Electrical Conductivity and Ohms Law

Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 35

ELECTRICAL

CONDUCTIVITY AND
OHM’S LAW
ELECTRICAL CONDUCTIVITY AND OHM’S LAW

• The momentum of a free electron is related to the wavevector by mv


= ħk.
• In an electric field E and magnetic field B the force F on an electron of
charge -e is -e[E + (1/c)v x B], so that Newton’s second law of motion
becomes
• In the absence of collisions the Fermi sphere (Fig. 10) moves in k space at a uniform rate by a
constant applied electric field. We integrate (39) with B = 0 to obtain

• If the force F = -eE is applied at time t = 0 to an electron gas that fills the Fermi sphere centered at
the origin of k space, then at a later time t the sphere will be displaced to a new center at

• Notice that the Fermi sphere is displaced as a whole because every electron is displaced by the
same k.
• Because of collisions of electrons with impurities,
lattice imperfections, and phonons, the displaced
sphere may be maintained in a steady state in an
electricfield.
• If the collision time is , the displacement of the Fermi
sphere in the steady state is given by (41) with t = .
• The incremental velocity is v = k/m = -eE/m.
• If in a constant electric field E there are n electrons of
charge q = -e per unit volume, the electric current
density is

(This is Ohm’s law.)


Values of the electrical conductivity and resistivity of the
elements are given in Table 3. In Gaussian units has the
dimensions of frequency.
Experimental Electrical Resistivity of Metals

• The electrical resistivity of most metals is dominated at room temperature (300


K) by collisions of the conduction electrons with lattice phonons and at liquid
helium temperature (4 K) by collisions with impurity atoms and mechanical
imperfections in the lattice (Fig. 11).
• The rates of these collisions are often independent to a good approximation, so
that if the electric field were switched off the momentum distribution would
relax back to its ground state with the net relaxation rate

where L and i are the collision times for scattering by phonons and by
imperfections, respectively.
• The net resistivity is given by

• where L is the resistivity caused by the thermal phonons, and i is the


resistivity caused by scattering of the electron waves by static defects that
disturb the periodicity of the lattice.
• Often L is independent of the number of defects when their
concentration is small, and often i is independent of temperature.
• This empirical observation expresses Matthiessen’s rule, which is
convenient in analyzing experimental data (Fig. 12).
• The residual resistivity, i(0), is the extrapolated resistivity at 0 K because L vanishes as T →
0.
• The lattice resistivity, L(T) =  - i(0), is the same for different specimens of a metal, even
though i(0) may itself vary widely.
• The resistivity ratio of a specimen is usually defined as the ratio of its resistivity at room
temperature to its residual resistivity.
• It is a convenient approximate indicator of sample purity: for many materials an impurity in
solid solution creates a residual resistivity of about 1 ohm-cm (1  106 ohm-cm) per
atomic percent of impurity.
• A copper specimen with a resistivity ratio of 1000 will have a residual resistivity of 1.7  10-3
ohm-cm, corresponding to an impurity concentration of about 20 ppm.
• In exceptionally pure specimens the resistivity ratio may be as high as 10 6, whereas in some
alloys (e.g., manganin) it is as low as 1.1.
• It is possible to obtain crystals of copper so pure that their conductivity at
liquid helium temperatures (4 K) is nearly 105 times that at room temperature;
for these conditions   2  10-9 s at 4 K.
• The mean free path of a conduction electron is defined as

• where vF is the velocity at the Fermi surface, because all collisions involve only
electrons near the Fermi surface. From Table 1 we have vF = 1.57  108 cm-s-1
for Cu, thus the mean free path is l(4 K)  0.3 cm.
• Mean free paths as long as 10 cm have been observed in very pure metals in
the liquid helium temperature range.
• The temperature-dependent part of the electrical resistivity is
proportional to the rate at which an electron collides with thermal
phonons and thermal electrons.
• The collision rate with phonons is proportional to the concentration of
thermal phonons.
• One simple limit is at temperatures over the Debye temperature :
here the phonon concentration is proportional to the temperature T,
so that   T for T > . A sketch of the theory is given in Appendix J.
Umklapp Scattering
• Umklapp scattering of electrons by phonons (Chapter 5) accounts for
most of the electrical resistivity of metals at low temperatures.
• These are electron-phonon scattering processes in which a reciprocal
lattice vector G is involved, so that electron momentum change in the
process may be much larger than in a normal electron-phonon
scattering process at low temperatures.
(In an umklapp process the wavevector of one particle may be “flipped
over.”)
• Consider a section perpendicular to [100] through two adjacent Brillouin zones in
bcc potassium, with the equivalent Fermi spheres inscribed within each (Fig. 13).
• The lower half of the figure shows the normal electron-phonon
• collision k’ = k + q, while the upper half shows a possible scattering process k’ = k
+ q + G involving the same phonon and terminating outside the first Brillouin
zone, at the point A.
• This point is exactly equivalent to the point A’ inside the original zone, where AA’
is a reciprocal lattice vector G.
• This scattering is an umklapp process, in analogy to phonons. Such collisions are
strong scatterers because the scattering angle can be close to .
• When the Fermi surface does not intersect the zone boundary, there
is some minimum phonon wavevector q0 for umklapp scattering.
• At low enough temperatures the number of phonons available for
umklapp scattering falls as exp(-U/T), where U is a characteristic
temperature calculable from the geometry of the Fermi surface inside
the Brillouin zone.
• For a spherical Fermi surface with one electron orbital per atom
inside the bcc Brillouin zone, one
• shows by geometry that q0 = 0.267kF.
• The experimental data (Fig. 12) for potassium have the expected
exponential form with U = 23 K compared with the Debye  = 91 K.
• At the very lowest temperatures (below about 2 K in potassium) the
number of umklapp processes is negligible and the lattice resistivity is
then caused only by small angle scattering, which is the normal (not
umklapp) scattering.
MOTION IN MAGNETIC FIELDS
• By the arguments of (39) and (41) we are led to the
equation of motion for the displacement k of a Fermi
sphere of particles acted on by a force F and by friction
as represented by collisions at a rate 1/:

• The free particle acceleration term is (ħd/dt)k and the


effect of collisions (the friction) is represented by ħk/
where  is the collision time.
Consider now the motion of the system in a uniform magnetic field
B. The Lorentz force on an electron is
Hall Effect
• The Hall field is the electric field developed across two faces of a
conductor, in the direction j  B, when a current j flows across a
magnetic field B.
• Consider a rod-shaped specimen in a longitudinal electric field Ex and
a transverse magnetic field, as in Fig. 14. If current cannot flow out of
the rod in the y direction we must have vy = 0. From (52) this is
possible only if there is a transverse electric field
• The lower the carrier concentration, the greater the magnitude of the
Hall coefficient. Measuring RH is an important way of measuring the
carrier concentration.
• Note: The symbol RH denotes the Hall coefficient (54), but the same
symbol is sometimes used with a different meaning, that of Hall
resistance in two-dimensional problems.
• The simple result (55) follows from the assumption that all relaxation
times are equal, independent of the velocity of the electron.
• A numerical factor of order unity enters if the relaxation time is a
function of the velocity.
• The expression becomes somewhat more complicated if both
electrons and holes contribute to the conductivity.
• In Table 4 observed values of the Hall coefficient are compared with
values calculated from the carrier concentration. The most accurate
measurements are made by the method of helicon resonance which
is treated as a problem in Chapter 14.
• The accurate values for sodium and potassium are in excellent
agreement with values calculated for one conduction electron per
atom, using (55).
• Notice, however, the experimental values for the trivalent elements
aluminium and indium: these agree with values calculated for one
positive charge carrier per atom and thus disagree in magnitude and
sign with values calculated for the expected three negative charge
carriers.
• The problem of an apparent positive sign for the charge carriers arises
also for Be and As, as seen in the table.
• The anomaly of the sign was explained by Peierls (1928).
• The motion of carriers of apparent positive sign, which Heisenberg
later called “holes,” cannot be explained by a free electron gas, but
finds a natural explanation in terms of the energy band theory to be
developed in Chapters 7–9.
• Band theory also accounts for the occurrence of very large values of
the Hall coefficient, as for As, Sb, and Bi.
THERMAL CONDUCTIVITY OF
METALS
• In Chapter 5 we found an expression for the thermal conductivity of
particles of velocity v, heat capacity C per unit volume, and mean free
path .
• The thermal conductivity of a Fermi gas follows from (36) for the heat
capacity, and with F = ½ mvF2

Here is l = vF; the electron concentration is n, and  is the collision time.


• Do the electrons or the phonons carry the greater part of the heat
current in a metal?
• In pure metals the electronic contribution is dominant at all
temperatures.
• In impure metals or in disordered alloys, the electron mean free path is
reduced by collisions with impurities, and the phonon contribution
may be comparable with the electronic contribution.
Ratio of Thermal to Electrical Conductivity
• The Wiedemann-Franz law states that for metals at
not too low temperatures the ratio of the thermal
conductivity to the electrical conductivity is directly
proportional to the temperature, with the value of the
constant of proportionality independent of the
particular metal.
• This result was important in the history of the theory
of metals, for it supported the picture of an electron
gas as the carrier of charge and energy. It can be
explained by using (43) for 
and (56) for K:

You might also like