0% found this document useful (0 votes)
24 views

Ch2 Macroscopic Fields V8

This document discusses macroscopic electric and magnetic fields. It begins by introducing Maxwell's macroscopic equations, which define the electric displacement field D and magnetic intensity field H in terms of the electric field E, polarization P, magnetic B field, and magnetization M. These auxiliary fields allow Maxwell's equations to be formulated without needing to consider atomic-scale charges and currents. The document then reviews electrostatics concepts like Gauss' law and the effect of dielectrics on capacitance. It introduces polarization and susceptibility and defines the electric displacement field D in terms of E and P. The document continues to discuss the electric field in dielectric media and formulate Gauss' law using D.

Uploaded by

christian26brown
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
24 views

Ch2 Macroscopic Fields V8

This document discusses macroscopic electric and magnetic fields. It begins by introducing Maxwell's macroscopic equations, which define the electric displacement field D and magnetic intensity field H in terms of the electric field E, polarization P, magnetic B field, and magnetization M. These auxiliary fields allow Maxwell's equations to be formulated without needing to consider atomic-scale charges and currents. The document then reviews electrostatics concepts like Gauss' law and the effect of dielectrics on capacitance. It introduces polarization and susceptibility and defines the electric displacement field D in terms of E and P. The document continues to discuss the electric field in dielectric media and formulate Gauss' law using D.

Uploaded by

christian26brown
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

Contents

2 Macroscopic Fields 3

2.1 Electric Field: Revision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2.1.1 Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2.1.2 Dielectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2 Electric Field in Dielectric Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.2.1 External Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.2.2 Energy Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.3 Magnetic Field: Revision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3.1 Faraday’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.4 Magnetic Vector Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.5 Magnetic Intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.5.1 Magnetic Susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.6 Interfaces and Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.6.1 Normal Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.6.2 Tangential Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.7 Summary of Linear Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1
PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

PHAS0038: Term 1 Macroscopic Fields 2


Chapter 2

Macroscopic Fields
• Maxwell’s equations have two major variants: the microscopic set of these equations use total charge and total
current from all sources, including the difficult-to-calculate atomic-scale charges and currents in materials. The
macroscopic set help with this problem, by defining two new auxiliary fields that can ‘sidestep’ having to know
in detail these atomic-scale charges and currents.
• 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 in form to those that Maxwell himself introduced. The cost of this
approach is that additional fields need to be defined. These are: (i) the electric displacement field D, which is
defined in terms of the electric field E and the polarization P (electric dipole moment per unit volume) of the
material; and (ii) the magnetic intensity or magnetizing field H, which is defined in terms of the magnetic B
field and the magnetization M (magnetic dipole moment per unit volume) of the material. In this chapter, we
will look in more detail at these macroscopic fields, D and H.

2.1 Electric Field: Revision


We begin this section of the course by going over electrostatic concepts which should hopefully be familiar from
previous studies, including Gauss’ Law and the effect of dielectrics on capacitance.

2.1.1 Electrostatics

• Consider a single charge, q, at position r′ in space (the ‘source point’). At another position (‘test point’),
the electric field due to q is:
{\bf E}({\bf r})=q({\bf r}-{\bf r}^{\prime })/(4\pi \epsilon _0 |{\bf r}-{\bf r}^{\prime }|^3) (2.1)

• Taking a surface integral gives


H
S
E · ndS = q/ϵ0 , for any surface enclosing q.

• Increasing the number of charges, and using the principle of superposition, we obtain:
\oint _S {\bf E}\cdot {\bf n} dS = \sum _{i} q_i / \epsilon _0
(2.2)

• For a continuous medium on scales much larger than atomic, we can use the concept of charge density ρ
and the integral becomes: \oint _S {\bf E}\cdot {\bf n} \, dS = \int _{\Vol } (\rho ({\bm {r^{\prime }}}) / \epsilon _0) \, d\Vol ^{\prime }
(2.3)

• This leads directly to a differential form of Gauss’ law for continuous media:

{{\bf \nabla \cdot E} = \rho /\epsilon _0} (2.4)

3
PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.1.2 Dielectrics

A dielectric is an electrical insulator that can be polarized by an applied (external) electric field. When a
dielectric is placed in an electric field, electric charges do not flow through the material, as in a conductor, but
only slightly shift from their average equilibrium positions, causing dielectric polarization. Positive charges are
displaced along the field direction and negative charges shift in the opposite direction. This creates an internal
electric field, which reduces the total field within the dielectric itself.

• Recall that capacitance is defined by Q = C ∆V where Q is total charge magnitude on a capacitor plate,
∆V is potential difference between plates, and C is the capacitance (‘charge-storing capability’).

• Capacitance changes when a dielectric material, characterised by a ‘dielectric constant’ κ, is added between
the plates:
C_{dielectric}=\kappa \, C_{vacuum} (2.5)

• Recall that a dielectric (insulator) has no free (mobile) charges.

• The polarization is defined as P = ϵ0 χe E. It is the electric dipole moment per unit volume.

• This introduces the quantity known as the susceptibility, χe .

• The dielectric constant is given by κ = 1 + χe . This is also referred to as the relative permittivity, indicated
by ϵr

• The absolute permittivity of the material is thus the product of the relative permittivity with the permittivity
of the vacuum, ϵ = ϵr ϵ0

Polarization reflects the fact that the atoms which make up the dielectric consist of separate positive (nucleus)
and negative (electrons) charges. These respond differently to the electric field, leading to a shift in the overall
charge distribution of the dielectric, while keeping it neutral on large scales. We will consider the microscopic
origin of polarization in detail further on.

Figure 2.1: Electronic polarization occurs due to displacement of the centre of the negatively charged electron
cloud relative to the positive nucleus of an atom by the imposed electric field.

PHAS0038: Term 1 Macroscopic Fields 4


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.2 Electric Field in Dielectric Media


• We want to develop a theory for electric fields in the presence of polarized media.
• We will start by considering the field outside a piece of polarized dielectric.
• This will introduce the idea of polarization charge density.
• Then we will consider the field inside a piece of polarized dielectric.
• We will thus find a useful reformulation of Gauss’ Law.

We start by finding the potential at a point r due to a small volume of polarized material at a point r′ . We will
then integrate this over the entire piece of dielectric material. First, note that the potential at r due to a single
dipole (of moment p) at r′ is:

\phi ({\bf r}) = \frac {1}{4\pi \epsilon _0}\frac {{\bf p}\cdot ({\bf r}-{\bf r}^{\prime })}{|{\bf r}-{\bf r}^{\prime }|^3} (2.6)

Recall that p = qd where q is magnitude of charge and and d is separation (pointing from excess negative charge
−q to excess positive charge +q within the dipole). For a quasi-continuous medium with many such dipoles,
we may write the total electric dipole moment of a volume element δV ′ as P′ δV ′ (moment per unit volume
multiplied by element volume). Hence we have:

\delta \phi ({\bf r}) = \frac {{\bf \delta }\Vol ^{\prime } \, {\bf P}({\bf r}^{\prime })\cdot ({\bf r}-{\bf r}^{\prime })} {4\pi \epsilon _0|{\bf r}-{\bf r}^{\prime }|^3} (2.7)

When we take the limit δV ′ → 0 and sum over the elements, we get an integral expression for the total potential
due to the polarized material:

\phi ({\bf r}) = \int _{\Vol '} \frac {d\Vol ^{\prime }\, {\bf P}({\bf r}^{\prime })\cdot ({\bf r}-{\bf r}^{\prime })} {4\pi \epsilon _0|{\bf r}-{\bf r}^{\prime }|^3} (2.8)

We use the following result involving the gradient of the function 1/|r−r′ |, with respect to the variable coordinates
of the source point r ′ (worth remembering):

{\bf \nabla }^{\prime } \biggl ( \frac {1}{|{\bf r}-{\bf r}^{\prime }|} \biggr ) = \frac {({\bf r}-{\bf r}^{\prime })} {|{\bf r}-{\bf r}^{\prime }|^3} (2.9)

to transform our integral into:

\phi ({\bf r}) = \frac {1}{4\pi \epsilon _0} \int _{\Vol '} {\bf P}({\bf r}^{\prime }) \cdot {\bf \nabla }^{\prime }\biggl ( \frac {1}{|{\bf r}-{\bf r}^{\prime }|}\biggr )d\Vol ^{\prime } (2.10)

Using the formula for the divergence of the product of a scalar and vector function from our Mathematical
Identities, we see that
{\bf P}({\bf r}^{\prime })\cdot \nabla ^{\prime } \xi = \nabla ^{\prime }\cdot (\xi \, {\bf P}({\bf r}^{\prime })) - \xi \, \nabla ^{\prime }\cdot {\bf P}({\bf r}^{\prime }), \notag
for ξ = 1/|r − r′ |. Our integral thus becomes:

\phi ({\bf r}) = \frac {1}{4\pi \epsilon _0} \int _{\Vol '} \biggl [ {\bf \nabla }^{\prime } \cdot \biggl ( \frac { {\bf P}({\bf r}^{\prime }}{|{\bf r}-{\bf r}^{\prime }|}\biggr ) -\frac {1}{|{\bf r}-{\bf r}^{\prime }|}{\bf \nabla ^{\prime } \cdot P}({\bf r}^{\prime })\biggr ]\,d\Vol ^{\prime } (2.11)

R H
Finally, we use the divergence theorem on the first term [ V
∇ · F dV = S
F · n dS], to give the potential outside
a polarized dielectric object:

\phi ({\bf r}) = \frac {1}{4\pi \epsilon _0} \oint _{S'} \frac { {\bf P}({\bf r}^{\prime })\cdot {\bf n}}{|{\bf r}-{\bf r}^{\prime }|}\,dS^{\prime } + \frac {1}{4\pi \epsilon _0} \int _{\Vol '} \frac {-{\bf \nabla \cdot P}({\bf r}^{\prime })}{|{\bf r}-{\bf r}^{\prime }|}\,d\Vol ^{\prime } (2.12)

PHAS0038: Term 1 Macroscopic Fields 5


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

• The surface density of polarization charge is thus defined as follows, at the discontinuous physical boundary
for the function P, between the inside (dielectric) and outside (empty space) of the material (see Figure
2.2):
\textcolor {red}{\sigma _P = {\bf P \cdot n}} (2.13)

• The volume polarization charge density is also defined. by the second integral above:

\textcolor {red}{\rho _P = -{\bf \nabla \cdot P}} (2.14)

• We can write the potential as the integrated surface charge density (charge per unit area) over its thin
physical surface, and the integrated volume charge density over its interior:

\phi ({\bf r}) &= \frac {1}{4\pi \epsilon _0} \biggl ( \oint _S \frac {\sigma _P}{|{\bf r}-{\bf r}^{\prime }|}dS^{\prime } + \int _\Vol \frac {\rho _P}{|{\bf r}-{\bf r}^{\prime }|}d\Vol ^{\prime }\biggr ) (2.15a)

For a uniformly polarized medium, ∇ · P = 0, so there is no net bound charge within the material on large scales,
but there will be bound charge on the physical surface (discontinuity in P) (see also Figure 2.2). The different
types of charge in a more general situation would be:
Bound charge: The charge within a material that is unable to move freely through the material. Small
displacements of bound charge are responsible for polarization of a material by an electric field.
Free charge: The charge in a conducting material associated with the mobile conduction electrons that are free
to move throughout the material. These electrons can also carry electric current.

Figure 2.2: Origin of surface charge density due to polarization.

We have considered the field due to a polarized dielectric, but only outside the dielectric. What is the field inside
a polarized dielectric?

• Consider three (small) charged conductors embedded in a volume of dielectric material (see Figure 2.3).

• They have charges q1 , q2 and q3 (sum to Q).

• Now use Gauss’ Law, noting that there is free charge Q and bound charge QP , and use an imaginary
surface S inside the dielectric material, which still encloses the conductors. Note that there are also three
‘contact’ surfaces S1 , S2 and S3 between the dielectric and the conductors. Finally, we denote by V the
volume which is filled with only dielectric material, i.e. it fills the ‘space’ between S, S1 , S2 and S3 .

• Gauss’ Law, applied to just the surface S, links the surface integral of E and the total amount of charge
(both free and bound) inside S:

PHAS0038: Term 1 Macroscopic Fields 6


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

Figure 2.3: Sketch of three conductors embedded in a dielectric.

\label {eq:216} \oint _{S} {\bf E \cdot n}\,dS = \frac {1}{\epsilon _0}(Q+Q_P)
(2.16)

We start by noting that, for the bound charge part, we can represent QP as an integral sum involving surface
charge density at the contact surfaces, and volume charge density throughout V:

Q_P = \int _{S_1, S_2, S_3}{\bf P \cdot n}\,dS + \int _{\Vol } - {\bf \nabla \cdot P} \, d\Vol (2.17)

It is important to realize that the arbitrary external bounding surface S we set does not enter into this integral
because it lies inside the continuous dielectric medium, and so there is no polarization charge surface density on
it (it is not a real, physical
R surface i.e.Hnot a boundary across which P changes discontinuously). We now use
the divergence theorem [ V ∇ · F dV = S F · n dS] to transform the second integral into a surface integral. But
we must take care: this time, we must include the surface S because it does, mathematically speaking, form one
of the boundaries of the volume V (the other boundaries being S1 , S2 , S3 ). It is also important to understand
the directions of the surface normals. They point outwards from the volume V. Explicitly, this gives:

Q_P &= \int _{S_1+S_2+S_3}{\bf P \cdot n}dS - \oint _S {\bf P \cdot n}dS - \int _{S_1+S_2+S_3}{\bf P \cdot n}dS \\ &= - \oint _S {\bf P \cdot n}dS

(2.18b)

Now we can use this expression for QP in Gauss’ law inside the dielectric, which was given as Eq. 2.16:
\oint _S {\bf E \cdot n}dS = \frac {1}{\epsilon _0}Q - \frac {1}{\epsilon _0} \oint _S {\bf P \cdot n}dS
(2.19)

After a little manipulation, we can rewrite this in terms of the free or external charge, Q.

• Using the divergence theorem yet again, we find that:

Q &= \oint _S (\epsilon _0 {\bf E}+{\bf P}){\bf \cdot n}\,dS, (2.20a)

becomes in the more general case of a distributed density of free charge, ρf :


\int _{V} \rho _f ({\bf r})\,dV = \int {\bf \nabla \cdot D} \, dV,
(2.21)

where we have defined the vector field D = ϵ0 E+P, and ρf (r) is now the volume density of free charge.
We have also used a slightly different symbol for volume, V , to emphasize that the volume integral applies
to the entire volume inside the exterior surface S.

PHAS0038: Term 1 Macroscopic Fields 7


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

• The electric displacement D is therefore the vector field whose divergence is only the free (or external)
charge density.

• So, if we consider a free charge density function, ρf (r), and use the divergence theorem, we obtain:

Divergence of the Electric Displacement, D

\bm {\nabla } \cdot \bm {D} = \rho _f (\bm {r}) (2.22)

2.2.1 External Charge

• We have talked about free charge (as opposed to the bound charge).

• With a dielectric, the difference is clear.

• Charge added from outside (external charge) is distinct from polarization charge, so it also contributes to
the divergence of D.

• But added charge is generally not free to move in a dielectric.

• Within a conductor, on the other hand, charge is free to move around.

• It is thus important to be aware of the difference between charge added and charge already present.

• In general, the polarization P is a function of the material and the field E.

• We write P = ϵ0 χe E for linear, isotropic, homogeneous media.

• In these media, as χe (the electric susceptibility) is constant (homogeneity):

{\bf D} = \epsilon _0{\bf E} + \epsilon _0 \chi _e {\bf E} = \epsilon {\bf E} (2.23)

• We call ϵ = ϵ0 (1 + χe ) the permittivity, and ϵr = ϵ/ϵ0 the relative permittivity or dielectric constant.

• Linear: P depends linearly on E

• Homogeneous: χe does not vary with position

• Isotropic: P and E are parallel

It is important to realize that a sufficiently strong electric field can break apart the charges in a material which
form the microscopic dipoles. At this point, called dielectric breakdown, all approximations discussed to this
point are invalid. For air, whose dielectric constant is 1.0006, the maximum field sustainable without breakdown
is around 3 × 106 V /m.
The reason that we refer to an isotropic dielectric for the relation P = ϵ0 χe (E) E is that it implies that the
polarization has the same direction as the external field. This is a good approximation for most media, but it
is necessary in some media to replace this with a tensor relationship, where the two vectors are not in the same
direction.
Under the action of extreme electric fields, the polarizability of a material can result in a non-linear dependence
in which case the material is referred to as non-linear and P is calculated in most cases as a Taylor expansion of
the generating field E. The coefficients of the expansion are the non-linear susceptibilities.

PHAS0038: Term 1 Macroscopic Fields 8


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.2.2 Energy Density

• What is the energy density of an electric field?


• We will answer this question using two approaches:
– Charge flowing into a capacitor;
– Adding a small charge to an existing electric field distribution.
• We will find that the final result in both cases is the same, i.e.

Energy density of an Electric Field


u = \frac {1}{2} \bm {D} \cdot \bm {E} (2.24)

Considering a capacitor first, we assume that it is in the process of being charged. If we start with the expression
for power (which is rate of change of energy with time) for a current I(t) flowing into a capacitor plate at voltage
V (t) for time t, then P (t) = V (t)I(t).
The energy at a given time is:

W = \int P(t)\,dt = \int V(t) \, I(t)\,dt = \int \frac {Q(t)}{C}\frac {dQ}{dt}dt = \frac {1}{2}\frac {Q^2}{C}, (2.25)

where Q denotes magnitude of charge on the plate and C denotes capacitance. For example, a parallel plate
capacitor with plates of area A separated by a distance d, the capacitance is given by C = ϵA d (for material of
permittivity ϵ). Using this expression and V = Q/C, we find that the electric field between the plates can be
written:
E=\frac {V}{d}=\frac {Q}{Cd}=\frac {Qd}{\epsilon Ad}=\frac {Q}{\epsilon A} (2.26)
Q
Of course, as D = ϵE, we find D = A. So the energy density is:

u &= \frac {W}{Ad} = \frac {1}{2}\frac {Q^2}{CAd} \\ &= \frac {1}{2}\frac {Q^2}{\epsilon A^2} \\ &= \frac {1}{2}{\bf D \cdot E}

(2.27c)

Another (more general) way to derive the same formula is to consider the work done in bringing a charge from
infinity to the point where the energy density is being determined. We know that the energy of a point charge,
q, in a potential ϕ is given by W = qϕ. This can be generalized for an entire charge distribution specified by the
charge density function ρ(r).
Now, what would be change in electrostatic energy when adding a very small amount of charge density, δρ, to
an existing configuration of charge density which has electrostatic potential ϕ? We can use our recent result for
Gauss’ theorem, ∇ · D = ρ, to work out the incremental change (to first order) in the system energy W :

\delta W &= \int _\Vol \delta \rho \, \varphi \, d\Vol \\ \mathrm {Since~~~} \delta \rho &= {\bf \nabla \cdot } \delta {\bf D} \\ \mathrm {and~~~} \varphi ({\bf \nabla \cdot \delta D}) &= {\bf \nabla \cdot } (\varphi \, {\bf \delta D})-{\bf \delta D \cdot }({\bf \nabla }\varphi ) \\ \delta W &= \int _\Vol ({\bf \nabla \cdot } (\varphi \, \delta {\bf D})-\delta {\bf D \cdot (\nabla }\varphi ) ) \, d\Vol \\ \delta W &= \int _S \varphi \, \delta {\bf D \cdot n}dS -\int _\Vol \delta {\bf D} \cdot (\nabla \varphi ) \, d\Vol

(2.28e)

PHAS0038: Term 1 Macroscopic Fields 9


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

where we have used the divergence theorem on the first part integral in the final line. But we know that
E = −∇φ adopting the usual sign convention, and we can also notice that the first integral will fall off rapidly
as we expand the surface further and further out (D ∝ 1/r2 , ϕ ∝ 1/r). This means that we can write, as our
volume of integration tends to infinity (using E = −∇ϕ):

\delta W = \int _\Vol \delta {\bf D \cdot E}\,d\Vol (2.29)

Now, if we assume a linear, dielectric medium, we know that D = ϵE, and we can integrate over the entire field
as it ‘builds up’ from zero to D, as the charges are brought together to form the final distribution:

W = \int _0 ^D \delta W =\int _0^D \int _\Vol \delta {\bf D \cdot E}\,d\Vol (2.30)

1
Now, since δD · E = δ(ϵE) E = ϵ × 2 δ(E 2 ), we can write:

W = \frac {1}{2} \int _0 ^E \int _\Vol \epsilon \, \delta (E^2)\,d\Vol =\frac {1}{2} \int _\Vol \epsilon E^2 \,d\Vol (2.31)

This of course gives us the result we derived above, namely u = E · D/2.

2.3 Magnetic Field: Revision


An important point to note as we start the area of magnetic fields is that this is where the essential link between
electric fields and magnetic fields (leading to the unified area of electromagnetism) becomes apparent. Thus far
we have considered electrostatics only.

• The magnetic field at test point r2 due to a circuit described by source points r1 , in both differential (single
current element) and integral (entire circuit) forms, is given by:

Biot-Savart Field Law


{d\bf B}({\bf r}_2) = \frac {\mu _0}{4\pi } I_1 \frac {d{\bf l}_1 \times {\bf r}_{12}}{|{\bf r}_{12}|^3} (2.32)

{\bf B}({\bf r}_2) = \frac {\mu _0}{4\pi } I_1 \oint _1 \frac {d{\bf l}_1 \times {\bf r}_{12}}{|{\bf r}_{12}|^3} (2.33)

• Note that this is empirically derived, and r12 denotes a vector running from a source point (‘1’) to the field
/ test point (‘2’).
• For a current density in a continuous medium, we can replace I1 dl1 with J1 dV1 :

{\bf B}({\bf r}_2) = \frac {\mu _0}{4\pi } \int _{\Vol _1} \frac {{\bf J}({\bf r}_1)\times {\bf r}_{12}}{|{\bf r}_{12}|^3}\, d\Vol _1 (2.34)

• This integral form implies that ∇2 ·B = 0, which indicates a lack of magnetic monopoles.

We can prove the last statement by appealing to the mathematical identity for the divergence of a vector cross
product, i.e. ∇·(F × G) = (∇ × F)·G − (∇ × G)·F :

\label {eq:243} {\bf \nabla }_2 {\bf \cdot B} &= \frac {\mu _0}{4\pi } \int _{\Vol _1} {\bf \nabla }_2 \cdot \frac {{\bf J}({\bf r}_1)\times {\bf r}_{12}}{|{\bf r}_{12}|^3} d\Vol _1 \\ &= \frac {\mu _0}{4\pi } \int _{\Vol _1} - {\bf J}({\bf r}_1) \cdot \biggl ( {\bf \nabla }_2 \times \frac {{\bf r}_{12}}{|{\bf r}_{12}|^3}\biggr ) d\Vol _1

(2.35b)

PHAS0038: Term 1 Macroscopic Fields 10


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

where, since we are taking the divergence with respect to point r2 , the term involving ∇2 × J(r1 ) is zero. But
now we can use two identities:

1. ∇2 (1/r12 ) = −r12 /|r12 |3


2. ∇ × (∇) ≡ 0

This shows that the integral on the right-hand size of equation (2.35a) is zero, and hence there are no magnetic
monopoles in this theory.

• The original, integral form of Ampère’s Law for the static case is:
\oint _C {\bf B} \cdot \, d{\bf \ell } = \mu _0I
(2.36)

where the current is that flowing through the area S enclosed by the path C.
• The differential form comes from writing I = S J · ndS. It follows that, for the static case (no explicit
R

time dependence):
{\bf \nabla } \times {\bf B} = \mu _0 {\bf J} (2.37)
• In the non-static case, we also have to account for time-varying electric field E as follows:

Ampère-Maxwell Law
{\bf \nabla } \times {\bf B} = \mu _0 {\bf J} + \mu _0 \epsilon _0 \frac {\partial {\bf E}}{\partial t} (2.38)

It is useful, in the context of this topic, to consider a capacitor being charged with a constant current, I. Using
Ampère’s law (in static form) we see: \oint {\bf B \cdot } d{\bf \ell } = \mu _0 \int _S {\bf J \cdot n} \, dS
(2.39)

Now consider a loop, C, around the wire leading to one plate of the capacitor, and two different surfaces bounded
by C, as shown in Figure 2.4:

1. A flat surface cutting across the wire;


2. A surface passing between the plates of the capacitor, but not cutting the wire

Figure 2.4: Amperian loops on a charging capacitor.

It is clear that these will give two different answers for the integral over the current density: in the first, the
answer will be I, and in the second it will be zero. This is clearly inconsistent – the full form of Ampère’s law
insists that the choice of surface be arbitrary. The resolution to the problem involves the time-dependence of the
electric field between the capacitor plates. We will investigate this further in Chapter 5, on Maxwell’s Equations.

PHAS0038: Term 1 Macroscopic Fields 11


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.3.1 Faraday’s Law

• Electromotive force (emf) is equivalent to a potential difference.


• Often encountered in terms of circuits, which have inductance that relates magnetic flux and current.
• Around a circuit, the emf, E, is defined by:

\mathcal {E} = \oint _C {\bf E^{\prime } \cdot } d{\bf l} (2.40)

where E′ is evaluated in the rest frame of a circuit that may itself be moving.
• Faraday’s Law (integral form) relates the emf to the rate of change of magnetic flux with time:

\mathcal {E} = -\frac {d\Phi }{dt} (2.41)

We define the magnetic flux, Φ, as:


\Phi = \int _S {\bf B \cdot n} dS (2.42)

in other words, the surface integral of the component of magnetic field crossing that surface. Now, using the
definition of emf we can relate the electric field to the derivative of the magnetic field:

\oint _C {\bf E^{\prime } \cdot } d{\bf l} = -\frac {d}{dt} \int _S {\bf B \cdot n} \, dS
(2.43)

Provided that the circuit being considered does not change or move with time, we can take the time derivative
inside the integral and make it partial. We can also use the electric H field E Rin our frame of reference, which is
the rest frame of the conductor. We can also use Stokes’ theorem [ C F·dl = S ∇ × F · ndS] on the line integral
of E to obtain the surface integral of ∇ × E. Hence:
\int _S {\bf \nabla \times E \cdot n} \, dS = -\int _S \frac {\partial {\bf B}}{\partial t} \cdot {\bf n} \, dS
(2.44)

Since this must be true for all fixed surfaces S, we find:

The differential form of Faraday’s Law

{\bf \nabla \times E} = - \frac {\partial {\bf B}}{\partial t} (2.45)

• When the magnetic field is static, this reduces to the conservative electric field E, i.e. ∇ × E = 0 as used
in electrostatic problems.
• Notice the minus sign: Lenz’s Law states that any induced currents arising from Faraday’s Law flow in
the sense that the magnetic field they produce (induced field) opposes the original change in magnetic flux.

2.4 Magnetic Vector Potential


The solution of many electrostatic problems is made easier by working in terms of the potential rather than the
electric field directly. The same idea can be applied to the magnetic field, though the eventual solution is rather
more complex.

• Since ∇ × ∇φ = 0 we know that we can write E = −∇ϕ when ∂B/∂t = 0

PHAS0038: Term 1 Macroscopic Fields 12


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

• Similarly, we know that ∇ · B = 0


• The relevant identity here is ∇ · (∇ × A) = 0
• We can then write generally:

The Magnetic Vector Potential


{\bf B} = {\bf \nabla } \times {\bf A} (2.46)

where A is the vector potential.


When we consider the form of the vector potential, it should be immediately apparent (by analogy with the
electric field as gradient of the potential) that there is a freedom in how we define it:
{\bf A}^{\prime } \rightarrow {\bf A} + {\bf \nabla } f, (2.47)
for any choice of scalar function f , results in the same B field since ∇ × (∇f ) = 0. This invariance under a
transformation is called gauge invariance. It should not be surprising: the electrostatic potential, φ, is not defined
up to an arbitrary additive constant (and thus all potentials are actually potential differences with respect to
some reference value).
There are different ways of choosing the vector potential which help with different situations. Consider a situation
where the electric field does not change with time. Then we write Ampère’s Law as:
{\bf \nabla \times B} = {\bf \nabla \times (\nabla \times A)} &= \mu _0 {\bf J} \\ {\bf \nabla (\nabla \cdot A)} - \nabla ^2{\bf A} &= \mu _0 {\bf J}
(2.48b)

• The Coulomb gauge is: ∇ · A = 0


• It leads to the following expression for finding solutions for the vector potential:
∇2 A = −µ0 J.
• By analogy with Poisson’s equation for electric potential V , ∇2 V = −ρ/ϵ0 , we can obtain a solution of
the form:
{\bf A}({\bf r}_2) = \frac {\mu _0}{4\pi } \int _{\Vol _1} \frac {{\bf J}({\bf r}_1)}{|{\bf r}_2-{\bf r}_1|}\,d\Vol _1 (2.49)

• The spatial distribution of current density thus determines the magnetic vector potential.
• There are other choices of gauge. For instance, the Lorenz gauge is ∇ · A = −µ0 ϵ0 .(∂V /∂t)
• Gauge invariance is a more general phenomenon.
• Solving for vector potential is (generally) more difficult than solving for the electrostatic potential.
• Note that the electric field can no longer be expressed as the gradient of a scalar potential if there is a
time-varying B field:
{\bf E} = -{\bf \nabla }\varphi - \frac {\partial {\bf A}}{\partial t} (2.50)

This last result can be proved rather easily. Consider the Maxwell equation for the curl of the electric field:

{\bf \nabla \times E} = - \frac {\partial {\bf B}}{\partial t} (2.51)

and substitute in the form of B = ∇ × A:

{\bf \nabla \times E} + \frac {\partial }{\partial t}{\bf \nabla \times A} = 0 (2.52)

The vector E + ∂A/∂t therefore has zero curl. We know from identities that it can be written as a gradient of
a scalar:
{\bf E} + \frac {\partial {\bf A}}{\partial t}= - {\bf \nabla }\varphi (2.53)
So, rearranging, we find that E = −∇ϕ − ∂A/∂t

PHAS0038: Term 1 Macroscopic Fields 13


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.5 Magnetic Intensity


As we saw with the electric field, E, the introduction of a medium other than vacuum results in changes to
Maxwell’s equations. These changes can be handled by using an alternative field which includes the effects of the
medium implicitly. We will now do the same for magnetic fields. A word of caution: non-linear magnetic media
are much more common than non-linear electric media; we will deal with these rather interesting materials in
Chapter 4 on Ferromagnetism.

• We introduced the polarization of a dielectric material, P ∝ E.

• Similarly, we introduce a quantity, proportional to the magnetic induction B (more commonly referred to
as the magnetic field).

• This quantity is the magnetization, M. It arises from microscopic, localized currents in the medium.

• It describes the response of a material to the magnetic induction.

• Electrons can be modelled as moving in tiny loops around atoms, thus forming microscopic currents: we
can use this concept to define M as the large-scale magnetic dipole moment per unit volume.

Let us consider the vector potential at a test point r2 due to a small volume of magnetised material at a source
point r1 (we will see later that this is given by the expression below). This small volume will have magnetic
moment ∆m = M(r1 ) δV1 . Then we can write:

{\bf A}({\bf r}_2) &= \frac {\mu _0}{4\pi } \int _\Vol \frac {{\bf \Delta m} \times {\bf r}_{12}}{|{\bf r}_{12}|^3} \\ &= \frac {\mu _0}{4\pi } \int _\Vol \frac {{\bf M}({\bf r}_1) \times {\bf r}_{12}}{|{\bf r}_{12}|^3}\,d\Vol _1 \\ &= \frac {\mu _0}{4\pi } \int _\Vol {\bf M}({\bf r}_1) \times {\bf \nabla }_1\frac {1}{|{\bf r}_{12}|}\,d\Vol _1

(2.54c)

Now we use the expansion of ∇ × (φF), with F = M and φ = 1/r12 to write

{\bf F} \times {\bf \nabla }\phi &= \phi {\bf \nabla } \times {\bf F} - {\bf \nabla } \times (\phi {\bf F}) \\ {\bf A}({\bf r}_2) &= \frac {\mu _0}{4\pi } \int _\Vol \biggl [ \frac {{\bf \nabla }_1 \times {\bf M}({\bf r}_1)}{r_{12}} - {\bf \nabla }_1 \times \biggl ( \frac {{\bf M} ({\bf r}_1)}{r_{12}} \biggr )\biggr ]\,d\Vol _1

(2.55b)

R R
Now we use the theorem V
∇ × F dV = S
n × F dS to write:

{\bf A}({\bf r}_2) &= \frac {\mu _0}{4\pi } \int _\Vol \biggl [ \frac {{\bf \nabla }_1 \times {\bf M}({\bf r}_1)}{r_{12}}\biggr ]\,d\Vol _1 - \frac {\mu _0}{4\pi } \int _S \frac {{\bf n \times M} ({\bf r}_1)}{r_{12}} \, dS_1 \\ &= \frac {\mu _0}{4\pi } \int _\Vol \biggl [ \frac {{\bf \nabla }_1 \times {\bf M}({\bf r}_1)}{r_{12}}\biggr ]\,d\Vol _1 + \frac {\mu _0}{4\pi } \int _S \frac {{\bf M}({\bf r}_1) \times {\bf n}}{r_{12}} \,dS_1

(2.56b)

• This then leads us to the magnetization current densities:

• We formally define:
{\bf J}_M = {\bf \nabla \times M} (2.57)

{\bf j}_M = {\bf M \times n} (2.58)

• JM is the volume magnetization current density

PHAS0038: Term 1 Macroscopic Fields 14


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

Figure 2.5: Origins of the magnetization surface current.

• jM is the surface magnetization current density – arising at boundaries where magnetization changes
discontinuously (see Figure 2.5).

It is clear that there will be no large-scale, (volume) bound current density where the magnetization is uniform.
So within the bulk of the rod shown below (cross section) there is a bound current density given by JM = ∇ × M,
and at the surface there is a bound surface current per unit length given by jM = M × n, where n is a unit
vector in the direction of the outward normal to the surface. JM is a current per unit area, where the area is
perpendicular to the direction of flow, and jM is a current per unit length, where the length is in the plane of the
surface and also perpendicular to the direction of the surface current. These bound currents are the net effect
of the collective motion of many microscopic currents (associated with magnetic dipoles) on larger scales.

• We now move on to considering how linear magnetic media behave.


• We know that ∇ × B = µ0 J, where J arises from free and magnetization current.
• Thus we have, in general, J = Jf + JM .
• Here Jf is due to the motion of free charges, and JM = ∇ × M, as described above.
• Hence ∇ × B = µ0 (Jf + ∇ × M) or ∇ × ( µB0 − M ) = Jf
• We then define H, the magnetic intensity, as

Magnetic Intensity
{\bf H} = \frac {{\bf B}}{\mu _0} - {\bf M} (2.59)

This yields ∇ × H = Jf .
The magnetic intensity serves a similar purpose to the electric displacement, in accounting for the response of
the medium as well as the magnetic induction. We can rewrite this, using Stokes’ theorem:
\int _S {\bf \nabla \times H \cdot n}dS &= \int _S {\bf J}_f \cdot {\bf n}\,dS \\ \oint _C {\bf H \cdot }d{\bf l} &= \int _S {\bf J}_f \cdot {\bf n}\,dS = I_f.

(2.60b)

This tells us that the integral of the magnetic intensity along a closed loop is equal to the free current flowing
across any surface defined by that loop. It also gives the units as amperes per metre (the same units as the
magnetization). It is important to note that the three quantities that we have defined so far (the magnetic
induction, B, the magnetization, M and the magnetic intensity, H) are not necessarily parallel; this will be
important when considering ferromagnetism in particular.

PHAS0038: Term 1 Macroscopic Fields 15


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

2.5.1 Magnetic Susceptibility

• For a linear, isotropic material, we assert (based on experimental observations):

M = \chi _m {\bf H} (2.61)

where χm is the magnetic susceptibility

• We can write B = µ0 (1 + χm )H

• If χm > 0 we have a paramagnetic material

• If χm < 0 we have a diamagnetic material

• Note that χm can depend on temperature, but is generally small for these materials (less than 10−5 )

2.6 Interfaces and Boundary Conditions


• Understanding how the different field vectors change at interfaces is important.

• We need to consider both medium/vacuum and medium/medium interfaces.

• We will consider the electric and magnetic fields in two groups:

– D and B together
– E and H together

• We want to know what is conserved

2.6.1 Normal Components

• First notice that we can write similar equations for D and B:

{\bf \nabla \cdot D} = \rho _f (2.62)

{\bf \nabla \cdot B} = 0 (2.63)

Figure 2.6: Small cylinder at interface.

• Consider an interface with no free charges

• Consider the small cylinder of Fig. 2.6, height dh, area dS.

PHAS0038: Term 1 Macroscopic Fields 16


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

Gauss’ theorem tells us:

\int _\Vol {\bf \nabla \cdot D} \, d\Vol = \int _\Vol \rho _f \, d\Vol
(2.64)

\rightarrow \oint _S {\bf D \cdot n} dS = \int _\Vol \rho _f \, d\Vol (2.65)

For the magnetic field, we find: \oint _S {\bf B \cdot n} \, dS = 0


(2.66)

What is the flux of D through the box? Take the limit dh → 0, and for an interface with no free charge we find:

\oint _S {\bf D \cdot n}\,dS &\rightarrow {\bf D}_2 \cdot {\bf n_S}dS -{\bf D}_1 \cdot {\bf n_S}dS \\ {\bf D}_2 \cdot {\bf n_S} &= {\bf D}_1 \cdot {\bf n_S} \\ {\bf D}_{1\perp } &= {\bf D}_{2\perp } \\ {\bf B}_{1\perp } &= {\bf B}_{2\perp }

(2.67d)

where the opposite signs on the displacement vectors come from their opposing directions (compared to the surface
normal nS pointing from medium 1 into 2). This implies that the normal components of D are continuous across
an interface with no free charges, while the normal components of B are always continuous. Note that, in fact:

{\bf D}_{2\perp } - {\bf D}_{1\perp } = \sigma _f (2.68)

2.6.2 Tangential Components

Figure 2.7: Small loop at interface.

• First notice that we can write similar equations for E and H:

{\bf \nabla \times E} = -\frac {\partial {\bf B}}{\partial t} (2.69)

{\bf \nabla \times H} = {\bf J_f} (2.70)

• Consider an interface with no free current.

• Consider the small loop of Fig.2.7, height dh, length dl.

PHAS0038: Term 1 Macroscopic Fields 17


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

Stokes’ theorem tells us


\int _S {\bf \nabla \times E \cdot n}\,dS &= - \int _S \frac {\partial {\bf B}}{\partial t} \cdot {\bf n} \, dS \\ \oint _C {\bf E} \cdot d{\bf l} &= - \int _S \frac {\partial {\bf B}}{\partial t} \cdot {\bf n} \, dS

(2.71b)

Taking the limit dh → 0, da = dl dh → 0, we find:

\int _S {\bf \nabla \times E \cdot n}\,dS = \oint _C {\bf E} \cdot d{\bf l}=0
(2.72)

−−→ −−→
But this can be written as E2 · AB + E1 · CD. As the vectors from A to B and from C to D have opposite
directions, we write:

{\bf E}_1 \cdot d{\bf l} &= {\bf E}_2 \cdot d{\bf l} \\ {\bf E}_{1\parallel } &= {\bf E}_{2\parallel }

(2.73b)

And for an interface with no free surface current (surface magnetization currents are irrelevant) we have a similar
result for H:

{\bf H}_1 \cdot d{\bf l} &= {\bf H}_2 \cdot d{\bf l} \\ {\bf H}_{1\parallel } &= {\bf H}_{2\parallel }

(2.74b)

In fact,
{H}_{2\parallel } - {H}_{1\parallel } = I_{enc} = J_{s\perp } (2.75)
or, using the surface normal nS pointing from medium 1 to 2:

{\bf n_S} \times ({\bf H}_{2} - {\bf H}_{1}) = {\bf J_{s}} (2.76)

This implies that the tangential components of the E and H fields are conserved subject to the conditions
explained above.

• Normal components of B are continuous across an interface


• Normal components of D are continuous across an interface with no free charges
• Tangential components of E are continuous across an interface
• Tangential components of H are continuous across an interface with no free currents

2.7 Summary of Linear Media

• Linear: χe is independent of E (or χm independent of B)


• Isotropic: P is parallel to E (or M is parallel to H)
• Homogeneous: χe is position-independent
• D = ϵ0 E + P
• P = ϵ0 χe E so D = ϵE, with ϵ = ϵ0 (1 + χe )
• ∇ · D = ρf

PHAS0038: Term 1 Macroscopic Fields 18


PHAS0038: Electromagnetic Theory Chapter 2 - Macroscopic Fields

• H = B/µ0 − M
• M = χm H so B = µ0 µr H with µr = 1 + χm

• ∇ × H = Jf
• continuous across an interface:
– B⊥
– D⊥ (when no free charges)
– E∥
– H∥ (when no free currents)

PHAS0038: Term 1 Macroscopic Fields 19

You might also like