You are on page 1of 39

Chapter 2

The macroscopic laws of electromagnetism

The matter of laws is organized in like this: firstly there are presented the
laws that rule the macroscopic theory of electromagnetism (in Chapter 2), and
then a few of the conclusions and consequences derived from the laws’ system
(in Chapter 3). In paragraph 3.5 there are presented the units of the electric and
magnetic quantities, and in Chapter 4 there is presented the use of these laws for
some significant problems computation.
The laws of macroscopic theory of electromagnetism represent a set
of mathematical equations coherent from the point of respecting the truth
criteria, completeness and non-contradiction, relation that connect the
quantities which characterize the electromagnetic field and the
electromagnetic state of the bodies in variable regime, gathering in a
mathematical “shape” the electromagnetic field phenomenology.
During the presentation of each law, after the statement and the comment
of its mathematical formulation, we follow to mark out its physical significance
and its most important consequences, firstly in variable regime, then in various
particular regimes: quasi-stationary regimes (in which the variations in time of
some of the quantities are sufficiently slow so they can be neglected), steady
state (stationary) regimes (in which the quantities are time-invariant, but there
are energy transformations) and static regimes (which are stationary regimes
without energy transformations).

2.1. Electric flux law (Gauss law of the electric flux)

Statement: Electric flux ΨΣ through any closed surface (Σ ) is equal at


any moment to the electric charge qVΣ form domain (VΣ ) bounded by the
surface (Σ ) :

ΨΣ = qVΣ . (2.1.1)

33
dA = dAext D

(VΣ ) ρv

dv
(Σ )

Figure 2.1.1
If the electric charge from domain (VΣ ) has the volume distribution
density ρ v (see Figure 2.1.1), then relation (2.1.1) can be explicitly written:

∫ D ⋅ dA = ∫ D ⋅dAext = ∫ ρ v ⋅ dv (2.1.2)
(Σ ) (Σ ) (VΣ )

and it represents the integral form electric flux law.


Applying Gauss-Ostrogradski relation, one obtains:

∫ D ⋅ dA = ∫ (div D ) ⋅ dv = ∫ ρ v ⋅ dv, (2.1.3)


(Σ ) (VΣ ) (VΣ )

relation which, being true for any arbitrary considered domain (VΣ ) ,
imposes the equality of the integrands:

div D = ρ v , (2.1.4)

expression which represent the local form of the electric flux law for
continuity domains.
Electric flux law emphasizes one of the causes that generates electric
field, namely charged bodies.
div D = 0

div D > 0 div D < 0

Figure 2.1.2

34
It is well-known that, for any field vector, the points from the space in
which the divergence is nonzero they represent end points of field lines (starting
point if the divergence is positive, and, respectively, end points id the
divergence is negative). It results that the lines of the electric field produced
by charges are open curves, starting from the positive charged bodies and
ending on the negative charged bodies (see Figure 2.1.2).

2.2. Magnetic flux law

Statement: Magnetic flux Φ Σ through any closed surface (Σ ) is zero in


any moment:
Φ Σ = 0. (2.2.1)
Replacing in equation (2.2.1) he expression of the magnetic flux, one
obtains the integral form of the magnetic flux law:

∫ B ⋅ dA = ∫ B ⋅ dAext = ∫ (div B ) ⋅ dv = 0. (2.2.2)


(Σ ) (Σ ) (VΣ )

Because this relation is valid for any space domain (VΣ ), it results the
local form for continuity domains of the magnetic flux law:

div B = 0. (2.2.3)
According to relation (2.2.3), the magnetic flux density field vector is a
solenoidal one (without sources). This fact underlines, on one hand, the non-
existence of the magnetic charges similar to electric charges and, on the other
hand, inexistence of some points – extremity of magnetic field lines.
Therefore the magnetic field lines are not open curves.
An immediate consequence of the magnetic flux law magnetic flux
through any open surface bounded by the same closed curve is the same.
In order to prove this statement one will consider an arbitrary closed
( )
curve (Γ ), and any two arbitrary open surfaces S Γ, j and (S Γ, k ) , which rest on
curve (Γ ) - see Figure 2.2.1.

35
dA j = dA ext

(SΓ , j )
dA k = −dAext (Σ )

(SΓ ,k )

(Γ ) dl

Figure 2.2.1
By convention we associate the line oriented element dl with each of the
area oriented elements dA k and dA j according to the right corkscrew rule, we
write the relation (2.2.2) for the closed curve (Σ ) , reunion of the open surfaces
( )
S Γ, j and (S Γ, k ) . In these conditions, the integral on (Σ ) will be the sum of the
( )
integral on S Γ, j and (S Γ, k ) :

∫ B ⋅ dAext = ∫ B ⋅ dAext + ∫ B ⋅ dAext = ∫ B ⋅ dA j − ∫ B ⋅ dAk = 0. (2.2.4)


(Σ ) (S Γ, j ) (S Γ, k ) (S Γ, j ) (S Γ, k )
Therefore

∫ B ⋅ dA j = ∫ B ⋅ dAk = ∫ B ⋅ dAk
(S Γ, j ) (S Γ, k ) (S Γ )
(2.2.5)

for any open surface (S Γ ) which sits on the closed curve (Γ ), which shows that
the magnetic flux has a unique value through all open surfaces bounded by
the same closed curve.
Moreover, the vector identity

(
div curl A = 0 ) (2.2.6)
allows the introduction of a new quantity, called magnetic vector potential and
denoted by A with relation:

curl A = B. (2.2.7)

36
As vector A is uniquely determined only if we know its divergence, it is
common that in stationary regime to adopt the calibration condition

div A = 0 (2.2.8)

such that the field vector A to be a solenoidal one too.


Applying the Stokes’ theorem, the magnetic flux which flows through an
arbitrary open surface (S Γ ) , which sits on a closed curve (Γ ) , can be expressed
by line integral of the of the magnetic potential vector on the curve (Γ ) :

Φ SΓ = ∫ B ⋅ dA = ∫ (curl A)⋅ dA = ∫ A ⋅ dl. (2.2.9)


( SΓ ) ( SΓ ) (Γ )

This shows that magnetic flux through an open surface depends only
on the closed curve that bounds it.

2.3. The magnetic field constitutive law


(the connection law between magnetic flux density, magnetic field
strength and magnetization)

Statement: For any point and for any time moment, between magnetic
flux density vector B , magnetic field strength vector H and magnetization
vector M exists the following relation:

( )
B = µ0 ⋅ H + M . (2.3.1)

We mention this law has only local form.


By expressing the magnetization vector M by its two components,
namely temporary component M t and permanent component M p , the
magnetic constitutive law has the following form:

B = µ0 ⋅ H + µ0 ⋅ M t + µ0 ⋅ M p (2.3.2)

and emphasizes as magnetic field generating cause the bodies that have
permanent magnetization.

37
2.4. Temporary magnetization law

Statement: For any point and for any time moment, temporary
magnetization M t depends on the magnetic field strength H :

Mt =Mt H . ( ) (2.4.1)

Function of magnetic materials’ nature, relation (2.4.1) has explicit


specific forms.
We remind a material is called isotropic if in any of its points there are
no privileged directions (in other words, its local properties does not depend on
the direction of a certain quantity) and is called linear if in any of its points
there are no privileged values (in other words, its local properties does not
depend on the value of a certain quantity).
Below there will be presented and discussed the various dependences
between the magnetic flux density and the magnetic field strength function of
magnetic materials’ nature.
♦ For isotropic linear materials, the dependence (2.4.1) has the form

M t = χm ⋅ H , (2.4.2)

where the proportionality coefficient χ m is a non-dimensional material constant


which is called magnetic susceptibility, and the magnetic constitutive law
becomes

B = µ0 ⋅ H + µ0 ⋅ M t + µ0 ⋅ M p = µ0 ⋅ H + µ0 ⋅ χ m ⋅ H + µ0 ⋅ M p =
(2.4.3)
= µ 0 ⋅ (1 + χ m ) ⋅ H + µ 0 ⋅ M p .

The non-dimensional material parameter


µr = 1 + χ m (2.4.4)
is called relative permeability, and the dimensional quantity
µ = µo ⋅ µ r (2.4.5)
is called absolute permeability.
Using these, the magnetic constitutive law in magnetic field becomes
B = µ ⋅ H + µ0 ⋅ M p = µ ⋅ H + I p, (2.4.6)

38
in which the quantity µ 0 ⋅ M p = I p is called magnetic polarization.

Ip B = μ⋅ H + I p

μ⋅ H B = μ⋅H
H H

Figure 2.4.1, a Figure 2.4.1, b

Vectors H , µ ⋅ H , B and I p = µ 0 ⋅ M p are placed from qualitative point


of view as in Figure 2.4.1,a case when M p ≠ 0 and, respectively, as in
Figure 2.4.1,b case for which M p = 0.

By the values of their relative permeability values, the materials are being
ordered as follows:
• diamagnetic materials (that have a µ r slightly less than one);
• paramagnetic (that have a µ r slightly more than one);
• ferromagnetic materials (with very big µ r , touch four figures and which,
care, from certain pair of values, they lose their linear character).

In the table below there are presented values of the relative permeability
for a couple of diamagnetic and paramagnetic materials, observing that for all of
them one can consider µ r ≅ 1.

Name of diamagnetic material µr

Hydrogen 1-0.063∙10-6
Copper 1-08.8∙10-6
Water 1-9∙10-6
Zinc 1-12∙10-6
Halite 1-12.6∙10-6

39
Name of diamagnetic material µr

Silver 1-19∙10-6
Mercury 1-25∙10-6
Bismuth 1-176∙10-6

Name of the paramagnetic material µr

Azoth 1+0.013∙10-6
Air 1+0.4∙10-6
Oxygen 1+1.9∙10-6
Aluminum 1+23∙10-6
Platinum 1+360∙10-6

♦ For linear anisotropic materials one notices that, if there is considered


a three-orthogonal coordinate axis system, in each point in the space each scalar
component of temporary magnetization vector depends in principle on all scalar
components of the magnetic field strength vector. In a Cartesian coordinate
system, for example, if H = Hx ⋅i + Hy ⋅ j + Hz ⋅k ,
M t = M t x ⋅ i + M t y ⋅ j + M t z ⋅ k and B = Bx ⋅ i + B y ⋅ j + Bz ⋅ k , then

M t = χ m ⋅ H x + χ m ⋅ H y + χ m ⋅ H z
 x xx xy xz

M t y = χ m yx ⋅ H x + χ m yy ⋅ H y + χ m yz ⋅ H z (2.4.7)

M t z = χ m zx ⋅ H x + χ m zy ⋅ H y + χ m zz ⋅ H z .

By defining the magnetic susceptibility tensor χ m by the attached


matrix:

40
χ m χ m xy χ m xz 
 xx 
χm =  χ m yx χ m yy χ m yz  , (2.4.8)
 
χ m χ m zy χ m zz 
 zx 
The law of temporary magnetization can be shortly written as

Mt = χm ⋅H (2.4.9)

and it leads to the following form of the magnetic constitutive law

B = µ0 ⋅ H + µ0 ⋅ M t + µ0 ⋅ M p = µ0 ⋅ H + µ0 ⋅ χ m ⋅ H + µ0 ⋅ M p =
( )
(2.4.10)
= µ0 ⋅ 1 + χ m ⋅ H + µ0 ⋅ M p = µ0 ⋅ µ r ⋅ H + µ0 ⋅ M p = µ ⋅ H + I p ,

where µ r = 1 + χ m is the relative permeability tensor, and µ = µ 0 ⋅ µ r is the


absolute permeability tensor.

Ip
B = μ⋅H + I p
B = μ⋅H

H
μ⋅H

Figure 2.4.2, a Figure 2.4.2, b

Relation (2.4.10) has the vector interpretation from Figure 2.4.2,a with
the particular case in Figure 2.4.2,b corresponding in which M p = 0.

If we explicitly write the relation (2.4.10):

 Bx   µ xx µ xy µ xz   H x   I p x 
       
 B y  =  µ yx µ yy µ yz  ⋅  H y  +  I p y  , (2.4.11)
 
 B  µ
 z   zx µ zy µ zz   H z   I p 
 z

41
we remark the fact that the matrix corresponding to the tensor µ is symmetrical
and positive defined. There are three orthogonal directions, called principal
magnetization directions, having the unit vectors u1 , u 2 , u 3 , for which the
scalar components of vectors B = B1 ⋅ u1 + B2 ⋅ u 2 + B3 ⋅ u 3 ,
H = H1 ⋅ u1 + H 2 ⋅ u 2 + H 3 ⋅ u 3 and I p = I p ⋅ u1 + I p ⋅ u 2 + I p ⋅ u 3 after
1 2 3
these three directions satisfy the matrix relation:

 B1   µ1 0 0   H1   I p1 
B  =  0 µ 0  ⋅ H  + I  .
 2  2   2   p2  (2.4.12)
 B3   0 0 µ 3   H 3   I p 
 3

♦ For nonlinear materials the dependence between the magnetic flux


density and magnetic field strength can be expressed (or at least approximated)
analytically (see Figure 2.4.3) by a real, nonlinear function f :

B = f (H ) . (2.4.13)

H
0
Figure 2.4.3
♦ One notices there are nonlinear materials (such as for example some
iron, nickel, cobalt alloys) for which the dependence of the relationship
between the magnetic flux density and the magnetic field strength cannot be
expressed (and not even approximated) analytically. These materials present a
specific phenomenon called magnetic hysteresis, which expresses the fact that
a functioning point from the curve which gives the dependence B = B(H )
has coordinates dependent on the succession of states followed to arrive in
that specific point, so it depends on the historical realizations of that state.

42
Below we present, from the qualitative point of view, this phenomenon
(see Figure 2.4.4). Starting from an initial state in which the material is not
magnetized, state corresponding to point O(0;0) in the plane (H ; B ) and
applying a magnetic field with increasing magnetic field strength, one notices
first a nonlinear increase of the magnetic flux density followed by a
saturation plateau until point A1 (H max ; Bmax ) . On this plateau, an increase of
the magnetic field strength does not lead to significant increases of the magnetic
flux density. OA1 curve is called first magnetization curve.

Figure 2.4.4
When the magnetic field strength is decreased, the values of the magnetic
flux density do not coincide to those corresponding to the first magnetization
curve for the same values of H . The decrease of the magnetic flux density is
realized by applying an increasing magnetic field but opposite as direction to the
initial one. To magnetic field cancellation it corresponds a non zero remanent
magnetic flux density – the functioning point A2 (0; Br ) . The magnetic field that
cancels the magnetic flux density is called coercive magnetic field –
functioning point A3 (− H c ;0 ).

43
The continuation of decreasing magnetic field strength until value − H max
is reached brings the functioning point A4 (− H max ;− Bmax ) in a point
A1 symmetric with respect to the origin of the coordinate system axis. Then, if
the magnetic field is increased, the functioning points are being placed on the
curve A4 A5 A6 A1 which is symmetric to curve A1 A2 A3 A4 with respect to the
origin. In this way the hysteresis cycle closes.
Ferromagnetic materials magnetize following irreversible evolutions,
which for a given value of H corresponding in principle three possible values
for B , corresponding to the first magnetization curve OA1 , to the descending
part A1 A2 A3 A4 or to ascending part A4 A5 A6 A1 of the hysteresis cycle.

We also mention that, once we arrive in point A , for example (H ; B )


coordinate, we decrease the value of the magnetic field by value ∆H , after
which we come back to the starting value point, we describe in this way a
secondary cycle. This cycle is very flat, it can be approximated by its median
and the transformation that took place can be considered as being quasi-
reversible.
In point P there are defined the following three relative permeabilities
(see Figure 2.4.4):
• static relative permeability µr, s , equal to the trigonometric tangent of the
angle between PO chord and OH axis:

B
µ r ,s = = tg α ; (2.4.14)
µ0 ⋅ H P
• dynamic relative permeability µ r, d , equal to the trigonometric tangent of the
angle between the trigonometric tangent in point P and the OH axis:

∆B
µ r ,d = lim = tg β ; (2.4.15)
∆H →0 µ 0 ⋅ ∆H
∆H > 0 P
• reversible relative permeability µ r, rev , equal to the trigonometric tangent
between the median of the secondary cycle and the OH axis:

∆B
µ r ,rev = lim = tg γ , (2.4.16)
∆H →0 µ 0 ⋅ ∆H
∆H < 0 P
with γ < Max {α ; β }.

44
According to the width of the hysteresis cycle, the ferromagnetic materials
can be divided into two big categories:
o soft magnetic materials, characterized by narrow cycles, respectively small
coercive fields (see dependence (a) from Figure 2.4.5);
o hard magnetic materials, characterized by wide cycles, respectively big
coercive fields (see dependence (b) from Figure 2.4.5).

0 H

Figure 2.4.5
The curves from Figure 2.4.5 gives us only a qualitative information, that
do not make out the real magnitude order of the ratio between the widths of the
hysteresis cycles. For comparison, we present below the values of the remanent
flux density (Br ) and of the coercive field (H c ) for a few of the soft
ferromagnetic materials, respectively for hard magnetic materials.

(B r ) (H c )
Name of the soft ferromagnetic material
[T ] [ A / m]
Super Malloy (79% Ni; 15% Fe; 5% Mo; 0.5% Mn) 0.6 0.4
Perm alloy (78.5% Ni; 21.5% Fe) 0.6 4
Pure Iron 1.4 4
Nickel – Zinc Ferrite 0.13 10

45
(Br ) (H c )
Name of the hard ferromagnetic material
[T ] [ A / m]
Electrotechnical Steel (with 4% Si) 1.8 40
Manganese – Zinc Ferrite 0.15 20
Steel (cu 1% C) .7 5∙103
Chrome Steel, Wolfram Steel 1.1 5∙103
Alnico (12% Al; 20% Ni; 5% Co; 63% Fe) 0.73 34∙103
Oerstit (20% Ni; 30% Co; 20% Ti; 30% Fe) 0.55 65∙103
Cobalt Ferrite 0.16 90∙103
Barium Ferrite 0.35 200∙103
Cobalt – Platinum Alloy (77% Pt; 23% Co) 0.45 260∙103

One can notice that, for both categories, the magnitude order of the
remanent flux densities is the same, while the values of coercive fields are
slightly different. For example, taking into consideration two materials with the
same value of the remanent flux density, manganese – zinc ferrite from the soft
materials category and cobalt ferrite from the hard materials category, one can
notice a ratio between the widths of the hysteresis cycles, expressed by the ratio
of their coercive fields, of 1/4.500.
Finally, it’s important to mention that the completion of a hysteresis
cycle is accompanied by energy transfer from the magnetic field to the
ferromagnetic body, the volume density of this energy being proportional to
the cycle’s area (as Warburg theorem proves, theorem that will be presented in a
following chapter of this book).
We also remember that the properties of the ferromagnetic material
are greatly influenced by temperature, these properties even disappearing
for certain limit values of temperature, proper for each material, called
Curie points. Curie point for cobalt is 1,137 0 C , for iron is 7530 C , and for
0
nickel is 376 C . Increasing the temperature over the Curie point value
leads to the transformation of ferromagnetic materials into paramagnetic
materials.

46
2.5. The electric field constitutive law
(the connection law between electric flux density, electric field
strength and electric polarization)

Statement: For any point and for any time moment, between electric
flux density vector D , electric field strength vector E and polarization
vector P there is the following relation:

D = ε 0 ⋅ E + P. (2.5.1)

As for the magnetic constitutive law, this law has only local form.
Expressing the electric polarization vector P by its two components,
namely the temporary P t and the permanent component P p , the electric
constitutive law has the following form:

D = ε 0 ⋅ E + Pt + Pp (2.5.2)

and emphasizes as electric field generating cause – bodies that have


permanent electric polarization.

2.6. The law of temporary electric polarization

Statement: For any point and for any time moment, temporary
electric polarization P t depends on electric field strength E :

()
Pt = Pt E . (2.6.1)

Depending on the electric material nature, relation (2.6.1) has explicit


particular forms.
As follows we present and comment various types of dependences
between electric displacement and electric field strength as a function of
magnetic materials nature.
♦ For linear and isotropic materials, the dependence (2.6.1) has the form

P t = ε 0 ⋅ χ e ⋅ E, (2.6.2)

47
where the proportionality coefficient χ e is a material numeric (non-
dimensional) constant called electric susceptibility, and the electric constitutive
law becomes

D = ε 0 ⋅ E + Pt + Pp = ε 0 ⋅ E + ε 0 ⋅ χ e ⋅ E + Pp = ε 0 ⋅ (1 + χ e ) ⋅ E + Pp . (2.6.3)
The numeric quantity
ε r = 1+ χ e (2.6.4)
is called relative permittivity, and the numeric quantity
ε = ε0 ⋅εr (2.6.5)
is called the absolute permittivity or electric constant.
Using these notations, the electric constitutive law has the form

D = ε ⋅ E + P p. (2.6.6)

Pp D =ε ⋅E + Pp

E D =ε ⋅E
E
ε ⋅E

Figure 2.6.1, a Figure 2.6.1, b


Vectors E , ε ⋅ E , D and P p re qualitative places as in Figure 2.6.1,a for
the case in which P p ≠ 0 and, respectively, as in Figure 2.6.1,b for the case in
which P p = 0.

Depending on their relative permittivities, materials can be ranked as


follows:
• diaelectric materials (which have ε r closed to 1);
• paraelectric materials (which have unit order ε r );
• ferroelectric materials (which have very big ε r , thousands order and
which, if bigger than certain pair of values, they loose their linear
characteristics).
In the table below e present the values of the relative permeabilities for
some of the diaelectric and paraelectric materials.

48
Name of the material Aggregation state εr

H2 1.0003
Air 1.0006
O2 1.0006
CO Gaseous 1.0007
CO2 1.001
CH4 1.001
C2H6 1.0015

Air ( − 1910 C ) 1.43

Transformer oil ( 200 C ) 2.2

Acetone ( 200 C ) 21.2

Ethyl alcohol (150 C ) 26


Liquid
0
Methyl alcohol (15 C ) 32.5

Nitrobenzene (180 C ) 36

Distilled water ( 200 C ) 81.1

Hydrocyanic acid (150 C ) 95


Paraffin 2.2
Polyethylene 2.3
Polyamide Solid 2.4
Insulant paper 2.4
Bakelite 2.8

Plexiglas Solid 3.0÷3.6

Pressboard 3.4÷4.3

49
Name of the material Aggregation state εr

Ebonite 2.5÷5.0
Rubber 3.0÷6.0
Quartz glass 4.0÷4.2
Porcelain 5.0÷6.5
ClNa Solid 5.5
Mica 5.0÷7.0
Glass 5.5÷8.0
SO4K2 8.35
Diamond 16.5

♦ For linear and anisotropic materials, one can notice that, if a three-
orthogonal coordinate system is taken into consideration, in each point from the
space, each scalar component of temporary electric polarization vector depends,
in principle, on all scalar components of electric field strength vector. In a
Cartesian coordinate system, for example, if E = E x ⋅ i + E y ⋅ j + E z ⋅ k ,
P t = Pt x ⋅ i + Pt y ⋅ j + Pt z ⋅ k and

D = Dx ⋅ i + D y ⋅ j + Dz ⋅ k , then

 Pt = ε 0 ⋅ χ e ⋅ E x + ε 0 ⋅ χ e ⋅ E y + ε 0 ⋅ χ e ⋅ E z
 x xx xy xz

 Pt y = ε 0 ⋅ χ e yx ⋅ E x + ε 0 ⋅ χ e yy ⋅ E y + ε 0 ⋅ χ e yz ⋅ E z (2.6.7)

 Pt z = ε 0 ⋅ χ e zx ⋅ E x + ε 0 ⋅ χ e zy ⋅ E y + ε 0 ⋅ χ e zz ⋅ E z .

Defining the electric susceptibility tensor χ e by the attached matrix:

χe χ e xy χ e xz 
 xx 
χ e =  χ e yx χ e yy χ e yz  , (2.6.8)
 
χe χ e zy χ e zz 
 zx 
the temporary electric polarization law can be shortly written

50
Pt = ε 0 ⋅ χ e ⋅ E (2.6.9)

and the following magnetic constitutive law results

D = ε 0 ⋅ E + Pt + P p = ε 0 ⋅ E + ε 0 ⋅ χ e ⋅ E + P p =
( )
(2.6.10)
= ε0 ⋅ 1+ χ e ⋅ E + P p = ε0 ⋅ε r ⋅ E + P p = ε ⋅ E + P p ,

where ε r = 1 + χ e is the relative permittivity tensor, and ε = ε 0 ⋅ ε r is the


absolute permittivity tensor.

Pp
D =ε ⋅E + Pp

D =ε ⋅E
E
ε ⋅E
E

Figure 2.6.2, a Figure 2.6.2, b


Relation (2.6.10) has the qualitative vector interpretation from
Figure 2.6.2,a with the particularization from Figure 2.6.2,b corresponding to
situation for which P p = 0.

If relation (2.6.10) is written explicitly:

 Dx  ε xx ε xy ε xz   E x   Pp x 
       
= ε
 y   yx
D ε yy ε ⋅
yz   y 
E +  Ppy  , (2.6.11)
 D  ε     
 z   zx ε zy ε zz   E z   Pp z 

it can be noticed that the matrix corresponding to tensor ε is symmetrical and


positive defined. So, there are three orthogonal directions (called electrization
principal directions), having the unit vectors u1 , u 2 , u 3 for which the scalar
components of the vectors D = D1 ⋅ u1 + D2 ⋅ u 2 + D3 ⋅ u 3 ,
E = E1 ⋅ u1 + E2 ⋅ u 2 + E3 ⋅ u 3 and P p = P p ⋅ u1 + P p ⋅ u 2 + P p ⋅ u 3 along to
1 2 3
these three directions satisfy the matrix relation:

51
 Dx  ε 1 0 0   E x   Pp x 
       
 D y  = 0 ε2 0  ⋅  E y  +  Pp y  . (2.6.12)
 
 D  0
 z 0 ε 3   E z   P 
 pz 

2.7. Electromagnetic induction law (Faraday law)

Statement: The electromotive force (emf) uΓ along a closed curve (Γ )


is equal to the rate of decrease (in time) of the magnetic flux Φ S Γ across any
surface (SΓ ) bordered by the closed curve (Γ ) :

dΦ S Γ
uΓ = − . (2.7.1)
dt

Considering the definition relations of emf and of the magnetic flux it


results the integral form of the law:
d
∫ E ⋅ dl = − dt ∫ B ⋅ dA. (2.7.2)
(Γ ) (S Γ )

B dA

(S Γ )

E
(Γ ) dl

Figure 2.7.1
The electromagnetic induction law has the above form only with the
condition that the reference direction of the closed curve (Γ ) (the reference
direction of the oriented line element dl ) and the direction of the normal to
the surface (SΓ ) (the oriented area element dA ) are associated according to
right corkscrew rule (see Figure 2.7.1).

52
For moving media, the integration domains follows the bodies in their
movement, and the derivative with respect to time of the magnetic flux is a
substantial derivative and it is computed using the following relation:

∂B 
d
dt (S∫ )
B ⋅ dA = ∫  ∂t + w ⋅ div B + curl B × w (
 ⋅ dA, ) (2.7.3)
Γ (S Γ)  

where w is the local speed vector of the medium.


Using Stoke relation, it results that

∫ E ⋅ dl = ∫ (curl E )⋅ dA (2.7.4)
(Γ ) (S Γ )

and

∫ [curl (B × w)]⋅ dA = ∫ (B × w)⋅ dl . (2.7.5)


(S Γ ) (Γ )

Taking into account the local form (2.2.3) of the magnetic flux law, one
obtains a new integral form of the electromagnetic induction law:

uΓ = ∫ E ⋅ dl = − ∫
∂B
∂t
(
⋅ dA − ∫ B × w ⋅ dl.) (2.7.6)
(Γ ) (S Γ ) (Γ )

Relation (2.7.6) emphasizes the physical significance of the law: the time
variable magnetic field produces (induces) an electric field by the
electromagnetic induction phenomenon. Therefore, the electromagnetic
induction is a physical phenomenon, unlike electric flux density D and
magnetic induction B , which are physical quantities.
Moreover, relation (2.7.6) allows the decomposition of the emf into two
components:
uΓ = ut + um , (2.7.7)
with

∂B  ∂B 
ut = − ∫ ∂
⋅ dA = ∫  −

 ⋅ dA (2.7.8)
Γ ) 
(S Γ ) t (S t

53
called emf induced by transformation and, respectively,

( )
u m = − ∫ B × w ⋅ dl = ∫ (w × B ) ⋅ dl (2.7.9)
(Γ ) (Γ )

called emf induced by movement.


The two components correspond to the two way of electromagnetic
induction phenomenon production: the time variation of the magnetic
induction B at rest (no movement) ( ut is not zero in this case), respectively,
the movement of at least a portion of a closed curve (Γ ) in a magnetic field
in such a way that the field lines are cut away by the field (only in this case
u m is non zero).

We remark that the sum from relation (2.7.7) does not depend on the
reference system chosen for the movement, while the separation into two
components depends in general on the adopted reference system.
The reference system of the emf can be arbitrary chosen. To determine
the real sense of the induced emf, Lenz formulated his famous rule: the real
direction of emf is such that its effects oppose the generating causes.
Therefore, the minus sign from relation (2.7.1) does not have to be
assimilated from mathematical point of view with Lenz’s law, this coming from
the adopted convention when we associate the direction for dl with the direction
of dA according to the corkscrew rule.
For continuity domains the equality:

 ∂B 
∫ (curl E )⋅ dA = − ∫ 


∂t
+ curl(w × B

)
 ⋅ dA (2.7.10)
(S Γ ) (S Γ )

is valid for any surface (S Γ ) and it imposes the equality of the integrands:

curl E = −
∂B
∂t
(
+ curl w × B . ) (2.7.11)

Relation (2.7.11) represents the local form for continuity domains of the
electromagnetic induction law.

( )
For media at rest w = 0 it results the local form

54
∂B
curl E = − (2.7.12)
∂t
and, respectively, the integral form

∂B
uΓ = − ∫ ∂t
⋅ dA. (2.7.13)
(S Γ )

Relation (2.7.12) shows that E is a rotational field and it underlines that


the lines of the electric field produced by electromagnetic induction
phenomenon are closed curves which surrounds the lines of the time
varying magnetic field which generated them.
In stationary regime, the electromagnetic induction law becomes the
theorem of the stationary electric potential, which has the local form

curl E = 0 (2.7.14)
and, respectively, the integral form

uΓ = ∫ E ⋅ dl = 0. (2.7.15)
(Γ )

The local form (2.7.14) of electromagnetic induction law in stationary


regime shows that, in this regime, the vector field E is non-rotational and,
according to the vector identity which states that the curl of the gradient for any
scalar field is zero, it can be introduced the scalar quantity V , called electric
potential, with relation:

E = − grad V . (2.7.16)

Figure 2.7.2

55
The integral form of the electromagnetic induction law in stationary
regime allows the demonstration of the following theorem: the voltage drop
between two points M k and M j from space (arbitrary point) does not
depend on the paths (integration curve) between them.
Indeed (see Figure 2.7.2), let consider two paths (arbitrary) between M k
and M j , represented by the open curves (C1 ) and (C2 ) , whose reunion is the
closed curve (Γ ) .
According to relation (2.7.15) it results that
Mj Mk Mj Mj
uΓ = ∫ E ⋅ dl = ∫ E ⋅ dl + ∫ E ⋅ dl = ∫ E ⋅ dl − ∫ E ⋅ dl = 0 (2.7.17)
(Γ ) Mk Mj Mk Mk
(C1 ) (C 2 ) (C1 ) (C 2 )

that is
Mj Mj

∫ E ⋅ dl = ∫ E ⋅ dl = u M k M j . (2.7.18)
Mk Mk
(C1 ) (C 2 )

The voltage is therefore a scalar physical quantity referring to an


ordered pair consisting of two points from the space, while the electric
potential is a physical scalar quantity associated to each point in the space.
As relation (2.7.16) is a differential type relation, it means that electric
potential is defined up to an arbitrary additive constant. This constant represents
the value V (M 0 ) = VM 0 , arbitrary, of potential M 0 from the space, arbitrary,
considered to be reference value for all potentials. Then, the potential of any
point from the space, for example M j , is

Mj
( )
V M j = VM j = VM 0 − ∫ E ⋅ dl. (2.7.19)
M0

This relation is immediately computed if it is computed the voltage


between points M j and M 0 . For easiness, we use Cartesian coordinate
system:

56
M0 M0
u P j P0 = ∫ E ⋅ dl = − ∫ (grad V ) ⋅ dl =
Mj Mj

( )
M0
 ∂V ∂V ∂V 
=− ∫  ∂x ∂y
 ⋅ i + ⋅ j +
∂z
⋅ k  ⋅ dx ⋅ i + dy ⋅ j + dz ⋅ k =
 (2.7.20)
Mj
M0 M
 ∂V ∂V ∂V  0
Mj
= − ∫  ⋅ dx + ⋅ dy + ⋅ dz  = − ∫ dV = V = VM j − VM 0 ,
Mj  ∂x ∂y ∂z  Mj
M0

which represents exactly relation (2.7.19).


Computing the voltage between two points M k and M j from the space
following a path (C3 ) , arbitrary, but crossing the point M 0 (see Figure 2.7.2),
one obtains:
Mj M0 Mj
uMkM j = ∫ E ⋅ dl = ∫ E ⋅ dl + ∫ E ⋅ dl + (VM 0 − VM 0 = )
Mk Mk M0
(C3 ) (2.7.21)
 Mk   Mj 
= V M 0 − ∫ E ⋅ dl − V M 0 − ∫ E ⋅ dl  = V M k − V M j ,
  
   
 M0   M0 
Relation which represents the expression of the voltage between any two
points function of their potentials.
If one considers VM 0 = 0 , it is said that point M 0 is ”grounded” or
“earthed” and in an electric circuit it has a specific symbol (see Figure 2.7.2).
In these conditions
M0
( )
V M j = VM j = ∫ E ⋅ dl. (2.7.22)
Mj

57
2.8. The magnetic circuit law (Ampère law)

Statement: The magnetomotive force (mmf) umΓ along any closed


curve (Γ ) is equal to the sum between the conduction electric current
strength iSΓ through an open surface (SΓ ) , arbitrary, bordered by the
closed curve (Γ ) and the time derivative of the electric flux ΨS Γ over the
same surface (SΓ ) :

dΨSΓ
u mΓ = iSΓ + . (2.8.1)
dt
The term iSΓ is also called turn and it is denoted by Θ S Γ .

By rewriting relation (2.8.1) one obtains the integral form of the law:
d
∫ H ⋅ dl = ∫ J ⋅ dA + dt ∫ D ⋅ dA (2.8.2)
(Γ ) ( SΓ ) ( SΓ )

which has the above presented form only if the association between the
reference direction of the orientation the oriented line element dl and the
oriented area element dA is according to the corkscrew rule (as for the
electromagnetic induction law).
For moving media, the integration domains follows the bodies in their
movement, following a similar path as the one described for the electromagnetic
induction law, and the integral developed form of magnetic circuit law is
obtained:

∫ H ⋅ dl = ∫ (curl H )⋅ dA = ∫ J ⋅ dA +
(Γ ) ( SΓ ) ( SΓ )
∂D 
+ ∫  (
+ w ⋅ div D + curl D × w  ⋅ dA = )
(SΓ )  ∂t 
(2.8.3)
∂D
= ∫ J ⋅ dA + ∫ ⋅ dA +
(S ) (S ) ∂t
Γ Γ

+ ∫ (w ⋅ div D )⋅ dA + ∫ [curl (D × w)]⋅ dA ,


( SΓ ) ( SΓ )

58
respectively,
u m Γ = i S Γ + id S + iv S + iRS . (2.8.4)
Γ Γ Γ

On the right side of equality (2.8.4) we have four terms:


-the conduction current strength:

i SΓ = ∫ J ⋅ dA; (2.8.5)
( SΓ )

- the displacement current strength:

∂D
id S =
Γ ∫ ∂t ⋅ dA , (2.8.6)
(S )
Γ

where

∂D
= Jd (2.8.7)
∂t
is the displacement current density;

- the convection current:

ivS =
Γ ∫ (w ⋅ div D )⋅ dA = ∫ w ⋅ ρ v ⋅ dA , (2.8.8)
( SΓ ) ( SΓ )

where

w ⋅ ρv = J v (2.8.9)

is the convection current density;

- Roentgen current strength

i RS =
Γ ∫ [curl (D × w)]⋅ dA = ∫ (D × w)⋅ dl . (2.8.10)
( SΓ ) (Γ )

The magnetic circuit law emphasizes two causes that can generate the
magnetic field: conducting bodies transited by conduction currents and/or
time variable electric fields (by displacement currents, convection and
Roentgen currents).
59
For continuity domains, the equality (2.8.3), true for any surface (SΓ ) ,
results in:

curl H = J +
∂D
∂t
( )
+ w ⋅ ρ v + curl D × w , (2.8.11)

relation that represents the local form of the magnetic circuit law.

( )
For stationary media w = 0 the local form becomes:

∂D
curl H = J + (2.8.12)
∂t
and it shows that the closed lines of the magnetic field surrounds the
conductors transited by conduction currents, respectively the lines of the
time variable electric field that generate them.
In steady state regime the magnetic circuit law becomes the Ampère
theorem, and it has the local form:

curl H = J (2.8.13)
and, respectively, the global form
u mΓ = is Γ . (2.8.14)

In the particular cases of the point from the space for which the vector
field H is a non-rotational ( curl H = 0 ), according to the vector identity which
states that the curl of the gradient of any scalar field is zero, the scalar quantity
Vm can be introduced, called scalar magnetic potential, using the relation:

H = − grad Vm . (2.8.15)

2.9. The electric charge conservation law

Statement: The conduction current iΣ emerging from a closed


surface (Σ ) , arbitrary, is equal to the rate of decrease (in time) of the
electric charge qVΣ contained in a domain VΣ bordered by the closed
surface (Σ ) :

60
dqVΣ
iΣ = − . (2.9.1)
dt
By rewriting the relation (2.9.1) one obtains the integral form of the law:
d
∫ J ⋅ dA = ∫ J ⋅ dAext = − dt (V∫ )
ρ v ⋅ dv (2.9.2)
(Σ ) (Σ ) Σ

According to the adopted convention referring to the orientation of the


unit normal to a closed surface (Σ ) (towards its exterior as in Figure 2.9.1), it
results that the reference direction of the current iΣ is towards the outside of
the space domain (VΣ ).
dA = dAext J

(Σ ) dv
ρv
(VΣ )

Figure 2.9.1
For moving media, the time derivative from relation (2.9.2) is a
substantial derivative, which is computed using the relation:

 ∂ρ
d
∫ ρ v ⋅ dv = ∫  v + div w ⋅ ρ v ( ) ⋅ dv, (2.9.3)
dt (V )
Σ (V )  ∂t
Σ

such that the developed integral form of the electric charge conservation law
becomes:

 ∂ρ v
∫ J ⋅ dA = ∫ − ∂t − div w ⋅ ρ v ( ) ⋅ dv, (2.9.4)
(Σ ) (VΣ ) 

or, using Gauss-Ostrogradski relation:

61
 ∂ρ v
∫ (div J )⋅ dv = ∫ − ∂t − div w ⋅ ρ v ( ) ⋅ dv , (2.9.5)
(VΣ ) (VΣ )  

in which

w ⋅ ρv = J v (2.9.6)

is the convection current density.


As relation (2.9.5) is true for any domain (VΣ ) , from the integrands
equality one obtains the local form of the electric charge conservation law for
continuity domains:
∂ρ v
div J = − − div J v (2.9.7)
∂t
or
∂ρ v
(
div J + J v = − ) ∂t
. (2.9.8)

( )
For stationary media w = 0 the local form of the law is:

∂ρv
div J = − . (2.9.9)
∂t
In steady state regime the local form becomes:

div J = 0 , (2.9.10)

The global form has the expression


iΣ = 0 (2.9.11)
and it shows that, in this regime, for any closed surface (Σ ) the electric
conduction current is preserved.
The electric charge conservation law underlines from macroscopic point
of view that the electric convection current is generated by moving bodies.

62
2.10. The constitutive law of electric conduction

Statement: In any point and at any time moment, the electric


conduction current density J depends on the electric field strength E :

J=J E .( ) (2.10.1)

Function of the nature of the various conducting materials, the


relationship (2.10.1) has explicit forms. As follows, there will be studied only
the conducting linear and isotropic materials. In this case the dependence
(2.10.1) becomes:

(
J = σ ⋅ E + Ei , ) (2.10.2)

in which σ is a dimensional material quantity called electric conductivity, and


E i represents impressed electric field strength.
Despite it has the attribute electric, the impressed electric field in not an
electric field in the accepted notion as it was presented up to now.
For understand it, it is very useful to resort to the microscopic
configuration of the conductors, noticing that upon the free charge carriers, in
certain conditions (if, for example, the conductors are heterogeneous, or they
are accelerated, or they are not isotherms), beside the electric nature forces,
there are also acting non-electric forces.
Admitting as a simplified model of the microscopic structure of a
homogenous, not-accelerated and isotherm conductor an ionic network positive
charged, placed in an electronic “fluid” negatively charged and without ordered
component for its constitutive particles’ movement, the volume electric charge
density of the conductors is zero.
Supposing that upon this conductor there is applied a non electric
perturbation, it has a new regime: due to non electric forces, the electrons will
migrate and they will generate in this way a separation of the opposite sign
charges, showing an electric field strength E , oriented from the region with
excess positive charges towards the region with excess of negative charges.
Upon each electron (having the charge q0 ) act two types of forces: an electric
nature one:

F el = q0 ⋅ E (2.10.3)

63
opposite direction to the electric field strength E (because q0 < 0 ) and a non
electric nature one F neel having the direction of the field. The electron will
move under the action of the non zero resulting force

 F neel 
F el + F neel = q0 ⋅ E + F neel = q0 ⋅  E +  ≠ 0 . (2.10.4)
 q 0 

The term

F neel
= Ei (2.10.5)
q0
has the dimension of an electric field strength and it is named impressed
electric field strength. Impressed electric field is a quantity that captures
from the electric point of view the non electric nature actions upon the
charge carriers from the conductors, in their heterogeneous areas of the
chemical-physical properties.
Function on the cause nature that generate them, the impressed electric
fields can be of mechanical, thermal, chemical, etc. nature, and function of
space repartition of heterogeneities, the impressed electric fields can be
surface or interface (contact) ones.
Relation (2.10.2) can be also written under the form:

E + Ei = ρ ⋅ J, (2.10.6)

in which is emphasized the dimensional material quantity called electric


resistivity

ρ = σ −1. (2.10.7)

The temperature dependence of the resistivity is well approximated by the


linear relation:
ρ (θ ) = ρ (θ 0 ) ⋅ [1 + α ⋅ (θ − θ 0 )], (2.10.8)

where θ 0 is a reference temperature, and α represents the resistivity


temperature variation coefficient.
For the majority of the metals α > 0 (resistivity increases with
temperature), but there are other conducting materials, such as coal for example,
for which α < 0 (their resistivity decreases while the temperature increases).

64
In the table below there are presented the values of the resistivity ρ ad of
the resistivity temperature variation coefficient α for a few of conducting
materials.

Name of the ρ at 200 C α ⋅ 103 between


material
[ Ω ⋅ mm2 / m ] 0 0 C and 1000 C
Silver 0.0161 4
Technical Cooper 0.0175 4.45
Electrolytic Cooper 0.0175 4.4
Gold 0.0237 3.77
Aluminum 0.0278 4.23
Yellow Cooper 0.08 1.5
Platinum 0.0866 2.47
Iron 0.0918 6.25
Nickel 0.138 6.21
Plumb 0.221 4.11
Cooper Nickel 0.4 0.2
Manganin 0.43 0.6
Constantan 0.45 0.4
Retort carbon 7.025 -0.3
Carbon 40 ÷ 100 -0.2
Uranium Oxide 50 ÷ 1000 -1.5
Copper Oxide 0.05 ÷ 1000 -2.6

Function of resistivity values, respectively of conductivity, the materials can


be classified into three big categories, namely:
• insulators, having big values for resistivity (respectively vary small values
for conductivity); a perfect insulator is the material for which σ = 0 ;

65
• semiconductors, with average values for resistivity (respectively for
conductivity) and with exponential shape for their temperature dependence;
• conductors, with very small values for resistivity (respectively very high
values for conductivity); a perfect conductor is the material for which
ρ = 0.

The physical significance of the electric conduction low is to underline as


a cause for the generation of the electric field the bodies that possess
impressed electric field.
For all constitutive lows, the local forms are those that give the consistent
information. The constitutive law of electric conduction is not an exception from
this rule. The integral form of this law is very useful for electric circuits study
and it will be presented for the particular case of the conductors having
sufficiently small cross sections in order to consider that the repartitions of the
conduction electric currents through the conductors’ sections are uniform.
These conductors are called filiform.

R M2

uf ub
e

M1

Figure 2.10.1, a Figure 2.10.1, b

Let consider a section from such a filiform conductor, having a constant


A cross section (see Figure 2.10.1, a), through which flows a current i . The
module of the current density J will be then
i
J= . (2.10.9)
A

66
Integrating the local form of the constitutive law of the electric conduction
along the axis (C ) of the conducting section, between the extremities M 1 and
M 2 of this section, one obtains:

∫ (E + E i )⋅ dl = ∫ E ⋅ dl + ∫ E i ⋅ dl = ∫ (ρ ⋅ J )⋅ dl =
M2 M2 M2 M2

M1 M1 M1 M1
(C ) (C ) (C ) (C )
M2 M2 M2
(2.10.10)
i dl
= ∫ (ρ ⋅ J ) ⋅ dl = ∫ ρ ⋅ ⋅ dl = i ⋅ ∫ ρ ⋅
A A
M1 M1 M 1
(C ) (C ) (C )

because the vectors J and dl are omo-parallel, and the electric conduction
current strength is the same along the conductor. In this relationship the
following quantities are underlined:
• electric voltage along the conductor (line?):
M2
uf = ∫ E ⋅ dl ; (2.10.11)
M1
(C )

• impressed electromotive force


M2
uei = ei = e = ∫ E i ⋅ dl ; (2.10.12)
M1
(C )

• electric resistance of the conducting section


M2
dl
R= ∫ ρ⋅
A
.
(2.10.13)
M1
(C )

So, the integral form of the constitutive law of electric conduction for
filiform conductors is:
u f + e = R ⋅i (2.10.14)

and it has the following remarks.

67
▪ The reference directions for global quantities u f , e and i that appear
in relation (2.10.14) are associated using the convention depicted in the
graphical representation from Figure 2.10.1,b. Because in general the
reference directions are arbitrary chosen, they can be optionally chosen; if one
uses other reference directions that the one presented in Figure 2.10.1,b, the
integral form of the constitutive law of electric conduction changes by changing
the corresponding signs of the terms with changed reference directions.
▪ In the particular case of the steady state regime, one proved that the voltage
does not depend on the path and, as a consequence, the voltage along the path
u f is equal to the voltage ub computed along any open curve (Cb ) having as
extremities the point M 1 and M 2 (called terminals in the electric circuit
theory); the electric voltage
M2
ub = ∫ E ⋅ dl (2.10.15)
M1
(Cb )

is called the terminals voltage, and the integral form of the constitutive law of
electric conduction becomes

ub + e = R ⋅ i . (2.10.16)

2.11. The law of energy conversion associated to electric


conduction

Statement: The power volume density p of the power transferred


from electromagnetic field to substance in the electric conduction process is
equal, at any moment of time, to the dot product between the electric field
strength E and the conduction electric current density J :
p = E ⋅ J. (2.11.1)

We remark that the sign of the dot product shows the real direction of
the transferred power: from the field toward the conducting body if p > 0 ,
respectively from the conducting body towards the field if p < 0 .

68
From relation (2.10.6) and (2.11.1) one can obtain a new expression of the
local form of the law, for linear and isotropic conductors:

( )
p = ρ ⋅ J − Ei ⋅ J = ρ ⋅ J − E i ⋅ J .
2
(2.11.2)

The two terms from the right side of equation (2.11.2) have the following
interpretation:
2
• Term ρ ⋅ J is positive and it represents the power volume density
irreversibly transferred by the electromagnetic field of the conductor.
• Term E i ⋅ J has the direction imposed by the orientation of the two vectors
E i and J ; if γ is the smallest angle between the two vectors, then we can
have the following situations: γ ∈ [0;π 2 ) , so E i ⋅ J > 0 and the power is
transferred from the impressed electric field sources (of the conductor) to
( ]
the electromagnetic field or γ ∈ π ;π , so E i ⋅ J < 0 and the power is
2
transferred from the electromagnetic field to impressed electric field
sources (of the conductor) or γ = π 2 and there is no power transfer
between the electromagnetic field and the conductor.

The integral form of the law is obtained by first integrating the local form
(2.11.1) on domain (VΣ ) in which the electric conduction process takes place.
The total power transferred by the electromagnetic field to the carrying current
conductors from (VΣ ) is

P= ∫ p ⋅ dv = ∫ (E ⋅ J ) ⋅ dv, (2.11.3)
(VΣ ) (VΣ )

and the energy corresponding to this phenomenon, between two time moments
t1 and t 2 , is
t2 t2  
(
W = ∫ P ⋅ dt = ∫  ∫ E ⋅ J ⋅ dv  ⋅ dt . ) (2.11.4)
t1 t1 
(VΣ ) 

The measure units I.S. for the quantities p , P and W are, respectively,
Watt per cubic meter ( W / m 3 ), Watt ( W ) and Watt multiplied by
second (W ⋅ s ).
For a filiform conductor in steady state regime (see Figure 2.10.1,a),
taking into account that dv = A ⋅ dl , it results that the integral relation (2.11.3)
takes the particular form

69
P= ∫ (E ⋅ J ) ⋅ dv = ∫ E ⋅ J ⋅ dv = ∫ E ⋅ J ⋅ A ⋅ dl =
(VΣ ) (VΣ ) (VΣ )
(2.11.5)
=i⋅ ∫ E ⋅ dl = u f ⋅ i = ub ⋅ i = Pb ,
(C )

that shows that the power transferred by the electromagnetic field to the
conducting section in the electric conduction process is equal to the power Pb
called received power at terminals.
Regarding the physical significance, the law characterizes the thermal
effect of the electromagnetic field, referring to heat producing phenomenon
(Joule-Lenz effect) which is associated to the electric conduction current
flow through the conductors, that corresponds to the transformation of
electromagnetic energy into caloric energy.

2.12. The electrolysis law

Statement: The substance mass m deposited at one of the electrodes of


an electrolytic bath in a time interval [t1 ;t 2 ] depends on the electric
current i (t ) flowing through the bath according to relationship:
t
1 A 2
⋅ ⋅ i (t ) ⋅ dt ,
F0 v t∫
m= (2.12.1)
1

where F0 is a universal constant called Faraday’s constant, that has the value

C
F0 = 96,484.6 , (2.12.2)
ech.gram

A is the mol’s mass of the deposited substance, and v is valence of a ion from
the deposited substance.
The ratio A / v is called the chemical equivalent, and its values are
presented in the table below.

70
Atomic mass Valence Chemical
Substance equivalent A / v [g ]
A[g ] v
Aluminum 1 1 1
Oxygen 16 2 8
Aluminum 27 3 9
Magnesium 24 2 12
Iron 3 55.9 3 18.6
Nickel 58.6 2 29.3
Cooper 2 63.2 2 31.6
Zinc 68.8 2 34.4
Potassium 39 1 39
tin 117.4 2 58.7
Cooper 1 63.2 1 63.2
Gold 196.1 3 65.4
Platinum 194.4 2 97.2
Mercury 199.8 2 99.9
Plumb 206.4 2 103.2
Silver 107.7 1 107.7

The physical significance of the electrolysis law is that it shows the mass
transport phenomenon associated to electric current flow through certain
substances.
The electrolysis process has many practical applications such that electro-
metallurgy (metal extraction from ore, metal refining, substance preparation
etc.) and galvanotechnics (galvanostegy – the coating of some object by nickel
plating, chrome plating, cadmium plating a.s.o and galvanoplasty – object
shape reproduction).

71

You might also like