International
Tables for Crystallography Volume C Mathematical, physical and chemical tables Edited by E. Prince © International Union of Crystallography 2006 |
International Tables for Crystallography (2006). Vol. C. ch. 4.3, pp. 397-404
Section 4.3.4.3. Excitation spectrum of valence electrons
C. Colliexa
|
Most inelastic interaction of fast incident electrons is with outer atomic shells in atoms, or in solids with valence electrons (referred to as conduction electrons in metals). These involve excitations in the 0–50 eV range, but, in a few cases, interband transitions from low-binding-energy shells may also contribute.
The basic concept introduced by the many-body theory in the interacting free electron gas is the volume plasmon. In a condensed material, the assembly of loosely bound electrons behaves as a plasma in which collective oscillations can be induced by a fast external charged particle. These eigenmodes, known as volume plasmons, are longitudinal charge-density fluctuations around the average bulk density in the plasma n 1028 e−/m3). Their eigen frequency is given, in the free electron gas, as
The corresponding
energy, measured in an energy-loss spectrum (see the famous example of the plasmon in aluminium in Fig. 4.3.4.3
), is the plasmon energy, for which typical values in a selection of pure solid elements are gathered in Table 4.3.4.2
. The accuracies of the measured values depend on several instrumental parameters. Moreover, they are sensitive to the specimen crystalline state and to its degree of purity. Consequently, there exist slight discrepancies between published values. Numbers listed in Table 4.3.4.2
must therefore be accepted with a 0.1 eV confidence. Some specific cases require comments: amorphous boron, when prepared by vacuum evaporation, is not a well defined specimen. Carbon exists in several allotropic varieties. The selection of the diamond type in the table is made for direct comparison with the other tetravalent specimens (Si, Ge, Sn). The results for lead (Pb) are still subject to confirmation. The volumic mass density is an important factor (through n) in governing the value of the plasmon energy. It varies with temperature and may be different in the crystal, in the amorphous, and in the liquid phases. In simple metals, the amorphous state is generally less dense than the crystalline one, so that its plasmon energy shifts to lower energies.
|
The above description applies only to very small scattering vectors q. In fact, the plasmon energy increases with scattering angle (and with momentum transfer ). This dependence is known as the dispersion relation, in which two distinct behaviours can be described:
|
Plasmon lifetime
is inversely proportional to the energy width of the plasmon peak . Even for Al, with one of the smallest plasmon energy widths (
eV), the lifetime is very short: after about five oscillations, their amplitude is reduced to 1/e. Such a damping demonstrates the strength of the coupling of the collective modes with other processes. Several mechanisms compete for plasmon decay:
|
Within this free-electron-gas description, the differential cross section for the excitation of bulk plasmons by incident electrons of velocity v is given by where N is the density of atoms per volume unit and
is the characteristic inelastic angle defined as
in the non-relativistic description and as
{with
} in the relativistic case. The angular dependence of the differential cross section for plasmon scattering is shown in Fig. 4.3.4.14
. The integral cross section up to an angle
is
The total plasmon cross section is calculated for
. Converted into mean free path, this becomes
and
The behaviour of as a function of the primary electron energy is shown in Fig. 4.3.4.15
.
The description of the bulk plasmon in the free-electron gas can be extended to any type of condensed material by introducing the dielectric response function , which describes the frequency and wavevector-dependent polarizability of the medium; cf. Daniels et al. (1970
). One associates, respectively, the
and
functions with the propagation of transverse and longitudinal EM modes through matter. In the small-q limit, these tend towards the same value:
As transverse dielectric functions are only used for wavevectors close to zero, the T and L indices can be omitted so that:
The transverse solution corresponds to the normal propagation of EM waves in a medium of dielectric coefficient
, i.e. to
For longitudinal fields, the only solution is
, which is basically the dispersion relation for the bulk plasmon.
In the framework of the Maxwell description of wave propagation in matter, it has been shown by several authors [see, for instance, Ritchie (1957)] that the transfer of energy between the beam electron and the electrons in the solid is governed by the magnitude of the energy-loss function
, so that
One can deduce (4.3.4.14)
by introducing a δ function at energy loss
for the energy-loss function:
As a consequence of the causality principle, a knowledge of the energy-loss function
over the complete frequency (or energy-loss) range enables one to calculate
by Kramers–Kronig analysis:
where PP denotes the principal part of the integral. For details of efficient practical evaluation of the above equation, see Johnson (1975
).
The dielectric functions can be easily calculated for simple descriptions of the electron gas. In the Drude model, i.e. for a free-electron plasma with a relaxation time τ, the dielectric function at long wavelengths is
with
, as above. The behaviour of the different functions, the real and imaginary terms in
, and the energy-loss function are shown in Fig. 4.3.4.16
. The energy-loss term exhibits a sharp Lorentzian profile centred at
and of width 1/τ. The narrower and more intense this plasmon peak, the more the involved valence electrons behave like free electrons.
In the Lorentz model, i.e. for a gas of bound electrons with one or several excitation eigenfrequencies , the dielectric function is
where
denotes the density of electrons oscillating with the frequency
and
is the associated relaxation time. The characteristic
,
, and
behaviours are displayed in Fig. 4.3.4.17
: a typical `interband' transition (in solid-state terminology) can be revealed as a maximum in the
function, simultaneous with a `plasmon' mode associated with a maximum in the energy-loss function and slightly shifted to higher energies with respect to the annulation conditions of the
function.
In most practical situations, there coexist a family of free electrons (with plasma frequency
and one or several families of
bound electrons (with eigenfrequencies
. The influence of bound electrons is to shift the plasma frequency towards lower values if
and to higher values if
. As a special case, in an insulator,
and all the electrons
have a binding energy at least equal to the band gap
, giving
.
This description constitutes a satisfactory first step into the world of real solids with a complex system of valence and conduction bands between which there is a strong transition rate of individual electrons under the influence of photon or electron beams. In optical spectroscopy, for instance, this transition rate, which governs the absorption coefficient, can be deduced from the calculation of the factor as
where
is the matrix element for the transition from the occupied level j in the valence band to the unoccupied level
in the conduction band, both with the same k value (which means for a vertical transition).
is the joint density of states (JDOS) with the energy difference
. This formula is also valid for small-angle-scattering electron inelastic processes. When parabolic bands are used to represent, respectively, the upper part of the valence band and the lower part of the conduction band in a semiconductor, the dominant JDOS term close to the onset of the interband transitions takes the form
where
is the band-gap energy. This concept has been successfully used by Batson (1987
) for the detection of gap energy variations between the bulk and the vicinity of a single misfit dislocation in a GaAs specimen. The case of non-vertical transitions involving integration over k-space has also been considered (Fink et al., 1984
; Fink & Leising, 1986
).
The dielectric constants of many solids have been deduced from a number of methods involving either primary photon or electron beams. In optical measurements, one obtains the values of and
from a Krakers–Kronig analysis of the optical absorption and reflection curves, while in electron energy-loss measurements they are deduced from Kramers–Kronig analysis of energy-loss functions.
Fig. 4.3.4.18
shows typical behaviours of the dielectric and energy-loss functions.
|
Čerenkov radiation is emitted when the velocity v of an electron travelling through a medium exceeds the speed of light for a particular frequency in this medium. The criterion for Čerenkov emission is
In an insulator, is positive at low energies and can considerably exceed unity, so that a `radiation peak' can be detected in the corresponding energy-loss range (between 2 and 4 eV in Si, Ge, III–V compounds, diamond,
); see Von Festenberg (1968
), Kröger (1970
), and Chen & Silcox (1971
). The associated scattering angle,
for high-energy electrons, is very small and this contribution can only be detected using a limited forward-scattering angular acceptance.
In an anisotropic crystal, the dielectric function has the character of a tensor, so that the energy-loss function is expressed as
If it is transformed to its orthogonal principal axes , and if the q components in this system are
, the above expression simplifies to
In a uniaxial crystal, such as a graphite, and
(i.e. parallel to the c axis):
where
is the angle between q and the c axis. The spectrum depends on the direction of q, either parallel or perpendicular to the c axis, as shown in Fig. 4.3.4.19
from Venghaus (1975
). These experimental conditions may be achieved by tilting the graphite layer at 45° with respect to the incident axis, and recording spectra in two directions at
with respect to it (see Fig. 4.3.4.20
).
Volume plasmons are longitudinal waves of charge density propagating through the bulk of the solid. Similarly, three exist longitudinal waves of charge density travelling along the surface between two media A and B (one may be a vacuum): these are the surface plasmons (Kliewer & Fuchs, 1974). Boundary conditions imply that
The corresponding charge-density fluctuation is contained within the (x) boundary plane, z being normal to the surface:
and the associated electrostatic potential oscillates in space and time as
The characteristic energy
of this surface mode is estimated in the free electron case as:
In the spherical interface case: (metal sphere in vacuum – the modes are now quantified following the l quantum number in spherical geometry);
(spherical void within metal).
Thin-film geometry: (metal layer of thickness t embedded in dielectric films of constant
). The two solutions result from the coupling of the oscillations on the two surfaces, the electric field being symmetric for the
mode and antisymmetric for the
.
In a real solid, the surface plasmon modes are determined by the roots of the equation for vacuum coating [or
for dielectric coating].
The probability of surface-loss excitation is mostly governed by the
energy-loss function, which is analogous for surface modes to the bulk
energy-loss function. In normal incidence, the differential scattering cross section
is zero in the forward direction, reaches a maximum for
, and decreases as
at large angles. In non-normal incidence, the angular distribution is asymmetrical, goes through a zero value for momentum transfer
in a direction perpendicular to the interface, and the total probability increases as
where
is the incidence angle between the primary beam and the normal to the surface. As a consequence, the probability of producing one (and several) surface losses increases rapidly for grazing incidences.
References


























