Mag Field
Mag Field
Mag Field
the magnetic influence on moving electric charges, electric currents,[2]: ch1 [3]
and magnetic materials. A moving charge in a magnetic field experiences a force
perpendicular to its own velocity and to the magnetic field.[2]: ch13 [4]: 278 A
permanent magnet's magnetic field pulls on ferromagnetic materials such as iron,
and attracts or repels other magnets. In addition, a nonuniform magnetic field
exerts minuscule forces on "nonmagnetic" materials by three other magnetic effects:
paramagnetism, diamagnetism, and antiferromagnetism, although these forces are
usually so small they can only be detected by laboratory equipment. Magnetic fields
surround magnetized materials, electric currents, and electric fields varying in
time. Since both strength and direction of a magnetic field may vary with location,
it is described mathematically by a function assigning a vector to each point of
space, called a vector field (more precisely, a pseudovector field).
In electromagnetics, the term magnetic field is used for two distinct but closely
related vector fields denoted by the symbols B and H. In the International System
of Units, the unit of B, magnetic flux density, is the tesla (in SI base units:
kilogram per second2 per ampere),[5]: 21 which is equivalent to newton per meter
per ampere. The unit of H, magnetic field strength, is ampere per meter (A/m).[5]:
22 B and H differ in how they take the medium and/or magnetization into account.
In vacuum, the two fields are related through the vacuum permeability,
𝐵
/
𝜇
0
=
𝐻
{\displaystyle \mathbf {B} /\mu _{0}=\mathbf {H} }; in a magnetized material, the
quantities on each side of this equation differ by the magnetization field of the
material.
Magnetic fields are produced by moving electric charges and the intrinsic magnetic
moments of elementary particles associated with a fundamental quantum property,
their spin.[6][2]: ch1 Magnetic fields and electric fields are interrelated and are
both components of the electromagnetic force, one of the four fundamental forces of
nature.
Description
The force on an electric charge depends on its location, speed, and direction; two
vector fields are used to describe this force.[2]: ch1 The first is the electric
field, which describes the force acting on a stationary charge and gives the
component of the force that is independent of motion. The magnetic field, in
contrast, describes the component of the force that is proportional to both the
speed and direction of charged particles.[2]: ch13 The field is defined by the
Lorentz force law and is, at each instant, perpendicular to both the motion of the
charge and the force it experiences.
There are two different, but closely related vector fields which are both sometimes
called the "magnetic field" written B and H.[note 1] While both the best names for
these fields and exact interpretation of what these fields represent has been the
subject of long running debate, there is wide agreement about how the underlying
physics work.[7] Historically, the term "magnetic field" was reserved for H while
using other terms for B, but many recent textbooks use the term "magnetic field" to
describe B as well as or in place of H.[note 2] There are many alternative names
for both (see sidebars).
The B-field
Finding the magnetic force
A charged particle that is moving with velocity v in a magnetic field B will feel a
magnetic force F. Since the magnetic force always pulls sideways to the direction
of motion, the particle moves in a circle.
Since these three vectors are related to each other by a cross product, the
direction of this force can be found using the right hand rule.
Alternative names for B[8]
Magnetic flux density[5]: 138
Magnetic induction[9]
Magnetic field (ambiguous)
The magnetic field vector B at any point can be defined as the vector that, when
used in the Lorentz force law, correctly predicts the force on a charged particle
at that point:[10][11]: 204
Here F is the force on the particle, q is the particle's electric charge, v, is the
particle's velocity, and × denotes the cross product. The direction of force on the
charge can be determined by a mnemonic known as the right-hand rule (see the
figure).[note 3] Using the right hand, pointing the thumb in the direction of the
current, and the fingers in the direction of the magnetic field, the resulting
force on the charge points outwards from the palm. The force on a negatively
charged particle is in the opposite direction. If both the speed and the charge are
reversed then the direction of the force remains the same. For that reason a
magnetic field measurement (by itself) cannot distinguish whether there is a
positive charge moving to the right or a negative charge moving to the left. (Both
of these cases produce the same current.) On the other hand, a magnetic field
combined with an electric field can distinguish between these, see Hall effect
below.
The first term in the Lorentz equation is from the theory of electrostatics, and
says that a particle of charge q in an electric field E experiences an electric
force:
𝐹
electric
=
𝑞
𝐸
.
{\displaystyle \mathbf {F} _{\text{electric}}=q\mathbf {E} .}
The second term is the magnetic force:[11]
𝐹
magnetic
=
𝑞
(
𝑣
×
𝐵
)
.
{\displaystyle \mathbf {F} _{\text{magnetic}}=q(\mathbf {v} \times \mathbf {B} ).}
Using the definition of the cross product, the magnetic force can also be written
as a scalar equation:[10]: 357
𝐹
magnetic
=
𝑞
𝑣
𝐵
sin
(
𝜃
)
{\displaystyle F_{\text{magnetic}}=qvB\sin(\theta )}
where Fmagnetic, v, and B are the scalar magnitude of their respective vectors, and
θ is the angle between the velocity of the particle and the magnetic field. The
vector B is defined as the vector field necessary to make the Lorentz force law
correctly describe the motion of a charged particle. In other words,[10]: 173–4
[T]he command, "Measure the direction and magnitude of the vector B at such and
such a place," calls for the following operations: Take a particle of known charge
q. Measure the force on q at rest, to determine E. Then measure the force on the
particle when its velocity is v; repeat with v in some other direction. Now find a
B that makes the Lorentz force law fit all these results—that is the magnetic field
at the place in question.
The B field can also be defined by the torque on a magnetic dipole, m.[12]: 174
The H-field
Alternative names for H[8]
Magnetic field intensity[9]
Magnetic field strength[5]: 139
Magnetic field
Magnetizing field
Auxiliary magnetic field
The magnetic H field is defined:[11]: 269 [12]: 192 [2]: ch36
where
𝜇
0
{\displaystyle \mu _{0}} is the vacuum permeability, and M is the magnetization
vector. In a vacuum, B and H are proportional to each other. Inside a material they
are different (see H and B inside and outside magnetic materials). The SI unit of
the H-field is the ampere per metre (A/m),[15] and the CGS unit is the oersted
(Oe).[13][10]: 286
Measurement
Main article: Magnetometer
An instrument used to measure the local magnetic field is known as a magnetometer.
Important classes of magnetometers include using induction magnetometers (or
search-coil magnetometers) which measure only varying magnetic fields, rotating
coil magnetometers, Hall effect magnetometers, NMR magnetometers, SQUID
magnetometers, and fluxgate magnetometers. The magnetic fields of distant
astronomical objects are measured through their effects on local charged particles.
For instance, electrons spiraling around a field line produce synchrotron radiation
that is detectable in radio waves. The finest precision for a magnetic field
measurement was attained by Gravity Probe B at 5 aT (5×10−18 T).[16]
Visualization
Main article: Field line
Visualizing magnetic fields
Left: the direction of magnetic field lines represented by iron filings sprinkled
on paper placed above a bar magnet.
Right: compass needles point in the direction of the local magnetic field, towards
a magnet's south pole and away from its north pole.
The field can be visualized by a set of magnetic field lines, that follow the
direction of the field at each point. The lines can be constructed by measuring the
strength and direction of the magnetic field at a large number of points (or at
every point in space). Then, mark each location with an arrow (called a vector)
pointing in the direction of the local magnetic field with its magnitude
proportional to the strength of the magnetic field. Connecting these arrows then
forms a set of magnetic field lines. The direction of the magnetic field at any
point is parallel to the direction of nearby field lines, and the local density of
field lines can be made proportional to its strength. Magnetic field lines are like
streamlines in fluid flow, in that they represent a continuous distribution, and a
different resolution would show more or fewer lines.
An advantage of using magnetic field lines as a representation is that many laws of
magnetism (and electromagnetism) can be stated completely and concisely using
simple concepts such as the "number" of field lines through a surface. These
concepts can be quickly "translated" to their mathematical form. For example, the
number of field lines through a given surface is the surface integral of the
magnetic field.[10]: 237
Various phenomena "display" magnetic field lines as though the field lines were
physical phenomena. For example, iron filings placed in a magnetic field form lines
that correspond to "field lines".[note 5] Magnetic field "lines" are also visually
displayed in polar auroras, in which plasma particle dipole interactions create
visible streaks of light that line up with the local direction of Earth's magnetic
field.
The magnetic field of permanent magnets can be quite complicated, especially near
the magnet. The magnetic field of a small[note 6] straight magnet is proportional
to the magnet's strength (called its magnetic dipole moment m). The equations are
non-trivial and depend on the distance from the magnet and the orientation of the
magnet. For simple magnets, m points in the direction of a line drawn from the
south to the north pole of the magnet. Flipping a bar magnet is equivalent to
rotating its m by 180 degrees.
There are two simplified models for the nature of these dipoles: the magnetic pole
model and the Amperian loop model. These two models produce two different magnetic
fields, H and B. Outside a material, though, the two are identical (to a
multiplicative constant) so that in many cases the distinction can be ignored. This
is particularly true for magnetic fields, such as those due to electric currents,
that are not generated by magnetic materials.
The magnetic pole model: two opposing poles, North (+) and South (−), separated by
a distance d produce a H-field (lines).
Historically, early physics textbooks would model the force and torques between two
magnets as due to magnetic poles repelling or attracting each other in the same
manner as the Coulomb force between electric charges. At the microscopic level,
this model contradicts the experimental evidence, and the pole model of magnetism
is no longer the typical way to introduce the concept.[11]: 258 However, it is
still sometimes used as a macroscopic model for ferromagnetism due to its
mathematical simplicity.[17]
In the magnetic pole model, the elementary magnetic dipole m is formed by two
opposite magnetic poles of pole strength qm separated by a small distance vector d,
such that m = qm d. The magnetic pole model predicts correctly the field H both
inside and outside magnetic materials, in particular the fact that H is opposite to
the magnetization field M inside a permanent magnet.
Since it is based on the fictitious idea of a magnetic charge density, the pole
model has limitations. Magnetic poles cannot exist apart from each other as
electric charges can, but always come in north–south pairs. If a magnetized object
is divided in half, a new pole appears on the surface of each piece, so each has a
pair of complementary poles. The magnetic pole model does not account for magnetism
that is produced by electric currents, nor the inherent connection between angular
momentum and magnetism.
The pole model usually treats magnetic charge as a mathematical abstraction, rather
than a physical property of particles. However, a magnetic monopole is a
hypothetical particle (or class of particles) that physically has only one magnetic
pole (either a north pole or a south pole). In other words, it would possess a
"magnetic charge" analogous to an electric charge. Magnetic field lines would start
or end on magnetic monopoles, so if they exist, they would give exceptions to the
rule that magnetic field lines neither start nor end. Some theories (such as Grand
Unified Theories) have predicted the existence of magnetic monopoles, but so far,
none have been observed.
A current loop (ring) that goes into the page at the x and comes out at the dot
produces a B-field (lines). As the radius of the current loop shrinks, the fields
produced become identical to an abstract "magnetostatic dipole" (represented by an
arrow pointing to the right).
In the model developed by Ampere, the elementary magnetic dipole that makes up all
magnets is a sufficiently small Amperian loop with current I and loop area A. The
dipole moment of this loop is m = IA.
The magnetic field of a magnetic dipole is depicted in the figure. From outside,
the ideal magnetic dipole is identical to that of an ideal electric dipole of the
same strength. Unlike the electric dipole, a magnetic dipole is properly modeled as
a current loop having a current I and an area a. Such a current loop has a magnetic
moment of
𝑚
=
𝐼
𝑎
,
{\displaystyle m=Ia,}
where the direction of m is perpendicular to the area of the loop and depends on
the direction of the current using the right-hand rule. An ideal magnetic dipole is
modeled as a real magnetic dipole whose area a has been reduced to zero and its
current I increased to infinity such that the product m = Ia is finite. This model
clarifies the connection between angular momentum and magnetic moment, which is the
basis of the Einstein–de Haas effect rotation by magnetization and its inverse, the
Barnett effect or magnetization by rotation.[18] Rotating the loop faster (in the
same direction) increases the current and therefore the magnetic moment, for
example.
Interactions with magnets
Force between magnets
Main article: Force between magnets
Specifying the force between two small magnets is quite complicated because it
depends on the strength and orientation of both magnets and their distance and
direction relative to each other. The force is particularly sensitive to rotations
of the magnets due to magnetic torque. The force on each magnet depends on its
magnetic moment and the magnetic field[note 7] of the other.
To understand the force between magnets, it is useful to examine the magnetic pole
model given above. In this model, the H-field of one magnet pushes and pulls on
both poles of a second magnet. If this H-field is the same at both poles of the
second magnet then there is no net force on that magnet since the force is opposite
for opposite poles. If, however, the magnetic field of the first magnet is
nonuniform (such as the H near one of its poles), each pole of the second magnet
sees a different field and is subject to a different force. This difference in the
two forces moves the magnet in the direction of increasing magnetic field and may
also cause a net torque.
This is a specific example of a general rule that magnets are attracted (or
repulsed depending on the orientation of the magnet) into regions of higher
magnetic field. Any non-uniform magnetic field, whether caused by permanent magnets
or electric currents, exerts a force on a small magnet in this way.
The details of the Amperian loop model are different and more complicated but yield
the same result: that magnetic dipoles are attracted/repelled into regions of
higher magnetic field. Mathematically, the force on a small magnet having a
magnetic moment m due to a magnetic field B is:[19]: Eq. 11.42
𝐹
=
𝛻
(
𝑚
⋅
𝐵
)
,
{\displaystyle \mathbf {F} ={\boldsymbol {\nabla }}\left(\mathbf {m} \cdot \mathbf
{B} \right),}
where the gradient ∇ is the change of the quantity m · B per unit distance and the
direction is that of maximum increase of m · B. The dot product m · B = mBcos(θ),
where m and B represent the magnitude of the m and B vectors and θ is the angle
between them. If m is in the same direction as B then the dot product is positive
and the gradient points "uphill" pulling the magnet into regions of higher B-field
(more strictly larger m · B). This equation is strictly only valid for magnets of
zero size, but is often a good approximation for not too large magnets. The
magnetic force on larger magnets is determined by dividing them into smaller
regions each having their own m then summing up the forces on each of these very
small regions.
Torque on a dipole
In the pole model of a dipole, an H field (to right) causes equal but opposite
forces on a N pole (+q) and a S pole (−q) creating a torque.
Equivalently, a B field induces the same torque on a current loop with the same
magnetic dipole moment.
In terms of the pole model, two equal and opposite magnetic charges experiencing
the same H also experience equal and opposite forces. Since these equal and
opposite forces are in different locations, this produces a torque proportional to
the distance (perpendicular to the force) between them. With the definition of m as
the pole strength times the distance between the poles, this leads to τ = μ0 m H
sin θ, where μ0 is a constant called the vacuum permeability, measuring 4π×10−7
V·s/(A·m) and θ is the angle between H and m.
𝜏
=
𝑚
×
𝐵
=
𝜇
0
𝑚
×
𝐻
,
{\displaystyle {\boldsymbol {\tau }}=\mathbf {m} \times \mathbf {B} =\mu _{0}\
mathbf {m} \times \mathbf {H} ,\,}
where × represents the vector cross product. This equation includes all of the
qualitative information included above. There is no torque on a magnet if m is in
the same direction as the magnetic field, since the cross product is zero for two
vectors that are in the same direction. Further, all other orientations feel a
torque that twists them toward the direction of magnetic field.
Right hand grip rule: a current flowing in the direction of the white arrow
produces a magnetic field shown by the red arrows.
All moving charged particles produce magnetic fields. Moving point charges, such as
electrons, produce complicated but well known magnetic fields that depend on the
charge, velocity, and acceleration of the particles.[20]
𝐵
=
𝜇
0
𝐼
4
𝜋
∫
w
i
r
e
d
ℓ
×
𝑟
^
𝑟
2
,
{\displaystyle \mathbf {B} ={\frac {\mu _{0}I}{4\pi }}\int _{\mathrm {wire} }{\frac
{\mathrm {d} {\boldsymbol {\ell }}\times \mathbf {\hat {r}} }{r^{2}}},}
where the integral sums over the wire length where vector dℓ is the vector line
element with direction in the same sense as the current I, μ0 is the magnetic
constant, r is the distance between the location of dℓ and the location where the
magnetic field is calculated, and r̂ is a unit vector in the direction of r. For
example, in the case of a sufficiently long, straight wire, this becomes:
|
𝐵
|
=
𝜇
0
2
𝜋
𝑟
𝐼
{\displaystyle |\mathbf {B} |={\frac {\mu _{0}}{2\pi r}}I}
where r = |r|. The direction is tangent to a circle perpendicular to the wire
according to the right hand rule.[21]: 225
A slightly more general[22][note 9] way of relating the current
𝐼
{\displaystyle I} to the B-field is through Ampère's law:
∮
𝐵
⋅
d
ℓ
=
𝜇
0
𝐼
e
n
c
,
{\displaystyle \oint \mathbf {B} \cdot \mathrm {d} {\boldsymbol {\ell }}=\mu
_{0}I_{\mathrm {enc} },}
where the line integral is over any arbitrary loop and
𝐼
enc
{\displaystyle I_{\text{enc}}} is the current enclosed by that loop. Ampère's law
is always valid for steady currents and can be used to calculate the B-field for
certain highly symmetric situations such as an infinite wire or an infinite
solenoid.
In a modified form that accounts for time varying electric fields, Ampère's law is
one of four Maxwell's equations that describe electricity and magnetism.
𝐹
=
𝑞
𝐸
+
𝑞
𝑣
×
𝐵
,
{\displaystyle \mathbf {F} =q\mathbf {E} +q\mathbf {v} \times \mathbf {B} ,}
where F is the force, q is the electric charge of the particle, v is the
instantaneous velocity of the particle, and B is the magnetic field (in teslas).
The Lorentz force is always perpendicular to both the velocity of the particle and
the magnetic field that created it. When a charged particle moves in a static
magnetic field, it traces a helical path in which the helix axis is parallel to the
magnetic field, and in which the speed of the particle remains constant. Because
the magnetic force is always perpendicular to the motion, the magnetic field can do
no work on an isolated charge.[23][24] It can only do work indirectly, via the
electric field generated by a changing magnetic field. It is often claimed that the
magnetic force can do work to a non-elementary magnetic dipole, or to charged
particles whose motion is constrained by other forces, but this is incorrect[25]
because the work in those cases is performed by the electric forces of the charges
deflected by the magnetic field.
𝐹
=
𝑞
𝑣
𝐵
sin
𝜃
,
{\displaystyle F=qvB\sin \theta ,}
so, for N charges where
𝑁
=
𝑛
ℓ
𝐴
,
{\displaystyle N=n\ell A,}
the force exerted on the conductor is
𝑓
=
𝐹
𝑁
=
𝑞
𝑣
𝐵
𝑛
ℓ
𝐴
sin
𝜃
=
𝐵
𝑖
ℓ
sin
𝜃
,
{\displaystyle f=FN=qvBn\ell A\sin \theta =Bi\ell \sin \theta ,}
where i = nqvA.
Relation between H and B
The formulas derived for the magnetic field above are correct when dealing with the
entire current. A magnetic material placed inside a magnetic field, though,
generates its own bound current, which can be a challenge to calculate. (This bound
current is due to the sum of atomic sized current loops and the spin of the
subatomic particles such as electrons that make up the material.) The H-field as
defined above helps factor out this bound current; but to see how, it helps to
introduce the concept of magnetization first.
Magnetization
Main article: Magnetization
The magnetization vector field M represents how strongly a region of material is
magnetized. It is defined as the net magnetic dipole moment per unit volume of that
region. The magnetization of a uniform magnet is therefore a material constant,
equal to the magnetic moment m of the magnet divided by its volume. Since the SI
unit of magnetic moment is A⋅m2, the SI unit of magnetization M is ampere per
meter, identical to that of the H-field.
In the Amperian loop model, the magnetization is due to combining many tiny
Amperian loops to form a resultant current called bound current. This bound
current, then, is the source of the magnetic B field due to the magnet. Given the
definition of the magnetic dipole, the magnetization field follows a similar law to
that of Ampere's law:[26]
∮
𝑀
⋅
d
ℓ
=
𝐼
b
,
{\displaystyle \oint \mathbf {M} \cdot \mathrm {d} {\boldsymbol {\ell }}=I_{\mathrm
{b} },}
where the integral is a line integral over any closed loop and Ib is the bound
current enclosed by that closed loop.
In the magnetic pole model, magnetization begins at and ends at magnetic poles. If
a given region, therefore, has a net positive "magnetic pole strength"
(corresponding to a north pole) then it has more magnetization field lines entering
it than leaving it. Mathematically this is equivalent to:
∮
𝑆
𝜇
0
𝑀
⋅
d
𝐴
=
−
𝑞
M
,
{\displaystyle \oint _{S}\mu _{0}\mathbf {M} \cdot \mathrm {d} \mathbf {A} =-q_{\
mathrm {M} },}
where the integral is a closed surface integral over the closed surface S and qM is
the "magnetic charge" (in units of magnetic flux) enclosed by S. (A closed surface
completely surrounds a region with no holes to let any field lines escape.) The
negative sign occurs because the magnetization field moves from south to north.
H-field and magnetic materials
𝐵
𝜇
0
−
𝑀
.
{\displaystyle \mathbf {H} \ \equiv \ {\frac {\mathbf {B} }{\mu _{0}}}-\mathbf
{M} .}
In terms of the H-field, Ampere's law is
∮
𝐻
⋅
d
ℓ
=
∮
(
𝐵
𝜇
0
−
𝑀
)
⋅
d
ℓ
=
𝐼
t
o
t
−
𝐼
b
=
𝐼
f
,
{\displaystyle \oint \mathbf {H} \cdot \mathrm {d} {\boldsymbol {\ell }}=\oint \
left({\frac {\mathbf {B} }{\mu _{0}}}-\mathbf {M} \right)\cdot \mathrm {d} {\
boldsymbol {\ell }}=I_{\mathrm {tot} }-I_{\mathrm {b} }=I_{\mathrm {f} },}
where If represents the 'free current' enclosed by the loop so that the line
integral of H does not depend at all on the bound currents.[27]
For the differential equivalent of this equation see Maxwell's equations. Ampere's
law leads to the boundary condition
(
𝐻
1
∥
−
𝐻
2
∥
)
=
𝐾
f
×
𝑛
^
,
{\displaystyle \left(\mathbf {H_{1}^{\parallel }} -\mathbf {H_{2}^{\parallel }} \
right)=\mathbf {K} _{\mathrm {f} }\times {\hat {\mathbf {n} }},}
where Kf is the surface free current density and the unit normal
𝑛
^{\displaystyle {\hat {\mathbf {n} }}} points in the direction from medium 2 to
medium 1.[28]
Similarly, a surface integral of H over any closed surface is independent of the
free currents and picks out the "magnetic charges" within that closed surface:
∮
𝑆
𝜇
0
𝐻
⋅
d
𝐴
=
∮
𝑆
(
𝐵
−
𝜇
0
𝑀
)
⋅
d
𝐴
=
0
−
(
−
𝑞
M
)
=
𝑞
M
,
{\displaystyle \oint _{S}\mu _{0}\mathbf {H} \cdot \mathrm {d} \mathbf {A} =\oint
_{S}(\mathbf {B} -\mu _{0}\mathbf {M} )\cdot \mathrm {d} \mathbf {A} =0-(-q_{\
mathrm {M} })=q_{\mathrm {M} },}
which does not depend on the free currents.
The H-field, therefore, can be separated into two[note 10] independent parts:
𝐻
=
𝐻
0
+
𝐻
d
,
{\displaystyle \mathbf {H} =\mathbf {H} _{0}+\mathbf {H} _{\mathrm {d} },}
where H0 is the applied magnetic field due only to the free currents and Hd is the
demagnetizing field due only to the bound currents.
The magnetic H-field, therefore, re-factors the bound current in terms of "magnetic
charges". The H field lines loop only around "free current" and, unlike the
magnetic B field, begins and ends near magnetic poles as well.
Magnetism
Main article: Magnetism
Most materials respond to an applied B-field by producing their own magnetization M
and therefore their own B-fields. Typically, the response is weak and exists only
when the magnetic field is applied. The term magnetism describes how materials
respond on the microscopic level to an applied magnetic field and is used to
categorize the magnetic phase of a material. Materials are divided into groups
based upon their magnetic behavior:
𝐵
=
𝜇
𝐻
,
{\displaystyle \mathbf {B} =\mu \mathbf {H} ,}
where μ is a material dependent parameter called the permeability. In some cases
the permeability may be a second rank tensor so that H may not point in the same
direction as B. These relations between B and H are examples of constitutive
equations. However, superconductors and ferromagnets have a more complex B-to-H
relation; see magnetic hysteresis.
Stored energy
Main article: Magnetic energy
See also: Magnetic hysteresis
Energy is needed to generate a magnetic field both to work against the electric
field that a changing magnetic field creates and to change the magnetization of any
material within the magnetic field. For non-dispersive materials, this same energy
is released when the magnetic field is destroyed so that the energy can be modeled
as being stored in the magnetic field.
𝑢
=
𝐵
⋅
𝐻
2
=
𝐵
⋅
𝐵
2
𝜇
=
𝜇
𝐻
⋅
𝐻
2
.
{\displaystyle u={\frac {\mathbf {B} \cdot \mathbf {H} }{2}}={\frac {\mathbf {B} \
cdot \mathbf {B} }{2\mu }}={\frac {\mu \mathbf {H} \cdot \mathbf {H} }{2}}.}
If there are no magnetic materials around then μ can be replaced by μ0. The above
equation cannot be used for nonlinear materials, though; a more general expression
given below must be used.
In general, the incremental amount of work per unit volume δW needed to cause a
small change of magnetic field δB is:
𝛿
𝑊
=
𝐻
⋅
𝛿
𝐵
.
{\displaystyle \delta W=\mathbf {H} \cdot \delta \mathbf {B} .}
Once the relationship between H and B is known this equation is used to determine
the work needed to reach a given magnetic state. For hysteretic materials such as
ferromagnets and superconductors, the work needed also depends on how the magnetic
field is created. For linear non-dispersive materials, though, the general equation
leads directly to the simpler energy density equation given above.
The second mathematical property is called the curl, such that ∇ × A represents how
A curls or "circulates" around a given point. The result of the curl is called a
"circulation source". The equations for the curl of B and of E are called the
Ampère–Maxwell equation and Faraday's law respectively.
More formally, since all the magnetic field lines that enter any given region must
also leave that region, subtracting the "number"[note 12] of field lines that enter
the region from the number that exit gives identically zero. Mathematically this is
equivalent to Gauss's law for magnetism:
∮
𝑆
𝐵
⋅
d
𝐴
=
0
{\displaystyle \oint _{S}\mathbf {B} \cdot \mathrm {d} \mathbf {A} =0}
where the integral is a surface integral over the closed surface S (a closed
surface is one that completely surrounds a region with no holes to let any field
lines escape). Since dA points outward, the dot product in the integral is positive
for B-field pointing out and negative for B-field pointing in.
Faraday's Law
Main article: Faraday's law of induction
A changing magnetic field, such as a magnet moving through a conducting coil,
generates an electric field (and therefore tends to drive a current in such a
coil). This is known as Faraday's law and forms the basis of many electrical
generators and electric motors. Mathematically, Faraday's law is:
𝐸
=
−
d
Φ
d
𝑡
{\displaystyle {\mathcal {E}}=-{\frac {\mathrm {d} \Phi }{\mathrm {d} t}}}
where
𝐸
{\displaystyle {\mathcal {E}}} is the electromotive force (or EMF, the voltage
generated around a closed loop) and Φ is the magnetic flux—the product of the area
times the magnetic field normal to that area. (This definition of magnetic flux is
why B is often referred to as magnetic flux density.)[35]: 210 The negative sign
represents the fact that any current generated by a changing magnetic field in a
coil produces a magnetic field that opposes the change in the magnetic field that
induced it. This phenomenon is known as Lenz's law. This integral formulation of
Faraday's law can be converted[note 13] into a differential form, which applies
under slightly different conditions.
∇
×
𝐸
=
−
∂
𝐵
∂
𝑡
{\displaystyle \nabla \times \mathbf {E} =-{\frac {\partial \mathbf {B} }{\partial
t}}}
Ampère's Law and Maxwell's correction
Main article: Ampère's circuital law
Similar to the way that a changing magnetic field generates an electric field, a
changing electric field generates a magnetic field. This fact is known as Maxwell's
correction to Ampère's law and is applied as an additive term to Ampere's law as
given above. This additional term is proportional to the time rate of change of the
electric flux and is similar to Faraday's law above but with a different and
positive constant out front. (The electric flux through an area is proportional to
the area times the perpendicular part of the electric field.)
The full law including the correction term is known as the Maxwell–Ampère equation.
It is not commonly given in integral form because the effect is so small that it
can typically be ignored in most cases where the integral form is used.
∇
×
𝐵
=
𝜇
0
𝐽
+
𝜇
0
𝜀
0
∂
𝐸
∂
𝑡
{\displaystyle \nabla \times \mathbf {B} =\mu _{0}\mathbf {J} +\mu _{0}\varepsilon
_{0}{\frac {\partial \mathbf {E} }{\partial t}}}
where J is the complete microscopic current density, and ε0 is the vacuum
permittivity.
∇
×
𝐻
=
𝐽
f
+
∂
𝐷
∂
𝑡
{\displaystyle \nabla \times \mathbf {H} =\mathbf {J} _{\mathrm {f} }+{\frac {\
partial \mathbf {D} }{\partial t}}}
These equations are not any more general than the original equations (if the
"bound" charges and currents in the material are known). They also must be
supplemented by the relationship between B and H as well as that between E and D.
On the other hand, for simple relationships between these quantities this form of
Maxwell's equations can circumvent the need to calculate the bound charges and
currents.
The magnetic field existing as electric field in other frames can be shown by
consistency of equations obtained from Lorentz transformation of four force from
Coulomb's Law in particle's rest frame with Maxwell's laws considering definition
of fields from Lorentz force and for non accelerating condition. The form of
magnetic field hence obtained by Lorentz transformation of four-force from the form
of Coulomb's law in source's initial frame is given by:[36]
𝐵
=
𝑞
4
𝜋
𝜀
0
𝑟
3
1
−
𝛽
2
(
1
−
𝛽
2
sin
2
𝜃
)
3
/
2
𝑣
×
𝑟
𝑐
2
=
𝑣
×
𝐸
𝑐
2
{\displaystyle \mathbf {B} ={\frac {q}{4\pi \varepsilon _{0}r^{3}}}{\frac {1-\beta
^{2}}{(1-\beta ^{2}\sin ^{2}\theta )^{3/2}}}{\frac {\mathbf {v} \times \mathbf
{r} }{c^{2}}}={\frac {\mathbf {v} \times \mathbf {E} }{c^{2}}}}
where
𝑞
{\displaystyle q} is the charge of the point source,
𝜀
0
{\displaystyle \varepsilon _{0}} is the vacuum permittivity,
𝑟
{\displaystyle \mathbf {r} } is the position vector from the point source to the
point in space,
𝑣
{\displaystyle \mathbf {v} } is the velocity vector of the charged particle,
𝛽{\displaystyle \beta } is the ratio of speed of the charged particle divided by
the speed of light and
𝜃{\displaystyle \theta } is the angle between
𝑟
{\displaystyle \mathbf {r} } and
𝑣
{\displaystyle \mathbf {v} }. This form of magnetic field can be shown to satisfy
maxwell's laws within the constraint of particle being non accelerating.[37] The
above reduces to Biot-Savart law for non relativistic stream of current (
𝛽
≪
1
{\displaystyle \beta \ll 1}).
Formally, special relativity combines the electric and magnetic fields into a rank-
2 tensor, called the electromagnetic tensor. Changing reference frames mixes these
components. This is analogous to the way that special relativity mixes space and
time into spacetime, and mass, momentum, and energy into four-momentum.[38]
Similarly, the energy stored in a magnetic field is mixed with the energy stored in
an electric field in the electromagnetic stress–energy tensor.
𝐵
=
∇
×
𝐴
,
𝐸
=
−
∇
𝜑
−
∂
𝐴
∂
𝑡
.
{\displaystyle {\begin{aligned}\mathbf {B} &=\nabla \times \mathbf {A} ,\\\mathbf
{E} &=-\nabla \varphi -{\frac {\partial \mathbf {A} }{\partial t}}.\end{aligned}}}
The vector potential, A given by this form may be interpreted as a generalized
potential momentum per unit charge [39] just as φ is interpreted as a generalized
potential energy per unit charge. There are multiple choices one can make for the
potential fields that satisfy the above condition. However, the choice of
potentials is represented by its respective gauge condition.
Maxwell's equations when expressed in terms of the potentials in Lorenz gauge can
be cast into a form that agrees with special relativity.[40] In relativity, A
together with φ forms a four-potential regardless of the gauge condition, analogous
to the four-momentum that combines the momentum and energy of a particle. Using the
four potential instead of the electromagnetic tensor has the advantage of being
much simpler—and it can be easily modified to work with quantum mechanics.
𝑡
𝑟
=
𝑡
−
|
𝑟
−
𝑟
𝑠
(
𝑡
𝑟
)
|
𝑐
{\displaystyle t_{r}=\mathbf {t} -{\frac {|\mathbf {r} -\mathbf {r} _{s}(t_{r})|}
{c}}}
where
𝑡
𝑟
{\textstyle {t_{r}}} is retarded time or the time at which the source's
contribution of the field originated,
𝑟
𝑠
(
𝑡
)
{\textstyle {r}_{s}(t)} is the position vector of the particle as function of time,
𝑟
{\textstyle \mathbf {r} } is the point in space,
𝑡
{\textstyle \mathbf {t} } is the time at which fields are measured and
𝑐
{\textstyle c} is the speed of light. The equation subtracts the time taken for
light to travel from particle to the point in space from the time of measurement to
find time of origin of the fields. The uniqueness of solution for
𝑡
𝑟
{\textstyle {t_{r}}} for given
𝑡
{\displaystyle \mathbf {t} },
𝑟
{\displaystyle \mathbf {r} } and
𝑟
𝑠
(
𝑡
)
{\displaystyle r_{s}(t)} is valid for charged particles moving slower than speed of
light.[42]
Any arbitrary motion of point charge causes electric and magnetic fields found by
solving maxwell's equations using green's function for retarded potentials and
hence finding the fields to be as follows:
𝐴
(
𝑟
,
𝑡
)
=
𝜇
0
𝑐
4
𝜋
(
𝑞
𝛽
𝑠
(
1
−
𝑛
𝑠
⋅
𝛽
𝑠
)
|
𝑟
−
𝑟
𝑠
|
)
𝑡
=
𝑡
𝑟
=
𝛽
𝑠
(
𝑡
𝑟
)
𝑐
𝜑
(
𝑟
,
𝑡
)
{\displaystyle \mathbf {A} (\mathbf {r} ,\mathbf {t} )={\frac {\mu _{0}c}{4\pi }}\
left({\frac {q{\boldsymbol {\beta }}_{s}}{(1-\mathbf {n} _{s}\cdot {\boldsymbol {\
beta }}_{s})|\mathbf {r} -\mathbf {r} _{s}|}}\right)_{t=t_{r}}={\frac {{\boldsymbol
{\beta }}_{s}(t_{r})}{c}}\varphi (\mathbf {r} ,\mathbf {t} )}
𝐵
(
𝑟
,
𝑡
)
=
𝜇
0
4
𝜋
(
𝑞
𝑐
(
𝛽
𝑠
×
𝑛
𝑠
)
𝛾
2
(
1
−
𝑛
𝑠
⋅
𝛽
𝑠
)
3
|
𝑟
−
𝑟
𝑠
|
2
+
𝑞
𝑛
𝑠
×
(
𝑛
𝑠
×
(
(
𝑛
𝑠
−
𝛽
𝑠
)
×
𝛽
𝑠
˙
)
)
(
1
−
𝑛
𝑠
⋅
𝛽
𝑠
)
3
|
𝑟
−
𝑟
𝑠
|
)
𝑡
=
𝑡
𝑟
=
𝑛
𝑠
(
𝑡
𝑟
)
𝑐
×
𝐸
(
𝑟
,
𝑡
)
{\displaystyle \mathbf {B} (\mathbf {r} ,\mathbf {t} )={\frac {\mu _{0}}{4\pi }}\
left({\frac {qc({\boldsymbol {\beta }}_{s}\times \mathbf {n} _{s})}{\gamma ^{2}(1-\
mathbf {n} _{s}\cdot {\boldsymbol {\beta }}_{s})^{3}|\mathbf {r} -\mathbf {r} _{s}|
^{2}}}+{\frac {q\mathbf {n} _{s}\times {\Big (}\mathbf {n} _{s}\times {\big (}(\
mathbf {n} _{s}-{\boldsymbol {\beta }}_{s})\times {\dot {{\boldsymbol {\
beta }}_{s}}}{\big )}{\Big )}}{(1-\mathbf {n} _{s}\cdot {\boldsymbol {\
beta }}_{s})^{3}|\mathbf {r} -\mathbf {r} _{s}|}}\right)_{t=t_{r}}={\frac {\mathbf
{n} _{s}(t_{r})}{c}}\times \mathbf {E} (\mathbf {r} ,\mathbf {t} )}
where
𝜑
(
𝑟
,
𝑡
)
{\textstyle \varphi (\mathbf {r} ,\mathbf {t} )}and
𝐴
(
𝑟
,
𝑡
)
{\textstyle \mathbf {A} (\mathbf {r} ,\mathbf {t} )} are electric scalar potential
and magnetic vector potential in Lorentz gauge,
𝑞
{\displaystyle q} is the charge of the point source,
𝑛
𝑠
(
𝑟
,
𝑡
)
{\textstyle {n}_{s}(\mathbf {r} ,t)} is a unit vector pointing from charged
particle to the point in space,
𝛽
𝑠
(
𝑡
)
{\textstyle {\boldsymbol {\beta }}_{s}(t)} is the velocity of the particle divided
by the speed of light and
𝛾
(
𝑡
)
{\textstyle \gamma (t)} is the corresponding Lorentz factor. Hence by the principle
of superposition, the fields of a system of charges also obey principle of
locality.
Quantum electrodynamics
See also: Standard Model and quantum electrodynamics
The classical electromagnetic field incorporated into quantum mechanics forms what
is known as the semi-classical theory of radiation. However, it is not able to make
experimentally observed predictions such as spontaneous emission process or Lamb
shift implying the need for quantization of fields. In modern physics, the
electromagnetic field is understood to be not a classical field, but rather a
quantum field; it is represented not as a vector of three numbers at each point,
but as a vector of three quantum operators at each point. The most accurate modern
description of the electromagnetic interaction (and much else) is quantum
electrodynamics (QED),[43] which is incorporated into a more complete theory known
as the Standard Model of particle physics.
All equations in this article are in the classical approximation, which is less
accurate than the quantum description mentioned here. However, under most everyday
circumstances, the difference between the two theories is negligible.
The field at the surface of the Earth is approximately the same as if a giant bar
magnet were positioned at the center of the Earth and tilted at an angle of about
11° off the rotational axis of the Earth (see the figure).[45] The north pole of a
magnetic compass needle points roughly north, toward the North Magnetic Pole.
However, because a magnetic pole is attracted to its opposite, the North Magnetic
Pole is actually the south pole of the geomagnetic field. This confusion in
terminology arises because the pole of a magnet is defined by the geographical
direction it points.[46]
Earth's magnetic field is not constant—the strength of the field and the location
of its poles vary.[47] Moreover, the poles periodically reverse their orientation
in a process called geomagnetic reversal. The most recent reversal occurred 780,000
years ago.[48]
Magnetic torque is used to drive electric motors. In one simple motor design, a
magnet is fixed to a freely rotating shaft and subjected to a magnetic field from
an array of electromagnets. By continuously switching the electric current through
each of the electromagnets, thereby flipping the polarity of their magnetic fields,
like poles are kept next to the rotor; the resultant torque is transferred to the
shaft.
A rotating magnetic field can be constructed using two orthogonal coils with 90
degrees phase difference in their AC currents. However, in practice such a system
would be supplied through a three-wire arrangement with unequal currents.
Synchronous motors use DC-voltage-fed rotor windings, which lets the excitation of
the machine be controlled—and induction motors use short-circuited rotors (instead
of a magnet) following the rotating magnetic field of a multicoiled stator. The
short-circuited turns of the rotor develop eddy currents in the rotating field of
the stator, and these currents in turn move the rotor by the Lorentz force.
The Italian physicist Galileo Ferraris and the Serbian-American electrical engineer
Nikola Tesla independently researched the use of rotating magnetic fields in
electric motors. In 1888, Ferraris published his research in a paper to the Royal
Academy of Sciences in Turin and Tesla gained U.S. patent 381,968 for his work.
Hall effect
Main article: Hall effect
The charge carriers of a current-carrying conductor placed in a transverse magnetic
field experience a sideways Lorentz force; this results in a charge separation in a
direction perpendicular to the current and to the magnetic field. The resultant
voltage in that direction is proportional to the applied magnetic field. This is
known as the Hall effect.
The Hall effect is often used to measure the magnitude of a magnetic field. It is
used as well to find the sign of the dominant charge carriers in materials such as
semiconductors (negative electrons or positive holes).
Magnetic circuits
Main article: Magnetic circuit
An important use of H is in magnetic circuits where B = μH inside a linear
material. Here, μ is the magnetic permeability of the material. This result is
similar in form to Ohm's law J = σE, where J is the current density, σ is the
conductance and E is the electric field. Extending this analogy, the counterpart to
the macroscopic Ohm's law (I = V⁄R) is:
Φ
=
𝐹
𝑅
m
,
{\displaystyle \Phi ={\frac {F}{R}}_{\mathrm {m} },}
where
Φ
=
∫
𝐵
⋅
d
𝐴
{\textstyle \Phi =\int \mathbf {B} \cdot \mathrm {d} \mathbf {A} } is the magnetic
flux in the circuit,
𝐹
=
∫
𝐻
⋅
d
ℓ{\textstyle F=\int \mathbf {H} \cdot \mathrm {d} {\boldsymbol {\ell }}} is the
magnetomotive force applied to the circuit, and Rm is the reluctance of the
circuit. Here the reluctance Rm is a quantity similar in nature to resistance for
the flux. Using this analogy it is straightforward to calculate the magnetic flux
of complicated magnetic field geometries, by using all the available techniques of
circuit theory.
This section needs to be updated. Please help update this article to reflect recent
events or newly available information.
Last update: October 2018 (July 2021)
As of October 2018, the largest magnetic field produced over a macroscopic volume
outside a lab setting is 2.8 kT (VNIIEF in Sarov, Russia, 1998).[49][50] As of
October 2018, the largest magnetic field produced in a laboratory over a
macroscopic volume was 1.2 kT by researchers at the University of Tokyo in 2018.
[50] The largest magnetic fields produced in a laboratory occur in particle
accelerators, such as RHIC, inside the collisions of heavy ions, where microscopic
fields reach 1014 T.[51][52] Magnetars have the strongest known magnetic fields of
any naturally occurring object, ranging from 0.1 to 100 GT (108 to 1011 T).[53]
History
Main article: History of electromagnetic theory
See also: Timeline of electromagnetism and classical optics
One of the first drawings of a magnetic field, by René Descartes, 1644, showing the
Earth attracting lodestones. It illustrated his theory that magnetism was caused by
the circulation of tiny helical particles, "threaded parts", through threaded pores
in magnets.
Early developments
While magnets and some properties of magnetism were known to ancient societies, the
research of magnetic fields began in 1269 when French scholar Petrus Peregrinus de
Maricourt mapped out the magnetic field on the surface of a spherical magnet using
iron needles. Noting the resulting field lines crossed at two points he named those
points "poles" in analogy to Earth's poles. He also articulated the principle that
magnets always have both a north and south pole, no matter how finely one slices
them.[54][note 14]
Mathematical development
Extending these experiments, Ampère published his own successful model of magnetism
in 1825. In it, he showed the equivalence of electrical currents to magnets[55]: 88
and proposed that magnetism is due to perpetually flowing loops of current instead
of the dipoles of magnetic charge in Poisson's model.[note 16] Further, Ampère
derived both Ampère's force law describing the force between two currents and
Ampère's law, which, like the Biot–Savart law, correctly described the magnetic
field generated by a steady current. Also in this work, Ampère introduced the term
electrodynamics to describe the relationship between electricity and magnetism.
[55]: 88–92
In 1850, Lord Kelvin, then known as William Thomson, distinguished between two
magnetic fields now denoted H and B. The former applied to Poisson's model and the
latter to Ampère's model and induction.[55]: 224 Further, he derived how H and B
relate to each other and coined the term permeability.[55]: 245 [61]
Between 1861 and 1865, James Clerk Maxwell developed and published Maxwell's
equations, which explained and united all of classical electricity and magnetism.
The first set of these equations was published in a paper entitled On Physical
Lines of Force in 1861. These equations were valid but incomplete. Maxwell
completed his set of equations in his later 1865 paper A Dynamical Theory of the
Electromagnetic Field and demonstrated the fact that light is an electromagnetic
wave. Heinrich Hertz published papers in 1887 and 1888 experimentally confirming
this fact.[62][63]
Modern developments
In 1887, Tesla developed an induction motor that ran on alternating current. The
motor used polyphase current, which generated a rotating magnetic field to turn the
motor (a principle that Tesla claimed to have conceived in 1882).[64][65][66] Tesla
received a patent for his electric motor in May 1888.[67][68] In 1885, Galileo
Ferraris independently researched rotating magnetic fields and subsequently
published his research in a paper to the Royal Academy of Sciences in Turin, just
two months before Tesla was awarded his patent, in March 1888.[69]
See also
General
Magnetohydrodynamics – the study of the dynamics of electrically conducting fluids
Magnetic hysteresis – application to ferromagnetism
Magnetic nanoparticles – extremely small magnetic particles that are tens of atoms
wide
Magnetic reconnection – an effect that causes solar flares and auroras
Magnetic scalar potential
SI electromagnetism units – common units used in electromagnetism
Orders of magnitude (magnetic field) – list of magnetic field sources and
measurement devices from smallest magnetic fields to largest detected
Upward continuation
Moses Effect
Mathematics
Magnetic helicity – extent to which a magnetic field wraps around itself
Applications
Dynamo theory – a proposed mechanism for the creation of the Earth's magnetic field
Helmholtz coil – a device for producing a region of nearly uniform magnetic field
Magnetic field viewing film – Film used to view the magnetic field of an area
Magnetic pistol – a device on torpedoes or naval mines that detect the magnetic
field of their target
Maxwell coil – a device for producing a large volume of an almost constant magnetic
field
Stellar magnetic field – a discussion of the magnetic field of stars
Teltron tube – device used to display an electron beam and demonstrates effect of
electric and magnetic fields on moving charges