Sei sulla pagina 1di 16

with wires of finite width, a resistor as a rod of carbon, an inductor as a

coiled wire, and the capacitor as a pair of metal plates. We shall examine
briefly each lumped component model from a field theory perspective.

5 Relationship between Field Theory


and Circuit Theory
(ref: Ramo et al.)
At lower frequencies where physical circuit dimensions are small compared
to the wavelength1 of electromagnetic waves, the behaviour of circuits is accurately modelled using lumped elementcomponent models, together with
Kirchhoffs laws. At higher frequencies where the distances between components are a significant fraction of a wavelength and greater, the signals carrying information or power from one place in a circuit to another are treated
as waves. Signals must be routed from one point to another using transmission lines, modelled using transmission line theory. If the component
dimensions be comparable to the wavelength then accurate understanding
and prediction of behaviour may require modelling using electromagnetic
field and wave theory.
In this section we examine the relationship between Maxwells equations
and circuit theory. Both Kirchhoffs voltage law, relating to voltage drops
Vi around a loop
N
X
Vi = 0
i=1

and Kirchhoffs current law, relating relating currents Ii leaving a node


N
X

Ii = 0

can be explained in terms of field theory.


Consider for example, the circuit shown below shown firstly using standard
circuit symbols for R, L and C, and secondly as a physical representation
f
= c/f

50 Hz
6000 km

100 kHz
3 km

1 MHz
300 m

10 MHz
30 m

5-1

100 MHz
3m

I(t)
R

2
L

V(t)

0
0

3
C

1
I(t)

V(t)

3
0

Coil

3
C

Figure 5.1: Circuit diagram showing component symbols (above) and a more physical depiction of the components (below).

5.1 Resistors

i=1

1 GHz
30 cm

10 GHz
3 cm

A resistor can be constructed from a resistive material of conductivity ,


length l and cross-sectional area A, as depicted below.

100 GHz
3 mm

5-2

AJW, EEE3055F, UCT 2011

Area A
1

1
0
0
1
0
1

0
1
0
1

J = E

where has units [Sm1] and is a property of the medium.

l
Figure 5.2: Resistor made from a cylinder of carbon. A current flows as a consequence of
the (axial component) of the electric field.

If the material is subjected to an electric field orientated along the length


of the cylinder, a current will flow, explained as follows:
Electrons move under the influence of the electric field to reach an
average drift velocity.
A classical model explains this as follows:
Electrons initially accelerate under the influence of the field, but repeatedly collide with bound atoms, and bounce off, resulting in deceleration. The net result is a constant average velocity for the electrons.
This has some analogy to the terminal velocity reached by a falling
object as a results of the resistance from the air molecules.
Area A

1
0
0
1
0
1
0
1
0
1

The average current density in Am2 is proportional to the strength of


the electric field, i.e.
J = E
It is noted that any particular electron continually accelerates and decelerates with each collision as illustrated.
Energy is dissipated as a result of the collisions (i.e. in the form of
heat).
The voltage developed across the resistor is found by integrating the electric
field through it from node 1 to node 2:
Z 2
Z 2
Z 2
Z 2
J
l
I/A
Edl =
Edl =
V21 = V2 V1 =
dl =
dl = I

A
1
1
1
1
The constant

Average drive velocity


of the electrons

is identified as the resistance of the rod, i.e.


R=

l
A

In the labelled circuit loop,


V21 = V2 V1 = IR

J = E
e

l
A

Path of an accelerating
electron, which collides
with atoms.

Figure 5.3: The electrons accelerate, but are impeded by the atomic structure, hence reaching a finite (average) terminal velocity. An imagined path of a single electron
is shown.

Because of the high density of electrons, the average drift speed is surprisingly slow. For example, Halliday, Resnick & Walker 6th Ed, do an
example calculation (Problem 27.3) in which the drift velocity with a copper wire or radius 0.9 mm, carrying a current of 17 mA, is calculated to be
4.9 107m/s or 1.8 mm/hr.

5.1.1 Calculation of average drift velocity


To calculate the average drift velocity of the electrons in a cylindrical wire
conductor, one needs to know the current I, the charge on one electron
(qe = 1.6 1019 coulombs), the thickness of the wire, and the electron
density e [m3] (not to be confused with charge density). Copper contains
e = 8.5 1028 electrons per m3 - a very large number!
Imagine a slug of electrons moving along a wire rod in the x direction.
The electron drift velocity in metres per second (in the x direction) can be
expressed in terms of the current as
vdrif t =

dx
dx dQ
1
=
= dQ (I)
dt
dQ dt
dx

where

5-3

AJW, EEE3055F, UCT 2011

5-4

AJW, EEE3055F, UCT 2011

dx
dt

is the velocity of the leading edge of the slug as it passes some point

x0 ,
dQ
dx

is the ratio of charge passing the point x0 per distance dx moved


in the x direction. This is identical the charge density in coulombs per
metre of wire.

dQ
dt

is the charge per second passing point x0 per second. If the conventional current is I amperes moving in the negative x direction, then
dQ
dt = I.

, consider a wire segment of length dx, and of thickness 2r.


To determine dQ
dx
The volume of the segment is r2 dx. The number of coulombs per metre is
(e qe )(r2dx)
dQ (charge/vol) vol
=
=
= e qer2
dx
length
dx
and hence
vdrif t =

I
e qe r2

dQ
= e qe r2 = 8.5 1028 (1.6 1019) 3.14 (0.5 103)2 = 10676 Cm1
dx
If a current of I = 1 ampere flows in the wire, the electron drift velocity is
1 [Cs1 ]
I
=
= 9.37 105ms1 = 34 cm/hr.
e qe r2
10676 [Cm1]

The number of electrons making up one coulomb is 1/ |qe | = 6.25 1018.


Thus for a 1 ampere current, 6.25 1018 electrons pass per second.

5.2 Capacitors
Consider a parallel plate capacitor. As the current flows through the wires,
a surface charge builds up on the inner sides of the capacitors plates.

5-5

The charge Q on the plate onto which the conventional current flows is
found by integrating the current flowing onto the plate, i.e.
Z t
Q(t) =
I(t)dt + Q0
t0

where Q0 is the initial charge at some starting time t0 . The other plate will
have a charge of Q(t).
A potential difference builds up between the places. The potential difference can be shown by careful argument [Griffiths] to be proportional to the
charge Q on the plates, i.e.
1
Q(t)
C
where C is the constant known as the capacitance.
Substituting for Q(t), we get
Z
1 t
Q0
Vc (t) =
I(t)dt +
C t0
C
Vc (t) =

For copper wire of 1 mm thickness,

vdrif t =

Note that all excess charge will sit on the surface of the capacitor
plates, in a thin layer (not inside the metal). Recall that E = 0 inside
a perfect conductor, and since div D = or div E = this implies,
= 0 inside the conductor. All excess charge must therefore lie on the
surface, described by a surface charge density s in Cm2.

AJW, EEE3055F, UCT 2011

In the circuit loop, V03 = V0 V3 = Vc (t).

5.3 Inductors
Inductors are made by winding several turns of wire either in air, or around
some high permeability material (which boosts the inductance, requiring
fewer turns).
We shall explain the operation of an inductor by considering first a single
turn, and then a coil of several turns, in the context of the series circuit
under analysis.
As already discussed, we are interested in applying Faradays law around
the dashed loop shown in the physical circuit. For the inductor, we are

5-6

AJW, EEE3055F, UCT 2011

interested in the integral of the electric field through the air gap between
the terminals as indicated by the dotted line between nodes 2 and 3 in the
circuit.

is known as the inductance, i.e.


Z
or

5.3.1 Single-turn Inductor

Integration
contour C

I
+

B(t) dS M
=
I(t)
I
3
The units of inductance are henrys [H], The voltage across the inductor is
Z
dI(t)
d
d[LI]
=L
V =
B dS =
dt S
dt
dt
In the labelled series circuit,
dI(t)
V32 = V3 V2 = L
dt
L=

Consider a single turn inductor that forms part of the series circuit under
analysis.

B dS = LI

S
S

5.3.2 Multi-turn Inductor


A multi-turn inductor is constructed by winding a coil of wire as depicted
in the figure below.

Figure 5.4: Single turn inductor.

We can apply Faradays law locally to a closed contour C that goes clockwise
around the inside of the wire and then across the air gap (in the shape of
the dotted path),
I
Z
Z
Z
M
d
E dl =
B dS =
E dl +
E dl =
dt S
t
C
(air)
(wire)
where and dS points into the page, and M is the flux threading the integration loop (and cutting a chosen surface S, bounded by C).
Since E = 0 in the wire, the potential different is then
Z
Z
d
B dS
V =
E dl =
dt S
(air)
The magnetic field B is linearly
R proportional to the current I flowing in the
2
wire , i.e. B I, and so is S B dS I. The constant of proportionality
2

The magnetic field vector owing to a short current segment can be computed using the Biot-Savart
law (reviewed in Section 2.5). The total field is found by integration of all contributions from current
elements in the wire.

5-7

AJW, EEE3055F, UCT 2011

I
+

V3
V

V2
V1

Figure 5.5: Multiturn inductor (N = 3 here).

The voltage drop across the terminals is the line integral along the dashed
line:
Z +
E dl
V =
Z
Z
Z
= [
E dl +
E dl + ...
E dl]
gap1

gap2

gapN

= V1 + V2 + ...VN
3

The units of inductance are equivalently [H] = [Wb A1 ] = [VA1 s1 ] = [VCs2 ] = [NC1 mCs2 ] =
[Nms2 ].

5-8

AJW, EEE3055F, UCT 2011

If we further assume that the flux linking each turn is the same, then
V1 =

d
= V2 = ... = VN
dt

and

d
V = N V1 = N
dt
Because there are N turns, the flux theading the coil will be N times
stronger than the contribution from a single turn, i.e.
= N 1turn
where 1turn is the flux contribution from a single turn. Substituting, we
obtain
d
d(L1turnI)
d
dI
V = N V1 = N
= N 2 1turn = N 2
= N 2 L1turn
dt
dt
dt
dt
Thus the inductance for a tightly wound N -turn coil is

spiral staircase winding around an imaginary centre line. The total flux
M passing through S is proportional to the current I in the wire, and is
given by
Z
M =

B dS = LI

where L is the inductance of the multi-turn coil. Thus


R
B(t) dS
M
L=
= S
I
I(t)
where M must be carefully evaluated for the particular coil structure. For
a short, compact coil of N turns, it is not difficult to show that
L = N 2 L1turn
where L1turn is the inductance of a single turn coil of the same radius. To
see this, one must grasp two points:

where L1turn is the inductance of a single turn.

the spiral surface S through with the flux lines pass consists of a stack
of N identical contributing flatish discs (the total surface area is approximately N times larger than for a single turn)

It is worth remembering that the inductance increases as a function of N 2 .


If one doubles the number of turns, the inductances increases by a factor of
four.

the flux density on each component disc is N times stronger than the
flux density generated by a single turn (superposition of contributions
from N turns, each carrying current I)

L = N 2 L1turn

Alternative Explanation

Analysis of a multi-turn coil is similar to the case of a single turn coil, with
the added complication that the integration contour C is not a circle, but
rather made up of a spiral that follows the wire (and a short section in the
air gap between its terminals). As argued for a single turn case, the voltage
across the terminals is
Z
Z
d
B dS
V =
E dl =
dt S
(air)

Thus the total flux threading


the surface of the multi-turn contour is
Z

M = B dS N N 1turn
S

where 1turn is equal to the flux generated by a single turn coil carrying
current I.
The inductance of the multi-turn coil is then
L=

N 2 1turn
M
=
= N 2 L1turn
I
I

where surface S is now a complicated-to-visualise surface that is bounded


by the contour C. It helps to imagine the surface within the coil as a smooth

5-9

AJW, EEE3055F, UCT 2011

5-10

AJW, EEE3055F, UCT 2011

5.4 Formulas for Practical Coils


In order to calculate the inductance accurately, we need to consider both the
field inside, and outside of the wire making up the coil. The total inductance
is usually computed in two parts:
L = Linternal + Lexternal
where Linternal is the contribution arising from the magnetic field within the
wire, and Lexternal is the contribution from the field outside the wire.

Example Calculation

Calculate the inductance of a circular copper wire ring, radius r = 10 cm,


wire radius a = 0.5 mm. NOTE 0 in copper.
The external component is


 
8 0.1
8r
7
2] = 4 10 0.1[ln
2] = 0.7 106 H
Lext 0 r[ln
a
0.0005

5.4.1 Inductance of a Circular Wire Ring

wire
diameter
2a

for the case where a << r.


The internal inductance of a long straight wire (in henrys per metre of
wire), assuming uniform current density in the wire4, can be shown to be
independent of wire radius a, and for non-magnetic metal wire is given by
(ref Ramo et al.)
0
Lint =
= 0.5 107 Hm1
8

External field
in the surrounding
air.
wire

2a

The internal inductance in henrys per metre is

radius
of circle

Lint =
Internal field
inside the metal

The total internal inductance


Lint Lint2r = 0.050 106 0.63 = 0.03 106 H

Figure 5.6: Circular wire ring, and its cross section.

which is relatively small compared to Lext . The total inductance is

The inductance of a wire ring can be found from


R
B dS M
=
L= S
I
I

L = Lint + Lext = 0.73 106 H

where B can be found by integrating the field contributions (using the BiotSavart law, reviewed in Section 2.5) from each elemental current segment
around the ring.
Considering only the contribution from the flux outside the wire, it can
be shown, by integration, that M Ir[ln 8r
a 2] and hence the external
inductance (Ramo et al.) is
 
M
8r
Lext =
0 r[ln
2]
I
a
5-11

4 107
0
=
= 0.05 106 Hm1
8
8

AJW, EEE3055F, UCT 2011

At sufficiently low frequencies (kHz down to DC), the current density is fairly uniform across the cross
section of the wire. As the frequency increases from DC, the current tends to concentrate increasingly
towards the outside of the wire. This effect is known as the skin effect, and is discussed later in the
course.

5-12

AJW, EEE3055F, UCT 2011

5.4.2 Inductance of a Short Coil (short length to radius ratio)

5.4.3 Inductance of a Long Multi-turn Coil (long length to radius


ratio)

Radius

r
Integration contour
length l

B field

Radius

Figure 5.7: Short inductor.


I

B ~ 0 outside

For a short length to radius ratio, the external inductance of an N turn coil
is N 2 times that of a single turn, i.e.
 
8r
2
2]
L N 0 r[ln
a
When winding a coil, it is useful to remember that the inductance is
proportional to the square of the number of turns. The inductance may be
increased by winding the coil on an iron or ferrite rod or on a toroid, which
has a relative permeability of hundreds or thousands that of air.

Figure 5.8: Long inductor.

For a long coil of length l, containing N turns carrying current I, the coil
can be treated as a wrapped current sheet (row of dots in the illustraHtion), fromRwhich H can easily be obtained by application of Amperes law
H dl = S J dS. Consider the integration contour shown in the sketch.
The vertical side contributions to the integral are negligibly small because
the flux lines are perpendicular to the contour. Outside the coil, the flux
lines spread out, and H becomes negligibly small compared to inside the
coil. Consequently, we can also ignore
the horizontal segment of the integral
H
on the outside of the coil. Thus H dl H l where H is the magnetic field
inside
R the coil. The total current passing through the integration contour
is S J dS = N I.
Thus Amperes law implies H l N I and from which
H N I/l
and
B 0 N I/l

5-13

AJW, EEE3055F, UCT 2011

5-14

AJW, EEE3055F, UCT 2011

An additional point to note is that this result is independent of the exact


position of the the horizontal segment of the contour within the coil. This
implies that the field intensity is uniform in a cross section of a long coil,
and hence the flux through the cross section of coil is = Br2 = 0 Hr2
(i.e. linking one turn of the coil)
The inductance is the ratio of the total flux linking the coil to the current,
being
R
B dS N N Br2
N (0 N I/l)r2
0 r2 N 2
L= S
=
=

=
I
I
I
I
l
5.4.4 Inductance of an Intermediate Length Multi-turn Coil

Consider two coils in close proximity, one containing N1 turns, and the
other containing N2 turns. Although not shown in the sketch, these coils
are parts of circuits and carry currents.
Let 1 (t) be the component of the flux threading coil 1, resulting purely
from the current I1(t) flowing in coil 1. Let the 1 f rom 2 be the flux threading
coil 1 arising from the current I2(t) in coil 2. From Faradays law, the voltage
across the terminals of coil 1 is
V1 (t) =

where N11 is the total flux threading the multi-turn coiled surface. Since
1 is proportional to I1 and 1 f rom 2 is proportional to I2, we have,

In cases where the coil can neither be considered very long or very short,
the following approximate formula is commonly used:
0 r2N 2
l + 0.9r
The formula incorporates an empirical correction factor (+0.9r) in the denominator.
L

5.5 Mutual Inductance

d (N11 + N11 f rom 2 )


d1
d1 f rom 2
= N1
+ N1
dt
dt
dt

V1 (t) = L1

dI2
dI1
+ M12
dt
dt

where L1 is the self inductance constant of coil 1, and M12 is another cond f rom 2
= M12 dIdt2 ,
stant. Since N1 1 dt
M12 =

N11 f rom 2
I2

Similarly, the voltage across the terminals of coil 2 is

If two wire coils are close to one another, flux resulting from current flowing
in one coil will thread the coil of the other. Thus a changing current in one
coil, will result in a changing flux in the other and hence induce a voltage
across its terminals. This is the basis of a transformer.

V1 (t)

V2 (t)
I1 (t)

d (N22 + N22 f rom 1 )


d2
d2 f rom 1
= N2
+ N2
dt
dt
dt

and

dI1
dI2
+ M21
dt
dt
where L2 is the self inductance constant of coil 2, and M21 is a constant:
V2 (t) = L2

2 + 2f rom1

1 + 1f rom2

V2 (t) =

I2 (t)

M21 =

N22 f rom 1
I1

It can be shown (consult more detailed texts), that regardless of the geometry, M12 = M21. The constant M = M12 = M21 must have the same

Figure 5.9: Coupled coils.

5-15

AJW, EEE3055F, UCT 2011

5-16

AJW, EEE3055F, UCT 2011

units of L1 and L2 being henrys, and is known as the mutual inductance


between the coils5.
Mutual Inductance for Tightly Coupled Coils
A special case is when the coils are tightly coupled, e.g. stacked on top
of one another such that 1 f rom 2 = 2 and 2 f rom 1 = 1 (or coils would
around a common torroid). For this case,


d1 f rom 2
d2
L2 dI2
dI2
= N1
= N1
= N1
M12
dt
dt
dt
N2 dt
and hence M12 =

N1
N2 L2 .

Similarly, M21 =

5.6 Kirchhoffs Voltage Law


The relationship between Kirchhoffs law for a lumped element circuit model
and the physical component layout, is established by application of Faradays law.
Consider applying Faradays law to the closed contour indicated by the
dotted line in the following physical circuit representation. Define dS pointing into the page, which implies the integration direction is clockwise.
1

N2
N1 L1 .

M=
for tightly coupled coils.
The voltage ratio is

V32
3

V03

L1 L2

V10

V(t)

Coil

3
C

p
L1 dIdt1 + M dIdt2
N12 L1turn dIdt1 + N12 L1turnN22 L1turn dIdt2
V1
N1
p
= dI2
=
=
dI
dI
dI
V2
N2
L2 dt + M dt1
N22 L1turn dt2 + N12 L1turnN22 L1turn dt1

which is a well known result for tightly coupled coils.

V21

N1 N2
L2 L1 = L1 L2
N2 N1

or

I(t)

Since M = M12 = M21 , the product M12M21 yields


M2 =

Figure 5.10: Series circuit loop - Faradays law is applied along the dotted line to derive
Kirchhoffs voltage law.

Since E 0 in the wires, the voltage drops around the circuit occur
across the components. Thus, we can write
Z
Z 0
Z 3
Z 2
I
Z 1
B
dM
Edl =
Edl
Edl
Edl
Edl =
dS =
dt
S t
3
2
1
0
or
dM
V10 + V21 + V32 + V03 =
dt
The flux threading the loop can be split into three contributions:

One could in principle, calculate M12 = 1 1If2rom 2 for a given coil geometry by doing a surface inteR
gration of B(i2 ) dS to obtain 1 f rom 2 (I2 ) = S1 B(i2 ) dS. B(i2 ) can be obtained directly from the
Biot-Savart law (which requires a contour integration along coil 2). There is a better way to do it
(consult other texts for details).

5-17

AJW, EEE3055F, UCT 2011

M = applied + self + mutual


where

5-18

AJW, EEE3055F, UCT 2011

applied refers to any flux imposed on the circuit e.g. wave a bar magnet
past the circuit.

I
R

self refers to the flux generated by the current flowing in the circuit
loop itself (the circuit can be thought of as a single turn inductor).
self = Lself I where Lself is the self inductance of the loop, which
carries current I.
mutual refers any leakage flux from other parts of the circuit (notably
the inductive element) that threads the loop.

dS points into the page

applied
t

Substituting the lumped element relationships derived above,


Z
1 t
dM
dI
V (t) IR
I(t)dt L =
C t0
dt
dt
P
Kirchhoffs law N
i=1 Vi = 0 describes the circuit model, and hence we
must introduce additional model component(s) into the circuit model
to account for the term ddtM .
It is noted that the term ddtM will modify the current flowing in the
circuit, and should be included for accurate prediction of the behaviour
of the circuit.
In practice however, this term is usually small compared to the other
terms, and is often neglected in practical circuit design.
For the complete model shown below, the equation in the form of Kirchhoffs
law is written as
Z
1 t
dI dapplied dself dmutual
V (t) IR

=0
I(t)dt L
C t0
dt
dt
dt
dt
or

N
X

Vi = 0

i=1

5-19

AJW, EEE3055F, UCT 2011

V (t)

(the direction of positive flux)

Lself

C
Figure 5.11: Circuit modified to incorporate an additional series inductor Lself which models the series inductance of the loop, and an additional voltage source which
models unwanted external signals.
d

The term dtself = Lself dI


dt resulting from the current in the loop, is modelled by a (small) series inductance Lself . A feeling for the magnitude of
this self inductance can be gained by considering a circuit arranged in a
circular loop of radius 10 cm. We previously calculated the self inductance
of a wire ring of radius 10 cm and wire radius of 0.5 mm to be 0.7 H.
At an operating frequency of say 1 kHz, the AC reactance of this term
is XL = 2f Lself 4 103 ohms, which is usually small enough to be
neglected from calculations. At higher frequencies, this term may become
significant.
d
arises if the circuit is exposed to some externally genThe term applied
dt
erated AC field, e.g. a nearby transmitter like a cell phone, or perhaps a
motor, or close-by transformer. Usually this term can be neglected. Of

5-20

AJW, EEE3055F, UCT 2011

course radio waves are ever present, but their contribution is usually insignificant compared to the voltage levels in the circuit. Circuits can be
shielded from external sources by placing them in a metal enclosure known
as a Faraday cage 6.
in this context arises from the leakage flux from the
The term dmutual
dt
inductor L, and is in practice usually small compared to the voltage drop
across the (multi-turn) inductor. The net effect may either be to increase
or decrease the current in the circuit, depending on the physical orientation
of the inductor. R
NOTE: The flux S B dS requires the direction of dS to be defined. Since
the integral of E is taken clockwise around the loop, the right hand rule
tells us that dS points into the page. The flux will be a positive quantity if
B (threading the loop) points into the page.

5.7 Kirchhoffs Current Law at a Node


Consider the illustration in Figure 5.12 showing four wires connecting to a
node carrying currents I1, I2, I3 and I4. Kirchhoffs node current law states
that the sum of all currents leaving the node equals zero, i.e.
N
X

Ii = I1 + I2 + I3 + I4 = 0

i=1

I2

I1
I3

b
n

closed surface S1

dS

I4

closed surface S2

A Faraday cage will provide good shielding from DC electric fields. DC magnetic fields however do
penetrate metal enclosures. e.g. the earths magnetic field is still detected by a magnetic compass
within a Faraday cage. The degree of penetration of time-varying AC electromagnetic fields is a
function of a frequency dependent parameter of the metal known as the skin depth, which will be
studied later in this course. For good shielding at a particular frequency, the enclosure wall should be
considerably thicker that the skin depth (at that frequency).

5-21

AJW, EEE3055F, UCT 2011

Displacement current "flows" between plates

Figure 5.12: The relationship between Kirchhoffs cuurent law at a node the continuity
equation.

5-22

AJW, EEE3055F, UCT 2011

Consider now the continuity of charge relationship


Z
I
dQ
d
dV =
J dS =
dt V
dt
S
which states that the total conduction current leaving an arbitrary closed
surface S is equal to (minus) the rate of change of charge within the volume
V enclosed by S. One can re-express the continuity relationship in a form
that looks similar to Kirchhoffs law by moving the charge term to the left
hand side:
I
Z
d
J dS +
dV = 0
dt V
S
The 2nd term
R can beHexpressed as a surface integral over S by substituting
Gauss law dV = S D dS,
I
I
d
J dS +
D dS = 0
dt S
S
Moving the time derivative within the integral, the continuity equation becomes
I
I
D
dS = 0
J dS +
S t
S
which says that the sum of the conduction current Ic and the displacement
current Id leaving an arbitrary closed surface7 is zero. I.e. for any closed
surface S,
Ic + Id = 0
where
Ic =
and

S1

The displacement current is typically insignificant (there is no significant


charge build up within S1 , i.e.
I
dQ
D
dS =
0
Id =
dt
S1 t
Thus we have
I1 + I2 + I3 + I4 0
If one shrinks surface S1 to a tiny surface surrounding the node, the displacement current shrinks to zero and the relationship converges exactly to
Kirchhoffs law.
If however, we choose a surface S = S2 in such a way as to pass between the
plates of the capacitor as illustrated in Figure 5.12, then we have a slightly
more subtle situation.
As there is one less wire cutting the surface, the total conduction current is
I
Ic =
J dS = I1 + I2 + I3
S2

There is however a significant charge build-up on the plate(s) of the capacitor as a result of current I4. The charge Qplate on the plate (within S2 )
dQ
builds up at a rate of dtplate = I4.

J dS
S

D
dS
S t
Thus we have derived a generalised form of Kirchhoffs current law, which
can be applied to an arbitrary closed surface.
Id =

For example, consider the closed surface S = S1 surrounding the node in


Figure 5.12. There are N = 4 wires piercing the surface and joining at the
node.
For S1 , the total conduction current leaving the surface is
I
Ic =
J dS = I1 + I2 + I3 + I4

Thus the continuity relationship


I
dQ
J dS =
dt
S
applied to surface S2 becomes

It is worth noting that the total displacement current flowing


H out of a closed surface is equal to the time
rate of change of charge enclosed by the surface, i.e. Id = S D
t dS = dQ/dt.

5-23

AJW, EEE3055F, UCT 2011

I1 + I2 + I3

dQplate
= I4
dt

5-24

AJW, EEE3055F, UCT 2011

5.8 The Relaxation Time of Conducting Materials

Rearranging, we get
I1 + I2 + I3 + I4 0
which is consistent with the case where S = S1 and Kirchhoffs law.
The approximation ( ) is present in the above expression because a
small (and negligible) charge will exist on the surface8 of the conductors 1
to 4
Another way to look at the situation is to observe that the sum of all
currents, both conduction and displacement current, flowing out of a closed
surface is zero, i.e. for surface S4,
I1 + I2 + I3 + Id = 0
H

where Id = S D
t dS is the displacement current leaving the surface. Id
is concentrated primarily between the plates of the capacitor (where the
electric field is strongest).

The term conductor refers to a material that will carry current when subjected to an electric field. In solid materials, like metals, electrons are free
to move, and the net movement of electrons constitutes a current. In liquids
(e.g. a salt solution), charged ions in solution are free to move allowing a
current to exist. Insulating materials, in contrast, are materials for which
the electrons are tightly bound to particular atoms, and hence no current
can flow.
A perfect conductor is one for which there is an unlimited abundance
of free electrons. The conductivity of a perfect conductor is infinite - an
infinitesimal electric field will create a large current. Metals can often be
approximated as perfect conductors in the analysis of their behaviour under
certain conditions.
If a conducting object is placed in a stationary position within an electric
field, the electrons will, given time, rearrange themselves such that:

H
= S D
dS = Id (for any closed surface),
Also, since we have shown that dQ
dt
t
dQ
and if the only significant dt within S4 is the charge build up on the inner
plate of the capacitor due to I4, then
Id =

E goes to zero inside the conductor (electrons quickly re-arrange themselves until the total E = 0 inside conductor). Note: in the steady state
situation, the net force on the electrons must go to zero - the electrons
will rearrange themselves to achieve this. Since the conductor is not
moving (stationary), in the steady state situation, the magnetic force
on the electrons will be zero, and hence the electric force must also be
zero9.

dQ
dQplate

= I4
dt
dt

which again for S4 implies

The charge density = 0 inside the conductor (since div D = and


E = 0, it means = 0) .

I1 + I2 + I3 + I4 0

Any net charge (excess charge) resides on the surface in an infinitesimally thin layer (we refer to this charge as a surface charge).
9

Any excess charge must be the surface, because 0 very rapidly inside a metal conductor - see section
on relaxation time.

5-25

AJW, EEE3055F, UCT 2011

If a conductor is moving through a static magnetic field - then the E field inside the metal can be
non-zero - electrons will always rearrange themselves such that the sum of the magnetic and electric
forces equals zero. For example, a rod of length l moving at velocity v through a static magnetic field
B will experience a magnetic force on the electrons F = qv B. Electrons will re-arrange themselves
such that the total force on an electron of chage q is qv B + qE = 0, i.e. inside the metal, E = v B
once the electrons have rearranged themselves. There will also exist a potential difference between the
Rb
Rb
end points of the rod, i.e. b a = a E dl = a (v B) dl. If v is perpendicular to B and the rod
is orientated such that its length is perpendicular to v and B, then the potential difference between
the ends will be vBl.

5-26

AJW, EEE3055F, UCT 2011

The potential (x, y, z) is constant throughout the conductor (since


the electric field is zero inside).
E is perpendicular at the boundary (i.e. no tangential component).
In practice, one might wonder just how long it would take for the electrons
to re-arrange themselves. Imagine setting up an arrangement of charge
(x, y, z) within a homogeneous conducting material and then releasing the
charge at some instant. The charge will redistribute itself such that the
electric field goes to zero at every point within the conductor10. This happens very rapidly in metals, so fast that it can be considered instantaneous
in many practical situations.
Consider the charge within the conducting body. The movement of charge
will be governed by the continuity equation, for which the differential form
J=

describes the relationship between current leaving a small volume element,


and the rate of change of charge within the element. If we substitute
J = E, we get

(E) =
t
and then eliminate E via Gauss law ( E = /) we obtain a first order
differential equation

+
=0

t
which has solution

(t) = 0 e t
where 0 is the initial charge density at time t = 0.
Thus the charge density at any point within the material will dissipate
to zero with an exponential decay. The decay curve is characterised by the
time constant = , also known as the relaxation time, which is the time
at which the charge density has reduced to e1 36.8% of its initial value.
After 5 the charge density will have decayed to less than 1% of the initial
value.
10

Here we are ignoring the granularity of electrons and treating the charge as a kind of fluid.

5-27

AJW, EEE3055F, UCT 2011

To see how quickly this happens in practice, the time constant may be
calculated for various materials. For example for a metal conductor like
copper ( = 5.8 107 Sm1, 0 = 8.85 1012 Fm1), the time
constant is = = 1.5 1019 s, which is extremely short compared to
say the period of a 100 GHz microwave sinusoid, being 1011 seconds. In
electronic circuits, the charge in the wires rearranges itself very quickly in
response to the dynamics of the circuits (i.e. to a very good approximation,
we consider E 0 and 0 inside the connecting copper wires - a small
component of E must however exist to drive the current).
For a weakly conducting liquid like tap water ( 102 Sm1, 810
Fm1), the relaxation time is about 70 109 s. For a good insulator like
glass (e.g. = 1014 Sm1, = 50), the relaxation time is calculated to
be about 4000 seconds (67 minutes).
Exercise: Calculate the relaxation time of iron ( = 0.9 107 Sm1).

5.9 Shielding and The Faraday Cage


Circuitry may be shielded from external electric fields by enclosing the circuit inside a metal box known as a Faraday cage. External electric fields
have no influence on the circuitry within a box made from a perfect conductor - an electrically quiet zone exists within the box. On a larger scale,
Faraday cages can provide protection against lightning strikes. Low frequency magnetic fields can however penetrate a real metal enclosure and
influence the circuitry inside it. Try for yourself to see if a permanent magnet is able to attract iron pieces through the walls of a metal box. DC
magnetic fields penetrate through metal enclosures. For example, a magnetic compass will still detect the earths magnetic field inside a Faraday
cage. Effective shielding from a magnetic field at 50 Hz requires a thick
wall (several mm), preferably made from a high permeability material. As
the frequency increases, a metal wall becomes more effective in attenuating
magnetic fields. In the MHz range and higher, metal enclosures are very
effective (if well sealed) for shielding circuits from external electromagnetic
fields, and also for preventing radiation leakage from the circuitry within
the enclosure. Electromagnetic waves reflect off the enclosure, and what

5-28

AJW, EEE3055F, UCT 2011

does enter the metal, decays exponentially with a decay constant called the
skin depth (covered later in the course).

dm /dt

V (t)

5.10 Twisted Pair Cables


An interesting application of field theory concerns the understanding of
how twin-wire transmission lines are influenced by electric and magnetic
fields. The following figure depicts a parallel wire (non twisted) transmission
line, which could be used to carry a signal from one location to another.
Such parallel wire transmission lines are particularly susceptible to inductive
coupling of magnetic fields, especially when several signals need to be carried
in the same bundle. Changing magnetic flux d/dt within the circuit loop
induces an additional voltage which adds to the signal voltage in accordance
with Faradays law. A clever solution to minimising inductive coupling is
to reduce the net flux, by twisting the pair of wires as illustrated. The
flux contributions B dS in adjacent twists are opposite in polarity and
will tend to cancel, resulting in reduced d/dt and hence reduced magnetic
field interference. This type of cable is known as twisted pair and is very
commonly used for data networks.
In addition to the minimization of magnetic coupling, the twisting also
improves the immunity to capacitive coupling. If, for example, the cable
lies close to a conductor that is varying in potential relative to ground, like
the live wire wire 50 Hz mains supply, this 50 Hz signal will capacitively
couple to the conductors (imagine small valued capacitors between the 50
Hz conductor and cables two wires). Twisting the cable, creates a more
symmetrical coupling arrangement, independent of the of the orientation of
the pair, hence causing the effect to be common to both wires. A differential
amplifier at the receiver with a high common-mode rejection ratio extracts
the desired differential signal, and removes the common capacitively coupled
interference.

V (t)
net dm /dt 0
Figure 5.13:

Figure 5.14: Illustrations of capacitive coupling onto parallel wire transmission lines for the
case of balanced versus unbalanced driving circuitry.

5-29

AJW, EEE3055F, UCT 2011

5-30

AJW, EEE3055F, UCT 2011

Sometimes twisted pairs are also shielded, offering increased immunity to


electromagnetic interference and noise. The shielding also further reduces
radiation from the cable itself. Several twisted pairs are sometimes bundled
within the same cable. The use of twisted pairs offers significantly lower
cross talk between data channels compared to non-twisted side-by-side
wires within a cable.
Unshielded twisted pair (UTP) cable is now used for connecting standard
PCs in in-door local area networks (LANs). UTP network cables replaced
previously used 50 ohm coaxial cables for LANs because UTP cables are
cheaper to manufacture than coaxial cable, and offer adequate immunity
to electromagnetic interference. UTP network cable is typically used for
distances up 100m.
The Cat-5e series cable is the cable commonly used for PC LANs (for both
100 Mbit/s and gigabit ethernet networks), and is designed to carry frequencies up to 100 MHz. PC LAN network cables contain four unshielded twisted
pairs, with RJ-45 connectors on each end. The characteristic impedance of
Cat-5e is 100 ohms.
Mechanical arrangement within the cable can further reduce coupling between pairs. For example the Power Cat-6 four pair cable sold by RS
Electronics contains four unshielded twisted pair (UTP) cables, with a central separator, and is designed to support high speed data transmission
systems (frequencies up to 250 MHz).

Cat-5e

Cat-6

RJ-45 connector
Figure 5.15: Photos from from RS website; connector from Intel website

5-31

AJW, EEE3055F, UCT 2011

5-32

AJW, EEE3055F, UCT 2011

Potrebbero piacerti anche