You are on page 1of 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 1 2

I(t)
and Circuit Theory R 2
1
L
(ref: Ramo et al.) V(t)

At lower frequencies where physical circuit dimensions are small compared 3


0
to the wavelength1 of electromagnetic waves, the behaviour of circuits is ac-
curately modelled using “lumped element”component models, together with
0 3
Kirchhoff’s laws. At higher frequencies where the distances between compo-
nents are a significant fraction of a wavelength and greater, the signals car- C
rying information or power from one place in a circuit to another are treated 1 2
as waves. Signals must be routed from one point to another using trans- I(t)
mission lines, modelled using transmission line theory. If the component R
1 L
dimensions be comparable to the wavelength then accurate understanding 2
and prediction of behaviour may require modelling using electromagnetic V(t)
field and wave theory. 3
In this section we examine the relationship between Maxwell’s equations 0 Coil
and circuit theory. Both Kirchhoff’s voltage law, relating to voltage drops
Vi around a loop 0 3
XN
Vi = 0 C
i=1
Figure 5.1: Circuit diagram showing component symbols (above) and a more physical de-
and Kirchhoff’s current law, relating relating currents Ii leaving a node
piction of the components (below).
N
X
Ii = 0
i=1 5.1 Resistors
can be explained in terms of field theory.
Consider for example, the circuit shown below shown firstly using standard A resistor can be constructed from a resistive material of conductivity σ,
circuit symbols for R, L and C, and secondly as a physical representation length l and cross-sectional area A, as depicted below.

1 f 50 Hz 100 kHz 1 MHz 10 MHz 100 MHz 1 GHz 10 GHz 100 GHz
λ = c/f 6000 km 3 km 300 m 30 m 3m 30 cm 3 cm 3 mm

5-1 5-2 AJW, EEE3055F, UCT 2011


Area A E J = σE • The average current density in Am−2 is proportional to the strength of
1
0
0
1 the electric field, i.e.
1 0
1
0
1σ 2 J = σE
e
0
1
where σ has units [Sm−1] and is a property of the medium.
l
• It is noted that any particular electron continually accelerates and de-
Figure 5.2: Resistor made from a cylinder of carbon. A current flows as a consequence of
the (axial component) of the electric field.
celerates with each collision as illustrated.
• Energy is dissipated as a result of the collisions (i.e. in the form of
If the material is subjected to an electric field orientated along the length heat).
of the cylinder, a current will flow, explained as follows:
The voltage developed across the resistor is found by integrating the electric
• Electrons move under the influence of the electric field to reach an field through it from node 1 to node 2:
average drift velocity. Z 2 Z 2 Z 2 Z 2
J I/A l
• A classical model explains this as follows: V21 = V2 −V1 = − E·dl = − Edl = − dl = − dl = −I
1 1 1 σ 1 σ σA
Electrons initially accelerate under the influence of the field, but repeat- l
edly collide with bound atoms, and “bounce off”, resulting in deceler- The constant σA
is identified as the resistance of the rod, i.e.
ation. The net result is a constant average velocity for the electrons. l
R=
This has some analogy to the terminal velocity reached by a falling σA
object as a results of the resistance from the air molecules. In the labelled circuit loop,

Area A E J = σE V21 = V2 − V1 = −IR


1
0
0
1
0
1
0
1 e Path of an accelerating 5.1.1 Calculation of average drift velocity
0
1 electron, which collides
Average drive velocity with atoms. To calculate the average drift velocity of the electrons in a cylindrical wire
of the electrons conductor, one needs to know the current I, the charge on one electron
Figure 5.3: The electrons accelerate, but are impeded by the atomic structure, hence reach- (qe = −1.6 × 10−19 coulombs), the thickness of the wire, and the electron
ing a finite (average) terminal velocity. An imagined path of a single electron density ρe [m−3] (not to be confused with charge density). Copper contains
is shown. ρe = 8.5 × 1028 electrons per m3 - a very large number!
Imagine a slug of electrons moving along a wire rod in the x direction.
Because of the high density of electrons, the average drift speed is sur- The electron drift velocity in metres per second (in the x direction) can be
prisingly slow. For example, Halliday, Resnick & Walker 6th Ed, do an expressed in terms of the current as
example calculation (Problem 27.3) in which the drift velocity with a cop-
dx dx dQ 1
per wire or radius 0.9 mm, carrying a current of 17 mA, is calculated to be vdrif t = = = dQ (−I)
dt dQ dt
4.9 × 10−7m/s or 1.8 mm/hr. 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 • Note that all excess charge will sit on the surface of the capacitor
x0 , plates, in a thin layer (not inside the metal). Recall that E = 0 inside
dQ a perfect conductor, and since div D = ρ or div εE = ρ this implies,
• is the ratio of charge passing the point x0 per distance dx moved
dx ρ = 0 inside the conductor. All excess charge must therefore lie on the
in the x direction. This is identical the charge density in coulombs per
surface, described by a surface charge density ρs in Cm−2.
metre of wire.
dQ
The charge Q on the plate onto which the conventional current flows is
• dt is the charge per second passing point x0 per second. If the conven- found by integrating the current flowing onto the plate, i.e.
tional current is I amperes moving in the negative x direction, then Z t
dQ
dt = −I. Q(t) = I(t)dt + Q0
t0
To determine dQ
dx
, consider a wire segment of length dx, and of thickness 2r.
where Q0 is the initial charge at some starting time t0 . The other plate will
The volume of the segment is πr2 dx. The number of coulombs per metre is
have a charge of −Q(t).
dQ (charge/vol) × vol (ρe qe )(πr2dx) A potential difference builds up between the places. The potential differ-
= = = ρe qeπr2
dx length dx ence can be shown by careful argument [Griffiths] to be proportional to the
charge Q on the plates, i.e.
and hence
−I 1
vdrif t = Vc (t) = Q(t)
ρe qe πr2 C
For copper wire of 1 mm thickness, where C is the constant known as the capacitance.
dQ Substituting for Q(t), we get
= ρe qe πr2 = 8.5 × 1028 ∗ (−1.6 × 10−19) ∗ 3.14 ∗ (0.5 × 10−3)2 = −10676 Cm−1 Z
dx 1 t Q0
Vc (t) = I(t)dt +
If a current of I = 1 ampere flows in the wire, the electron drift velocity is C t0 C

−I −1 [Cs−1 ] In the circuit loop, V03 = V0 − V3 = −Vc (t).


vdrif t = = = 9.37 × 10−5ms−1 = 34 cm/hr.
ρe qe πr2 −10676 [Cm−1]
5.3 Inductors
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. Inductors are made by winding several turns of wire either in air, or around
some high permeability material (which boosts the inductance, requiring
5.2 Capacitors fewer turns).
We shall explain the operation of an inductor by considering first a single
Consider a parallel plate capacitor. As the current flows through the wires, turn, and then a coil of several turns, in the context of the series circuit
a surface charge builds up on the inner sides of the capacitor’s plates. under analysis.
As already discussed, we are interested in applying Faraday’s law around
the dashed loop shown in the physical circuit. For the inductor, we are

5-5 AJW, EEE3055F, UCT 2011 5-6 AJW, EEE3055F, UCT 2011
interested in the integral of the electric field through the air gap between is known as the inductance, i.e.
Z
the terminals as indicated by the dotted line between nodes 2 and 3 in the
circuit. B · dS = LI
S
or R
5.3.1 Single-turn Inductor B(t) · dS ΨM S
L= =
I(t) I
Consider a single turn inductor that forms part of the series circuit under 3
The units of inductance are henrys [H], The voltage across the inductor is
analysis. Z
d d[LI] dI(t)
V = B · dS = =L
B
Integration dt S dt dt
contour C B
In the labelled series circuit,
I
+
dI(t)
V32 = V3 − V2 = −L
V dt

5.3.2 Multi-turn Inductor

Figure 5.4: Single turn inductor. A multi-turn inductor is constructed by winding a coil of wire as depicted
in the figure below.
We can apply Faraday’s law locally to a closed contour C that goes clockwise I
+
around the inside of the wire and then across the air gap (in the shape of
V3
the dotted path),
V
I Z Z Z V2
d −∂ΨM
E · dl = E · dl + E · dl = − B · dS = V1
C (air) (wire) dt S ∂t

where and dS points into the page, and ΨM is the flux threading the inte- Figure 5.5: Multiturn inductor (N = 3 here).
gration loop (and cutting a chosen surface S, bounded by C).
Since E = 0 in the wire, the potential different is then The voltage drop across the terminals is the line integral along the dashed
Z Z line:
d Z +
V =− E · dl = B · dS
(air) dt S V =− E · dl
The magnetic field B is linearly Z− Z Z
2
R proportional to the current I flowing in the = −[ E · dl + E · dl + ... E · dl]
wire , i.e. B ∝ I, and so is S B · dS ∝ I. The constant of proportionality
gap1 gap2 gapN
= V1 + V2 + ...VN
2
The magnetic field vector owing to a short current segment can be computed using the Biot-Savart
3
law (reviewed in Section 2.5). The total field is found by integration of all contributions from current The units of inductance are equivalently [H] = [Wb A−1 ] = [VA−1 s−1 ] = [VCs−2 ] = [NC−1 mCs−2 ] =
elements in the wire. [Nms−2 ].

5-7 AJW, EEE3055F, UCT 2011 5-8 AJW, EEE3055F, UCT 2011
If we further assume that the flux Ψ linking each turn is the same, then “spiral staircase” winding around an imaginary centre line. The total flux
dΨ ΨM passing through S is proportional to the current I in the wire, and is
V1 = = V2 = ... = VN given by Z
dt
and ΨM = B · dS = LI
dΨ S
V = N V1 = N where L is the inductance of the multi-turn coil. Thus
dt
R
Because there are N turns, the flux Ψ theading the coil will be N times ΨM B(t) · dS
stronger than the contribution from a single turn, i.e. L= = S
I I(t)
Ψ = N × Ψ1turn where ΨM must be carefully evaluated for the particular coil structure. For
where Ψ1turn is the flux contribution from a single turn. Substituting, we a short, compact coil of N turns, it is not difficult to show that
obtain
L = N 2 L1turn
dΨ dΨ d(L1turnI) dI
V = N V1 = N = N 2 1turn = N 2 = N 2 L1turn where L1turn is the inductance of a single turn coil of the same radius. To
dt dt dt dt
Thus the inductance for a tightly wound N -turn coil is see this, one must grasp two points:

L = N 2 L1turn • 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 ap-
where L1turn is the inductance of a single turn. proximately N times larger than for a single turn)

It is worth remembering that the inductance increases as a function of N 2 . • the flux density on each component “disc” is N times stronger than the
If one doubles the number of turns, the inductances increases by a factor of flux density generated by a single turn (superposition of contributions
four. from N turns, each carrying current I)
Thus the total flux threading
Z the surface of the multi-turn contour is
Alternative Explanation 
ΨM = B · dS ≈ N × N × Ψ1turn
Analysis of a multi-turn coil is similar to the case of a single turn coil, with S

the added complication that the integration contour C is not a circle, but where Ψ1turn is equal to the flux generated by a single turn coil carrying
rather made up of a spiral that follows the wire (and a short section in the current I.
air gap between its terminals). As argued for a single turn case, the voltage The inductance of the multi-turn coil is then
across the terminals is ΨM N 2 Ψ1turn
Z Z L= = = N 2 L1turn
d I I
V =− E · dl = B · dS
(air) dt S
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 for the case where a << r.
The “internal” inductance of a long straight wire (in henrys per metre of
In order to calculate the inductance accurately, we need to consider both the wire), assuming uniform current density in the wire4, can be shown to be
field inside, and outside of the wire making up the coil. The total inductance independent of wire radius a, and for non-magnetic metal wire is given by
is usually computed in two parts: (ref Ramo et al.)
µ0
L = Linternal + Lexternal L′int = = 0.5 × 10−7 Hm−1

where Linternal is the contribution arising from the magnetic field within the Example Calculation
wire, and Lexternal is the contribution from the field outside the wire.
Calculate the inductance of a circular copper wire ring, radius r = 10 cm,
5.4.1 Inductance of a Circular Wire Ring wire radius a = 0.5 mm. NOTE µ ≈ µ0 in copper.
The external component is
External field    
8r −7 8 × 0.1
in the surrounding Lext ≈ µ0 r[ln − 2] = 4π × 10 × 0.1[ln − 2] = 0.7 × 10−6 H
wire
air. a 0.0005
diameter wire 2a
r
2a The internal inductance in henrys per metre is
radius µ0 4π × 10−7
of circle L′int = = = 0.05 × 10−6 Hm−1
8π 8π
Internal field
The total internal inductance
inside the metal

Figure 5.6: Circular wire ring, and its cross section. Lint ≈ L′int2πr = 0.050 × 10−6 ∗ 0.63 = 0.03 × 10−6 H

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


The inductance of a wire ring can be found from
R L = Lint + Lext = 0.73 × 10−6 H
B · dS ΨM
L= S =
I I
where B can be found by integrating the field contributions (using the Biot-
Savart 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 ΨM ≈ Iµr[ln 8r
a − 2] and hence the “external”
inductance (Ramo et al.) is 4
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
ΨM 8r towards the outside of the wire. This effect is known as the ’skin effect’, and is discussed later in the
Lext = ≈ µ0 r[ln − 2]
I a course.

5-11 AJW, EEE3055F, UCT 2011 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

Integration contour

length l I
B field

r
Radius
Figure 5.7: Short inductor.
B ~ 0 outside
I
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.
 
2 8r
L ≈ N µ0 r[ln − 2]
a
Figure 5.8: Long inductor.
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 For a long coil of length l, containing N turns carrying current I, the coil
increased by winding the coil on an iron or ferrite rod or on a toroid, which can be treated as a wrapped “current sheet” (row of dots in the illustra-
has a relative permeability of hundreds or thousands that of air.
Htion), fromRwhich H can easily be obtained by application of Ampere’s 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
H the horizontal segment of the integral
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 Ampere’s law implies H l ≈ N I and from which


H ≈ N I/l
and
B ≈ µ0 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 Consider two coils in close proximity, one containing N1 turns, and the
position of the the horizontal segment of the contour within the coil. This other containing N2 turns. Although not shown in the sketch, these coils
implies that the field intensity is uniform in a cross section of a long coil, are parts of circuits and carry currents.
and hence the flux through the cross section of coil is Ψ = Bπr2 = µ0 Hπr2 Let Ψ1 (t) be the component of the flux threading coil 1, resulting purely
(i.e. linking one turn of the coil) from the current I1(t) flowing in coil 1. Let the Ψ1 f rom 2 be the flux threading
coil 1 arising from the current I2(t) in coil 2. From Faraday’s law, the voltage
The inductance is the ratio of the total flux linking the coil to the current, across the terminals of coil 1 is
being d (N1Ψ1 + N1Ψ1 f rom 2 ) dΨ1 dΨ1 f rom 2
R V1 (t) = = N1 + N1
B · dS N Ψ N Bπr2 N (µ0 N I/l)πr2 πµ0 r2 N 2 dt dt dt
L= S = = ≈ =
I I I I l where N1Ψ1 is the total flux threading the multi-turn coiled surface. Since
Ψ1 is proportional to I1 and Ψ1 f rom 2 is proportional to I2, we have,
5.4.4 Inductance of an Intermediate Length Multi-turn Coil
dI1 dI2
In cases where the coil can neither be considered very long or very short, V1 (t) = L1 + M12
the following approximate formula is commonly used: dt dt
πµ0 r2N 2 where L1 is the self inductance constant of coil 1, and M12 is another con-
dΨ f rom 2
L≈ stant. Since N1 1 dt = M12 dIdt2 ,
l + 0.9r
The formula incorporates an empirical correction factor (+0.9r) in the de- N1Ψ1 f rom 2
M12 =
nominator. I2

5.5 Mutual Inductance


Similarly, the voltage across the terminals of coil 2 is
If two wire coils are close to one another, flux resulting from current flowing
d (N2Ψ2 + N2Ψ2 f rom 1 ) dΨ2 dΨ2 f rom 1
in one coil will thread the coil of the other. Thus a changing current in one V2 (t) = = N2 + N2
coil, will result in a changing flux in the other and hence induce a voltage dt dt dt
across its terminals. This is the basis of a transformer. and
dI2 dI1
ψ2 + ψ2f rom1 V2 (t) = L2+ M21
ψ1 + ψ1f rom2 dt dt
where L2 is the self inductance constant of coil 2, and M21 is a constant:
N2Ψ2 f rom 1
V1 (t) V2 (t) M21 =
I1
I1 (t) I2 (t)
It can be shown (consult more detailed texts), that regardless of the ge-
ometry, 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” 5.6 Kirchhoff’s Voltage Law
between the coils5.
The relationship between Kirchhoff’s law for a lumped element circuit model
Mutual Inductance for Tightly Coupled Coils and the physical component layout, is established by application of Fara-
day’s law.
A special case is when the coils are tightly coupled, e.g. stacked on top Consider applying Faraday’s law to the closed contour indicated by the
of one another such that Ψ1 f rom 2 = Ψ2 and Ψ2 f rom 1 = Ψ1 (or coils would dotted line in the following physical circuit representation. Define dS point-
around a common torroid). For this case, ing into the page, which implies the integration direction is clockwise.
 
dI2 dΨ1 f rom 2 dΨ2 L2 dI2
M12 = N1 = N1 = N1
dt dt dt N2 dt 1 R
2
N1 N2
and hence M12 = N2 L2 . Similarly, M21 = N1 L1 .
I(t)
V21 L
1 2
Since M = M12 = M21 , the product M12M21 yields V32
V(t) V10
N1 N2 3
M2 = L2 L1 = L1 L2 V03
N2 N1 0 Coil
or p
M= L1 L2 0 3
for tightly coupled coils. C
The voltage ratio is
p Figure 5.10: Series circuit loop - Faraday’s law is applied along the dotted line to derive
V1 L1 dIdt1 + M dIdt2 N12 L1turn dIdt1 + N12 L1turnN22 L1turn dIdt2 N1 Kirchhoff’s voltage law.
= dI2 dI
= dI
p dI
=
V2 L2 dt + M dt1 N22 L1turn dt2 + N12 L1turnN22 L1turn dt1 N2
Since E → 0 in the wires, the voltage drops around the circuit occur
which is a well known result for tightly coupled coils. across the components. Thus, we can write
I Z 1 Z 2 Z 3 Z 0 Z
∂B dΨM
− E·dl = − E·dl− E·dl− E·dl− E·dl = ·dS =
0 1 2 3 S ∂t dt
or
dΨM
V10 + V21 + V32 + V03 =
dt
The flux threading the loop can be split into three contributions:
5 N Ψ
One could in principle, calculate M12 = 1 1If2rom 2 for a given coil geometry by doing a surface inte-
R ΨM = Ψapplied + Ψself + Ψmutual
gration of B(i2 ) · dS to obtain Ψ1 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). where

5-17 AJW, EEE3055F, UCT 2011 5-18 AJW, EEE3055F, UCT 2011
• Ψapplied refers to any flux imposed on the circuit e.g. wave a bar magnet
I
past the circuit. R
• Ψself 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). V (t) L
Ψself = Lself I where Lself is the self inductance of the loop, which
carries current I. dS points into the page
(the direction of positive flux)
• Ψmutual refers any leakage flux from other parts of the circuit (notably −
∂ψapplied
∂t
the inductive element) that threads the loop.
Substituting the lumped element relationships derived above,
Z
1 t dI dΨM Lself
V (t) − IR − I(t)dt − L =
C t0 dt dt
P
• Kirchhoff’s 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 dΨdtM .
• It is noted that the term dΨdtM will modify the current flowing in the C
circuit, and should be included for accurate prediction of the behaviour
of the circuit. Figure 5.11: Circuit modified to incorporate an additional series inductor Lself which mod-
els the series inductance of the loop, and an additional voltage source which
• In practice however, this term is usually small compared to the other models unwanted external signals.
terms, and is often neglected in practical circuit design. dΨ
The term dtself = Lself dI
dt resulting from the current in the loop, is mod-
For the complete model shown below, the equation in the form of Kirchhoff’s elled by a (small) series inductance Lself . A feeling for the magnitude of
law is written as this self inductance can be gained by considering a circuit arranged in a
Z circular loop of radius 10 cm. We previously calculated the self inductance
1 t dI dΨapplied dΨself dΨmutual
V (t) − IR − I(t)dt − L − − − =0 of a wire ring of radius 10 cm and wire radius of 0.5 mm to be 0.7 µH.
C t0 dt dt dt dt
At an operating frequency of say 1 kHz, the AC reactance of this term
or is XL = 2πf Lself ≈ 4 × 10−3 ohms, which is usually small enough to be
N
X neglected from calculations. At higher frequencies, this term may become
Vi = 0
i=1
significant.

The term applied
dt arises if the circuit is exposed to some externally gen-
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

5-19 AJW, EEE3055F, UCT 2011 5-20 AJW, EEE3055F, UCT 2011
course radio waves are ever present, but their contribution is usually in- 5.7 Kirchhoff’s Current Law at a Node
significant compared to the voltage levels in the circuit. Circuits can be
shielded from external sources by placing them in a metal enclosure known Consider the illustration in Figure 5.12 showing four wires connecting to a
as a Faraday cage 6. node carrying currents I1, I2, I3 and I4. Kirchhoff’s node current law states
The term dΨmutual in this context arises from the leakage flux from the that the sum of all currents leaving the node equals zero, i.e.
dt
inductor L, and is in practice usually small compared to the voltage drop N
X
across the (multi-turn) inductor. The net effect may either be to increase Ii = I1 + I2 + I3 + I4 = 0
or decrease the current in the circuit, depending on the physical orientation i=1
of the inductor. R
NOTE: The flux S B · dS requires the direction of dS to be defined. Since
I2
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.

I1

I3

b
n
closed surface S1

dS

I4

closed surface S2

6
A Faraday cage will provide good shielding from DC electric fields. DC magnetic fields however do Displacement current "flows" between plates
penetrate metal enclosures. e.g. the earth’s 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 Figure 5.12: The relationship between Kirchhoffs cuurent law at a node the continuity
studied later in this course. For good shielding at a particular frequency, the enclosure wall should be equation.
considerably thicker that the skin depth (at that frequency).

5-21 AJW, EEE3055F, UCT 2011 5-22 AJW, EEE3055F, UCT 2011
Consider now the continuity of charge relationship For example, consider the closed surface S = S1 surrounding the node in
I Z Figure 5.12. There are N = 4 wires piercing the surface and joining at the
d dQ
J · dS = − ρdV = − node.
S dt V dt
For S1 , the total conduction current leaving the surface is
which states that the total conduction current leaving an arbitrary closed I
surface S is equal to (minus) the rate of change of charge within the volume Ic = J · dS = I1 + I2 + I3 + I4
S1
V enclosed by S. One can re-express the continuity relationship in a form
that looks similar to Kirchhoff’s law by moving the charge term to the left The displacement current is typically insignificant (there is no significant
hand side: I Z charge build up within S1 , i.e.
d I
J · dS + ρdV = 0 ∂D dQ
dt V Id = · dS = ≈0
S1 ∂t dt
S
The 2nd termR can beHexpressed as a surface integral over S by substituting Thus we have
Gauss’ law ρdV = S D · dS,
I1 + I2 + I3 + I4 ≈ 0
I I
d If one shrinks surface S1 to a tiny surface surrounding the node, the dis-
J · dS + D · dS = 0
S dt S placement current shrinks to zero and the relationship converges exactly to
Moving the time derivative within the integral, the continuity equation be- Kirchhoff’s law.
comes I I
∂D If however, we choose a surface S = S2 in such a way as to pass between the
J · dS + · dS = 0
S S ∂t plates of the capacitor as illustrated in Figure 5.12, then we have a slightly
which says that the sum of the conduction current Ic and the displacement more subtle situation.
current Id leaving an arbitrary closed surface7 is zero. I.e. for any closed As there is one less wire cutting the surface, the total conduction current is
I
surface S,
Ic + Id = 0 Ic = J · dS = I1 + I2 + I3
S2
where I There is however a significant charge build-up on the plate(s) of the capac-
Ic = J · dS itor as a result of current I4. The charge Qplate on the plate (within S2 )
S dQ
builds up at a rate of dtplate = I4.
and I
∂D
· dS
Id = Thus the continuity relationship
S ∂t I
Thus we have derived a generalised form of Kirchhoff’s current law, which dQ
J · dS = −
can be applied to an arbitrary closed surface. S dt
applied to surface S2 becomes
7
It is worth noting that the total displacement current flowing
H out of a closed surface is equal to the time
dQplate
rate of change of charge enclosed by the surface, i.e. Id = S ∂D I1 + I2 + I3 ≈ − = −I4
∂t · dS = dQ/dt. dt

5-23 AJW, EEE3055F, UCT 2011 5-24 AJW, EEE3055F, UCT 2011
Rearranging, we get 5.8 The Relaxation Time of Conducting Materials
I1 + I2 + I3 + I4 ≈ 0
The term ‘conductor’ refers to a material that will carry current when sub-
which is consistent with the case where S = S1 and Kirchhoff’s law.
jected to an electric field. In solid materials, like metals, electrons are free
The approximation (” ≈ ”) is present in the above expression because a
to move, and the net movement of electrons constitutes a current. In liquids
small (and negligible) charge will exist on the surface8 of the conductors 1
(e.g. a salt solution), charged ions in solution are free to move allowing a
to 4
current to exist. Insulating materials, in contrast, are materials for which
the electrons are tightly bound to particular atoms, and hence no current
Another way to look at the situation is to observe that the sum of all
can flow.
currents, both conduction and displacement current, flowing out of a closed
A “perfect conductor” is one for which there is an unlimited abundance
surface is zero, i.e. for surface S4,
of free electrons. The conductivity of a perfect conductor is infinite - an
I1 + I2 + I3 + Id = 0 infinitesimal electric field will create a large current. Metals can often be
H approximated as perfect conductors in the analysis of their behaviour under
where Id = S ∂D ∂t · dS is the displacement current leaving the surface. Id certain conditions.
is concentrated primarily between the plates of the capacitor (where the If a conducting object is placed in a stationary position within an electric
electric field is strongest). field, the electrons will, given time, rearrange themselves such that:
H • E goes to zero inside the conductor (electrons quickly re-arrange them-
Also, since we have shown that dQdt
= S ∂D
∂t
·dS = Id (for any closed surface),
dQ selves until the total E = 0 inside conductor). Note: in the steady state
and if the only significant dt within S4 is the charge build up on the inner
situation, the net force on the electrons must go to zero - the electrons
plate of the capacitor due to I4, then
will rearrange themselves to achieve this. Since the conductor is not
dQ dQplate moving (stationary), in the steady state situation, the magnetic force
Id = ≈ = I4
dt dt on the electrons will be zero, and hence the electric force must also be
which again for S4 implies zero9.
• The charge density ρ = 0 inside the conductor (since div D = ρ and
I1 + I2 + I3 + I4 ≈ 0
E = 0, it means ρ = 0) .
• Any net charge (excess charge) resides on the surface in an infinitesi-
mally thin layer (we refer to this charge as a ‘surface charge’).
9
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. Φb − Φa = a E · dl = a (v × B) · dl. If v is perpendicular to B and the rod
8
Any excess charge must be the surface, because ρ → 0 very rapidly inside a metal conductor - see section is orientated such that its length is perpendicular to v and B, then the potential difference between
on relaxation time. the ends will be vBl.

5-25 AJW, EEE3055F, UCT 2011 5-26 AJW, EEE3055F, UCT 2011
• The potential Φ(x, y, z) is constant throughout the conductor (since To see how quickly this happens in practice, the time constant may be
the electric field is zero inside). calculated for various materials. For example for a metal conductor like
copper (σ = 5.8 × 107 Sm−1, ε ≈ ε0 = 8.85 × 10−12 Fm−1), the time
• E is perpendicular at the boundary (i.e. no tangential component).
constant is τ = σε = 1.5 × 10−19 s, which is extremely short compared to
In practice, one might wonder just how long it would take for the electrons say the period of a 100 GHz microwave sinusoid, being 10−11 seconds. In
to re-arrange themselves. Imagine setting up an arrangement of charge electronic circuits, the charge in the wires rearranges itself very quickly in
ρ(x, y, z) within a homogeneous conducting material and then releasing the response to the dynamics of the circuits (i.e. to a very good approximation,
charge at some instant. The charge will redistribute itself such that the we consider E ≈ 0 and ρ ≈ 0 inside the connecting copper wires - a small
electric field goes to zero at every point within the conductor10. This hap- component of E must however exist to drive the current).
pens very rapidly in metals, so fast that it can be considered instantaneous For a weakly conducting liquid like tap water (σ ≈ 10−2 Sm−1, ε ≈ 81ε0
in many practical situations. Fm−1), the relaxation time is about 70 × 10−9 s. For a good insulator like
Consider the charge within the conducting body. The movement of charge glass (e.g. σ = 10−14 Sm−1, ε = 5ε0), the relaxation time is calculated to
will be governed by the continuity equation, for which the differential form be about 4000 seconds (67 minutes).
Exercise: Calculate the relaxation time of iron (σ = 0.9 × 107 Sm−1).
∂ρ
∇·J=−
∂t
describes the relationship between current leaving a small volume element, 5.9 Shielding and The Faraday Cage
and the rate of change of charge within the element. If we substitute
Circuitry may be shielded from external electric fields by enclosing the cir-
J = σE, we get
∂ρ cuit inside a metal box known as a Faraday cage. External electric fields
∇ · (σE) = − have no influence on the circuitry within a box made from a perfect con-
∂t
ductor - an electrically quiet zone exists within the box. On a larger scale,
and then eliminate E via Gauss’ law (∇ · E = ρ/ε) we obtain a first order
Faraday cages can provide protection against lightning strikes. “Low fre-
differential equation
σ ∂ρ quency” magnetic fields can however penetrate a real metal enclosure and
ρ+ =0 influence the circuitry inside it. Try for yourself to see if a permanent mag-
ε ∂t
which has solution net is able to attract iron pieces through the walls of a metal box. DC
σ
ρ(t) = ρ0 e− ε t magnetic fields penetrate through metal enclosures. For example, a mag-
netic compass will still detect the earth’s magnetic field inside a Faraday
where ρ0 is the initial charge density at time t = 0.
cage. Effective shielding from a magnetic field at 50 Hz requires a thick
Thus the charge density at any point within the material will dissipate
wall (several mm), preferably made from a high permeability material. As
to zero with an exponential decay. The decay curve is characterised by the
the frequency increases, a metal wall becomes more effective in attenuating
time constant τ = σε , also known as the relaxation time, which is the time
magnetic fields. In the MHz range and higher, metal enclosures are very
at which the charge density has reduced to e−1 ≡ 36.8% of its initial value.
effective (if well sealed) for shielding circuits from external electromagnetic
After 5τ the charge density will have decayed to less than 1% of the initial
fields, and also for preventing radiation leakage from the circuitry within
value.
the enclosure. Electromagnetic waves “reflect off” the enclosure, and what
10
Here we are ignoring the granularity of electrons and treating the charge as a kind of fluid.

5-27 AJW, EEE3055F, UCT 2011 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). V (t) dψm /dt

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. V (t)
Such parallel wire transmission lines are particularly susceptible to inductive
coupling of magnetic fields, especially when several signals need to be carried net dψm /dt ≈ 0
in the same bundle. Changing magnetic flux dΨ/dt within the circuit loop
induces an additional voltage which adds to the signal voltage in accordance Figure 5.13:
with Faraday’s 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 dΨ/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 cable’s 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.

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 Cat-5e
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 frequen-
cies 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-6
Cat-5e is 100 ohms.
Mechanical arrangement within the cable can further reduce coupling be-
tween pairs. For example the “Power Cat-6 four pair cable” sold by RS
Electronics contains four unshielded twisted pair (UTP) cables, with a cen-
tral separator, and is designed to support high speed data transmission
systems (frequencies up to 250 MHz).
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

You might also like