected by epoxy resin. By impregnating the

transformer with epoxy under a vacuum, one can replace air spaces within the windings with

epoxy, thus sealing the windings and helping to prevent the possible formation of corona and

absorption of dirt or water. This produces transformers more suited to damp or dirty

environments, but at increased manufacturing cost.[65]

Coolant

Cut-away view of three-phase oil-cooled transformer. The oil reservoir is visible at the top.

Radiative fins aid the dissipation of heat.

High temperatures will damage the winding insulation.[66]

Small transformers do not generate

significant heat and are cooled by air circulation and radiation of heat. Power transformers rated

up to several hundred kVA can be adequately cooled by natural convective air-cooling,

sometimes assisted by fans.[67]

In larger transformers, part of the design problem is removal of

heat. Some power transformers are immersed in transformer oil that both cools and insulates the

windings.[68]

The oil is a highly refined mineral oil that remains stable at transformer operating

temperature. Indoor liquid-filled transformers are required by building regulations in many

jurisdictions to use a non-flammable liquid, or to be located in fire-resistant rooms.[69]

Air-cooled

dry transformers are preferred for indoor applications even at capacity ratings where oil-cooled

construction would be more economical, because their cost is offset by the reduced building

construction cost.

The oil-filled tank often has radiators through which the oil circulates by natural convection;

some large transformers employ forced circulation of the oil by electric pumps, aided by external

Page 92: Basic Laws and Rules in Electrical Power

fans or water-cooled heat exchangers.[68]

Oil-filled transformers undergo prolonged drying

processes to ensure that the transformer is completely free of water vapor before the cooling oil

is introduced. This helps prevent electrical breakdown under load. Oil-filled transformers may be

equipped with Buchholz relays, which detect gas evolved during internal arcing and rapidly de-

energize the transformer to avert catastrophic failure.[59]

Oil-filled transformers may fail, rupture,

and burn, causing power outages and losses. Installations of oil-filled transformers usually

includes fire protection measures such as walls, oil containment, and fire-suppression sprinkler

systems.

Polychlorinated biphenyls have properties that once favored their use as a coolant, though

concerns over their environmental persistence led to a widespread ban on their use.[70]

Today,

non-toxic, stable silicone-based oils, or fluorinated hydrocarbons may be used where the expense

of a fire-resistant liquid offsets additional building cost for a transformer vault.[66][69]

Before

1977, even transformers that were nominally filled only with mineral oils may also have been

contaminated with polychlorinated biphenyls at 10-20 ppm. Since mineral oil and PCB fluid

mix, maintenance equipment used for both PCB and oil-filled transformers could carry over

small amounts of PCB, contaminating oil-filled transformers.[71]

Some "dry" transformers (containing no liquid) are enclosed in sealed, pressurized tanks and

cooled by nitrogen or sulfur hexafluoride gas.[66]

Experimental power transformers in the 2 MVA range have been built with superconducting

windings which eliminates the copper losses, but not the core steel loss. These are cooled by

liquid nitrogen or helium.[72]

Insulation drying

Construction of oil-filled transformers requires that the insulation covering the windings be

thoroughly dried before the oil is introduced. There are several different methods of drying.

Common for all is that they are carried out in vacuum environment. The vacuum makes it

difficult to transfer energy (heat) to the insulation. For this there are several different methods.

The traditional drying is done by circulating hot air over the active part and cycle this with

periods of vacuum (Hot Air Vacuum drying, HAV). More common for larger transformers is to

use evaporated solvent which condenses on the colder active part. The benefit is that the entire

process can be carried out at lower pressure and without influence of added oxygen. This process

is commonly called Vapour Phase Drying (VPD).

For distribution transformers which are smaller and have a smaller insulation weight, resistance

heating can be used. This is a method where current is injected in the windings and the resistance

in the windings is heating up the insulation. The benefit is that the heating can be controlled very

well and it is energy efficient. The method is called Low Frequency Heating (LFH) since the

current is injected at a much lower frequency than the nominal of the grid, which is normally 50

or 60 Hz. A lower frequency reduces the affect of the inductance in the transformer and the

voltage can be reduced.

Terminals

Page 93: Basic Laws and Rules in Electrical Power

Very small transformers will have wire leads connected directly to the ends of the coils, and

brought out to the base of the unit for circuit connections. Larger transformers may have heavy

bolted terminals, bus bars or high-voltage insulated bushings made of polymers or porcelain. A

large bushing can be a complex structure since it must provide careful control of the electric field

gradient without letting the transformer leak oil.[73]

Applications

Image of an electrical substation in Melbourne, Australia showing 3 of 5 220kV/66kV

transformers, each with a capacity of 185MVA

A major application of transformers is to increase voltage before transmitting electrical energy

over long distances through wires. Wires have resistance and so dissipate electrical energy at a

rate proportional to the square of the current through the wire. By transforming electrical power

to a high-voltage (and therefore low-current) form for transmission and back again afterward,

transformers enable economic transmission of power over long distances. Consequently,

transformers have shaped the electricity supply industry, permitting generation to be located

remotely from points of demand.[74]

All but a tiny fraction of the world's electrical power has

passed through a series of transformers by the time it reaches the consumer.[43]

Transformers are also used extensively in electronic products to step down the supply voltage to

a level suitable for the low voltage circuits they contain. The transformer also electrically isolates

the end user from contact with the supply voltage.

Signal and audio transformers are used to couple stages of amplifiers and to match devices such

as microphones and record players to the input of amplifiers. Audio transformers allowed

telephone circuits to carry on a two-way conversation over a single pair of wires. A balun

transformer converts a signal that is referenced to ground to a signal that has balanced voltages to

ground, such as between external cables and internal circuits.

The principle of open-circuit (unloaded) transformer is widely used for characterisation of soft

magnetic materials, for example in the internationally standardised Epstein frame method[75]

.

Page 94: Basic Laws and Rules in Electrical Power

Notes

1. ^ Poyser, Arthur William (1892), Magnetism and electricity: A manual for students in advanced

classes. London and New York; Longmans, Green, & Co., p. 285, fig. 248. Retrieved 2009-08-

06.

2. ^ "Joseph Henry". Distinguished Members Gallery, National Academy of Sciences.

http://www.nas.edu/history/members/henry.html. Retrieved November 30, 2006.

3. ^ Chow, Tai L. (2006), Introduction to electromagnetic theory: a modern perspective. Sudbury,

Massachusetts (USA), Jones and Bartlett Publishers, p. 171. ISBN 0763738271.

4. ^ Faraday, Michael (1834). "Experimental Researches on Electricity, 7th Series". Philosophical

Transactions of the Royal Society of London 124: 77–122. doi:10.1098/rstl.1834.0008.

5. ^ a b Stanley Transformer. Los Alamos National Laboratory; University of Florida.

http://www.magnet.fsu.edu/education/tutorials/museum/stanleytransformer.html. Retrieved

January 9, 2009.

6. ^ W. De Fonveille (1880-1-22). "Gas and Electricity in Paris". Nature 21 (534): 283.

http://books.google.com/?id=ksa-S7C8dT8C&pg=RA2-PA283. Retrieved January 9, 2009.

7. ^ Hughes, Thomas P. (1993). Networks of Power: Electrification in Western Society, 1880-1930.

Baltimore: The Johns Hopkins University Press. p. 96. ISBN 0-8018-2873-2.

http://books.google.com/?id=g07Q9M4agp4C&pg=PA96&dq=Networks+of+Power:+Electrificat

ion+in+Western+Society,+1880-1930+ganz#v=onepage&q=. Retrieved September 9, 2009.

8. ^ a b Allan, D.J.. "Power transformers – the second century". Power Engineering Journal.

9. ^ Uppenborn, F. J., History of the Transformer, E. & F. N. Spon, London, 1889. p. 41.

10. ^ Friedrich Uppenborn: History of the Transformer, page 35.

11. ^ Uppenborn, F. J., History of the Transformer, E. & F. N. Spon, London, 1889.

12. ^ a b Coltman, J. W. (January 1988). "The Transformer". Scientific American: pp. 86–95. OSTI

6851152.

13. ^ http://dspace.thapar.edu:8080/dspace/bitstream/123456789/91/1/P92087.pdf

14. ^ www.aaidu.org/phdsubmit/phdpdf/06PHEE201.pdf

15. ^ [1]|SÁNDOR JESZENSZKY: Electrostatics and Electrodynamics at Pest University in the mid-

19th Century, ( page: 178) University of Pavia

16. ^ a b Bláthy, Ottó Titusz (1860 - 1939), Hungarian Patent Office, January 29, 2004.

17. ^ Zipernowsky, K., M. Déri and O. T. Bláthy, Induction Coil, Patent No. 352,105, U.S. Patent

Office, November 2, 1886, retrieved July 8, 2009.

18. ^ a b Smil, Vaclav, Creating the Twentieth Century: Technical Innovations of 1867—1914 and

Their Lasting Impact, Oxford University Press, 2005, p. 71.

19. ^ Nagy, Árpád Zoltán, "Lecture to Mark the 100th Anniversary of the Discovery of the Electron

in 1897" (preliminary text), Budapest October 11, 1996, retrieved July 9, 2009.

20. ^ Oxford English Dictionary (2nd ed.). Oxford University Press. 1989.

21. ^ Hospitalier, Édouard, 1882, The Modern Applications of Electricity, Translated and Enlarged

by Julius Maier. New York, D. Appleton & Co., p. 103.

22. ^ "Ottó Bláthy, Miksa Déri, Károly Zipernowsky", IEC Techline. Retrieved 2010-04-16.

23. ^ a b Stanley, William, Jr., Induction Coil. Patent No. 0349611, United States Patent Office,

September 21, 1886. Retrieved July 13, 2009.

24. ^ Skrabec, Quentin R. (2007). George Westinghouse: Gentle Genius. Algora Publishing. p. 102.

ISBN 978-0875865089. http://books.google.com/?id=C3GYdiFM41oC&pg=PA102.

25. ^ Westinghouse, G. Jr., Electrical Converter, Patent No. 366362, United States Patent Office,

1887.

26. ^ Neidhöfer, Gerhard: Michael von Dolivo-Dobrowolsky und der Drehstrom. Geschichte der

Elektrotechnik, Volume 9, VDE VERLAG, Berlin Offenbach, ISBN 978-3-8007-3115-2.

(German)

Page 95: Basic Laws and Rules in Electrical Power

27. ^ Uth, Robert (December 12, 2000). "Tesla coil". Tesla: Master of Lightning. PBS.org.

http://www.pbs.org/tesla/ins/lab_tescoil.html. Retrieved May 20, 2008.

28. ^ Tesla, Nikola (1891). "System of elecrical lighting". United States Patent and Trademark

Office. http://patft.uspto.gov/netacgi/nph-

Parser?Sect1=PTO1&Sect2=HITOFF&d=PALL&p=1&u=%2Fnetahtml%2FPTO%2Fsrchnum.ht

m&r=1&f=G&l=50&s1=0454622.PN.&OS=PN/0454622&RS=PN/0454622.

29. ^ a b Flanagan, William M. (January 1, 1993). Handbook of Transformer Design and

Applications. McGraw-Hill Professional. Chap. 1, p. 1–2. ISBN 0070212910.

30. ^ a b Winders. Power Transformer Principles and Applications. pp. 20–21.

31. ^ a b c d Say, M. G. (February, 1984). Alternating Current Machines, Fifth Edition. Halsted Press.

ISBN 0470274514.

32. ^ Heathcote, Martin (November 3, 1998). J & P Transformer Book, Twelfth edition. Newnes.

pp. 2–3. ISBN 0750611588.

33. ^ a b c d e Calvert, James (2001). "Inside Transformers". University of Denver.

http://www.du.edu/~jcalvert/tech/transfor.htm. Retrieved May 19, 2007.

34. ^ a b McLaren, P. G. (January 1, 1984). Elementary Electric Power and Machines. pp. 68–74.

ISBN 0132576015.

35. ^ a b Say, M. G. (February, 1984). Alternating Current Machines, Fifth Edition. Halsted Press.

p. 485. ISBN 0470274514.

36. ^ a b c Dixon, Lloyd. "Magnetics Design Handbook". Texas Instruments.

http://focus.ti.com/lit/ml/slup126/slup126.pdf.

37. ^ a b Billings, Keith (1999). Switchmode Power Supply Handbook. McGraw-Hill.

ISBN 0070067198.

38. ^ "400 Hz Electrical Systems". Aerospaceweb.org.

http://www.aerospaceweb.org/question/electronics/q0219.shtml. Retrieved May 21, 2007.

39. ^ Kubo, T.; Sachs, H.; Nadel, S. (2001) (PDF). Opportunities for new appliance and equipment

efficiency standards. American Council for an Energy-Efficient Economy. p. 39.

http://www.aceee.org/pubs/a016full.pdf. Retrieved June 21, 2009.

40. ^ Riemersma, H., et al.; Eckels, P.; Barton, M.; Murphy, J.; Litz, D.; Roach, J. (1981).

"Application of Superconducting Technology to Power Transformers". IEEE Transactions on

Power Apparatus and Systems PAS-100 (7): 3398. doi:10.1109/TPAS.1981.316682.

http://md1.csa.com/partners/viewrecord.php?requester=gs&collection=TRD&recid=0043264EA

&q=superconducting+transformer&uid=790516502&setcookie=yes.

41. ^ a b c Heathcote, Martin (November 3, 1998). J & P Transformer Book, Twelfth edition. Newnes.

pp. 41–42. ISBN 0750611588.

42. ^ Pansini, Anthony J. (1999). Electrical Transformers and Power Equipment. Fairmont Press.

p. 23. ISBN 0881733113.

43. ^ a b Nailen, Richard (May 2005). "Why we must be concerned with transformers". Electrical

Apparatus. http://findarticles.com/p/articles/mi_qa3726/is_200505/ai_n13636839/pg_1.

44. ^ a b c Daniels, A. R.. Introduction to Electrical Machines. pp. 47–49.

45. ^ a b Say, M. G. (February, 1984). Alternating Current Machines, Fifth Edition. Halsted Press.

pp. 142–143. ISBN 0470274514.

46. ^ Pansini. Electrical Transformers and Power Equipment. pp. 89–91.

47. ^ Commercial site explaining why autotransformers are smaller

48. ^ Bakshi, M. V. and Bakshi, U. A.. Electrical Machines - I. p. 330. ISBN 8184310099.

49. ^ a b Kulkarni, S. V. and Khaparde, S. A. (May 24, 2004). Transformer Engineering: design and

practice. CRC. pp. 36–37. ISBN 0824756533.

50. ^ Say, M. G. (February, 1984). Alternating Current Machines, Fifth Edition. Halsted Press.

p. 166. ISBN 0470274514.

51. ^ Hindmarsh. Electrical Machines and their Applications. p. 173.

Page 96: Basic Laws and Rules in Electrical Power

52. ^ Abdel-Salam, M. et al.. High-Voltage Engineering: Theory and Practice. pp. 523–524.

ISBN 0824741528.

53. ^ Carr, Joseph. Secrets of RF Circuit Design. pp. 193–195. ISBN 0071370676.

54. ^ Guile, A. and Paterson, W. (1978). Electrical Power Systems, Volume One. Oxford: Pergamon

Press. pp. 330–331. ISBN 008021729X.

55. ^ Institution of Electrical Engineers (1995). Power System Protection. London: Institution of

Electrical Engineers. pp. 38–39. ISBN 0852968345.

56. ^ a b Hindmarsh, John (1984). Electrical Machines and their Applications. Pergamon. pp. 29–31.

ISBN 0080305733.

57. ^ Gottlieb, Irving (1998). Practical Transformer Handbook. Newnes. p. 4. ISBN 075063992X.

58. ^ a b c d McLyman, Colonel Wm. T. (2004). Transformer and Inductor Design Handbook. CRC.

Chap. 3, pp. 9–14. ISBN 0824753933.

59. ^ a b Harlow, James H.. Electric Power Transformer Engineering. Taylor & Francis. Chap. 2, pp.

20–21.

60. ^ Boteler, D. H.; Pirjola, R. J.; Nevanlinna, H. (1998). "The effects of geomagnetic disturbances

on electrical systems at the Earth's surface". Advances in Space Research 22: 17–27.

doi:10.1016/S0273-1177(97)01096-X.

61. ^ Hasegawa, Ryusuke (June 2, 2000). "Present status of amorphous soft magnetic alloys".

Journal of Magnetism and Magnetic Materials 215-216: 240–245. doi:10.1016/S0304-

8853(00)00126-8.

62. ^ McLyman. Transformer and Inductor Design Handbook. Chap. 3 p1.

63. ^ Lee, Reuben. "Air-Core Transformers". Electronic Transformers and Circuits.

http://www.vias.org/eltransformers/lee_electronic_transformers_07b_22.html. Retrieved May 22,

2007.

64. ^ a b c Central Electricity Generating Board (1982). Modern Power Station Practice. Pergamon

Press.

65. ^ Heathcote, Martin (November 3, 1998). J & P Transformer Book. Newnes. pp. 720–723.

ISBN 0750611588.

66. ^ a b c Kulkarni, S. V. and Khaparde, S. A. (May 24, 2004). Transformer Engineering: design and

practice. CRC. pp. 2–3. ISBN 0824756533.

67. ^ Pansini, Anthony J. (1999). Electrical Transformers and Power Equipment. Fairmont Press.

p. 32. ISBN 0881733113.

68. ^ a b Willis, H. Lee (2004). Power Distribution Planning Reference Book. CRC Press. p. 403.

ISBN 0824748751.

69. ^ a b ENERGIE (1999) (PDF). The scope for energy saving in the EU through the use of energy-

efficient electricity distribution transformers. http://www.leonardo-

energy.org/drupal/files/Full%20project%20report%20-%20Thermie.pdf?download.

70. ^ "ASTDR ToxFAQs for Polychlorinated Biphenyls". 2001.

http://www.atsdr.cdc.gov/tfacts17.html. Retrieved June 10, 2007.

71. ^ McDonald, C. J. and Tourangeau, R. E. (1986). PCBs: Question and Answer Guide Concerning

Polychlorinated Biphenyls. Government of Canada: Environment Canada Department. p. 9.

ISBN 066214595X. http://www.ec.gc.ca/wmd-dgd/default.asp?lang=En&n=AD2C1530-

1&offset=3&toc=show#anchor6. Retrieved November 7, 2007.

72. ^ Pansini, Anthony J. (1999). Electrical Transformers and Power Equipment. Fairmont Press.

pp. 66–67. ISBN 0881733113.

73. ^ Ryan, Hugh M. (2001). High Voltage Engineering and Testing. Institution Electrical Engineers.

pp. 416–417. ISBN 0852967756.

74. ^ Heathcote. J & P Transformer Book. p. 1.

75. ^ IEC 60404-2

Page 97: Basic Laws and Rules in Electrical Power

References

Central Electricity Generating Board (1982). Modern Power Station Practice. Pergamon.

ISBN 0-08-016436-6.

Daniels, A.R. (1985). Introduction to Electrical Machines. Macmillan. ISBN 0-333-

19627-9.

Flanagan, William (1993). Handbook of Transformer Design and Applications. McGraw-

Hill. ISBN 0-0702-1291-0.

Gottlieb, Irving (1998). Practical Transformer Handbook. Elsevier. ISBN 0-7506-3992-

X.

Hammond, John Winthrop. Men and Volts, the Story of General Electric, published 1941

by J.B.Lippincott. Citations: design, early types - 106-107; design, William Stanley, first

built - 178; oil-immersed, began use of - 238.

Harlow, James (2004). Electric Power Transformer Engineering. CRC Press. ISBN 0-

8493-1704-5.

Heathcote, Martin (1998). J & P Transformer Book, Twelfth edition. Newnes. ISBN 0-

7506-1158-8.

Hindmarsh, John (1977). Electrical Machines and their Applications, 4th edition. Exeter:

Pergammon. ISBN 0-08-030573-3.

Kulkarni, S.V. & Khaparde, S.A. (2004). Transformer Engineering: design and practice.

CRC Press. ISBN 0-8247-5653-3.

McLaren, Peter (1984). Elementary Electric Power and Machines. Ellis Horwood.

ISBN 0-4702-0057-X.

McLyman, Colonel William (2004). Transformer and Inductor Design Handbook. CRC.

ISBN 0-8247-5393-3.

Pansini, Anthony (1999). Electrical Transformers and Power Equipment. CRC Press.

p. 23. ISBN 0-8817-3311-3.

Ryan, H.M. (2004). High Voltage Engineering and Testing. CRC Press. ISBN 0-8529-6775-6.

Say, M.G. (1983). Alternating Current Machines, Fifth Edition. London: Pitman. ISBN 0-273-

01969-4.

Winders, John (2002). Power Transformer Principles and Applications. CRC. ISBN 0-8247-

0766-4.

Gururaj, B.I. (June 1963). "Natural Frequencies of 3-Phase Transformer Windings". IEEE

Transactions on Power Apparatus and Systems 82 (66): 318–329.

doi:10.1109/TPAS.1963.291359.

http://ieeexplore.ieee.org/xpl/freeabs_all.jsp?isnumber=4072786&arnumber=4072800&count=2

5&index=12.

Three-phase transformer circuits

Page 98: Basic Laws and Rules in Electrical Power

Since three-phase is used so often for power distribution systems, it makes sense that we would

need three-phase transformers to be able to step voltages up or down. This is only partially true,

as regular single-phase transformers can be ganged together to transform power between two

three-phase systems in a variety of configurations, eliminating the requirement for a special

three-phase transformer. However, special three-phase transformers are built for those tasks, and

are able to perform with less material requirement, less size, and less weight than their modular

counterparts.

A three-phase transformer is made of three sets of primary and secondary windings, each set

wound around one leg of an iron core assembly. Essentially it looks like three single-phase

transformers sharing a joined core as in Figure below.

Three phase transformer core has three sets of windings.

Those sets of primary and secondary windings will be connected in either Δ or Y configurations

to form a complete unit. The various combinations of ways that these windings can be connected

together in will be the focus of this section.

Whether the winding sets share a common core assembly or each winding pair is a separate

transformer, the winding connection options are the same:

Primary - Secondary

Y - Y

Y - Δ

Δ - Y

Δ - Δ

The reasons for choosing a Y or Δ configuration for transformer winding connections are the

same as for any other three-phase application: Y connections provide the opportunity for

multiple voltages, while Δ connections enjoy a higher level of reliability (if one winding fails

open, the other two can still maintain full line voltages to the load).

Probably the most important aspect of connecting three sets of primary and secondary windings

together to form a three-phase transformer bank is paying attention to proper winding phasing

Page 99: Basic Laws and Rules in Electrical Power

(the dots used to denote ―polarity‖ of windings). Remember the proper phase relationships

between the phase windings of Δ and Y: (Figure below)

(Y) The center point of the ―Y‖ must tie either all the ―-‖ or all the ―+‖ winding points together.

(Δ) The winding polarities must stack together in a complementary manner ( + to -).

Getting this phasing correct when the windings aren't shown in regular Y or Δ configuration can

be tricky. Let me illustrate, starting with Figure below.

Inputs A1, A2, A3 may be wired either ―Δ‖ or ―Y‖, as may outputs B1, B2, B3.

Three individual transformers are to be connected together to transform power from one three-

phase system to another. First, I'll show the wiring connections for a Y-Y configuration: Figure

below

Page 100: Basic Laws and Rules in Electrical Power

Phase wiring for ―Y-Y‖ transformer.

Note in Figure above how all the winding ends marked with dots are connected to their

respective phases A, B, and C, while the non-dot ends are connected together to form the centers

of each ―Y‖. Having both primary and secondary winding sets connected in ―Y‖ formations

allows for the use of neutral conductors (N1 and N2) in each power system.

Now, we'll take a look at a Y-Δ configuration: (Figure below)

Page 101: Basic Laws and Rules in Electrical Power

Phase wiring for ―Y-Δ‖ transformer.

Note how the secondary windings (bottom set, Figure above) are connected in a chain, the ―dot‖

side of one winding connected to the ―non-dot‖ side of the next, forming the Δ loop. At every

connection point between pairs of windings, a connection is made to a line of the second power

system (A, B, and C).

Now, let's examine a Δ-Y system in Figure below.

Phase wiring for ―Δ-Y‖ transformer.

Such a configuration (Figure above) would allow for the provision of multiple voltages (line-to-

line or line-to-neutral) in the second power system, from a source power system having no

neutral.

And finally, we turn to the Δ-Δ configuration: (Figure below)

Page 102: Basic Laws and Rules in Electrical Power

Phase wiring for ―Δ-Δ‖ transformer.

When there is no need for a neutral conductor in the secondary power system, Δ-Δ connection

schemes (Figure above) are preferred because of the inherent reliability of the Δ configuration.

Considering that a Δ configuration can operate satisfactorily missing one winding, some power

system designers choose to create a three-phase transformer bank with only two transformers,

representing a Δ-Δ configuration with a missing winding in both the primary and secondary

sides: (Figure below)

Page 103: Basic Laws and Rules in Electrical Power

―V‖ or ―open-Δ‖ provides 2-υ power with only two transformers.

This configuration is called ―V‖ or ―Open-Δ.‖ Of course, each of the two transformers have to be

oversized to handle the same amount of power as three in a standard Δ configuration, but the

overall size, weight, and cost advantages are often worth it. Bear in mind, however, that with one

winding set missing from the Δ shape, this system no longer provides the fault tolerance of a

normal Δ-Δ system. If one of the two transformers were to fail, the load voltage and current

would definitely be affected.