2.1. Integro-Differential Equation for the Power Spectrum
Luminescence, or equivalently the optical power density spectrum
, is described quantum mechanically by the Poynting vector, which can be directly related to the transverse polarization function
, which is the selfenergy in the Dyson equations for the transverse photon Green’s function components in the Keldysh formalism [
7,
19,
20]. All the quantities presented in this paper are considered in frequency space, i.e., evaluated at steady state.
The free photon Green’s function represents the photons propagating without any interaction with the medium. When carriers are injected the transverse polarization function
, which is the selfenergy in photon Green’s function Dyson equation, determines how the excited medium modifies the photon propagation. The lesser Keldysh component
is proportional to the carriers recombination rate and yields the number of emitted photons per unit area. It thus governs the power emission spectrum, as seen in Equation (1). The imaginary and real parts of
are, respectively, proportional to absorption and gain and refractive index changes, since the dielectric function of the medium reads
as shown in Ref. [
19]. The starting points for the results derived here are the equations in Refs. [
7,
19,
20].
Here,
, and
denote, respectively, the electron charge, the speed of light, the sample volume and the velocity matrix element, which is the expectation value of the velocity operator, i.e., the momentum operator divided by the electron mass. It stems directly from the fact that current is charge times velocity. The formal definition of the transverse polarization function selfenergy in terms of functional derivatives is
, where
and
are, respectively, expectation values of the induced current and vector potential operators. The full expression involves labels along the Keldysh contour and is tensorial. A complete discussion is beyond the scope of this paper. For details see Refs. [
19,
20].
The crystal momentum
is a consequence of Fourier transforming from real space. Likewise,
ω and thus the photon energy
stem from a corresponding Fourier transformation from time to the frequency domain. The matrix element satisfies the integro-differential equation [
20],
where
is the screened Hulthén potential [
21,
22,
23,
24]. Furthermore,
Electrons or holes are labelled, respectively, by , the renormalized energies , and dephasing are calculated from the real and imaginary parts of the selfenergy in the Dyson equation for the retarded carriers Green’s functions. This paper focuses on quasi-equilibrium luminescence and on three dimensional (bulk semiconductors) with one conduction and one valence band.
Under these conditions,
denotes a Fermi function characterized by a chemical potential
and the spectral function in Equation (4) for each particle, derived from components of the carriers’ Green’s function in the Keldysh formalism, reads
The next step is to re-write the last term in Equation (4) by means of the identity
and to approximate this factor by
, where
is the total chemical potential, where
T is the temperature in Kelvins and
is the Boltzmann constant. Different versions of this approximation has been used before in phenomenological approaches for absorption Refs. [
21,
22,
23] and delivered good agreement with experiments (see details and further references in Ref. [
23]). Within the Keldysh Green’s functions, context, a detailed derivation of its application is given in Ref. [
20]. The fully numerical solutions of the equations that use this version of the approximation have given very good agreement with both single beam and pump-probe luminescence [
20,
25]. Its usefulness has been further confirmed recently by the good agreement between the analytical solutions shown here and the experimental luminescence of dilute semiconductors [
7,
8,
9,
14].
Figure 1 and
Figure 2 depict typical ranges of parameters, showing that
is an excellent approximation. Note that this theory is applied for photon energies around the semiconductor bandgap and thus in
Figure 1 and
Figure 2,
.
Low temperature luminescence is typically performed with a small density of injected carriers. Very good agreement of this theory with results from different experimental teams for a variety of materials has been obtained with carrier densities around 10
15 carriers/cm
3 [
7,
8,
9], further justifying the range of densities in the
y-axis. The theory has also been used for high temperatures and high densities to investigate optical nonlinearites [
14,
15,
16], and this range is illustrated in
Figure 2.
Relevant dephasing mechanisms such as electron-electron, electron-phonon and electron-impurity scattering can be added to the selfenergy [
17,
18], and the resulting
is frequency and momentum dependent. However, in what follows, it is replaced by averaged values, leading to a simple approximation for
consistent with the Ansatz solution,
where
and
. In 3D, the material resonance energy is:
, where
The bandgap
is given by the sum of the fundamental band gap
, and a many body renormalisation term
where
denote, respectively the electron and hole effective masses. The equation for
simplifies to:
The total dephasing Γ will determine the luminescence linewidth. Thus, it can be treated as a phenomenological parameter used to interpret data, and at the same time estimate the strength of the scattering and dephasing processes [
7,
8,
9] by comparison of adjusted data with microscopic calculations derived from the relevant selfenergies [
17,
18]. At this point, the Kubo–Martin–Schwinger (KMS) relation under the form derived in Ref. [
20] can be applied to Equation (8),
together with the auxiliary variable:
, leading to the relation:
Expressing
from Equation (8) in terms of
the corresponding integro-differential equation becomes
where
.
Before proceeding, the Hulthén potential [
21,
22,
23,
24] should be revised. The usual approximation for a static 3D screened potential is the Yukawa potential,
. However, the corresponding Schrödinger equation does not have known analytical solutions. In contrast, the Hulthén potential:
, has known analytical solutions that have proven to be very useful for the description of bulk absorption [
22]. Recent applications have confirmed its relevance to explain experimental luminescence studies [
7,
8,
9].
Figure 3 shows that, in the range of carrier densities and temperatures of interest, the Yukawa potential can be replaced by the Hulthén potential with negligible differences in numerical values.
The Fourier transform of the Hulthén potential has an analytical expression,
where
is the sample volume,
is the Trigamma function [
26],
is the screening wavenumber and by including
in
at Equation (11),
. Analytical approximations for
and
are given in Ref. [
7]. Note that the bandgap renormalization including Coulomb hole and screened exchange corrections reads
The Fermi functions
are evaluated at the peak of the spectral function for each particle, i.e., in Equation (5),
. More details are given in Ref. [
7]. Equation (13) goes beyond phenomenological term for the bandgap shift [
21,
22,
23], and also, in contrast to those, here we can in principle take into account a reduction in the Coulomb interaction due to phase space filling through the factor
. Note however that in the range of carrier densities and temperatures of interest
i.e.,
, as shown in
Figure 4.
At quasi-equilibrium, used in Refs. [
7,
8,
9,
14,
15,
16,
22], the total chemical potential
is calculated self-consistently with the many body renormalization of the bandgap
and can be written exactly as
, where
is the total free carrier chemical potential calculated from the bottom of each band. In other words, the inversion factor can be equivalently written as
, and it is now clear why
Figure 4 has the detuning
in the
x-axis. Furthermore, the 3D exciton binding energy for GaAs is 4.2 meV and there is no luminescence of absorption for a detuning below 4.2 meV, unless there are deep levels due to impurities and defects, which are not considered here. Thus the approximation,
i.e.,
for the dielectric constant used in the Hulthén potential is clearly excellent in the low power luminescence case. Nonlinear absorption studies are only meaningful away from population inversion leading to optical gain, i.e., the studies are in the range
. Thus a decreasing occupation reflects phase space feeling and even for
the approach is valid. In order to study the gain regime, the approximation used in the literature is to make at the Hulthén potential
and consider the inversion factor only on the right hand side of Equation (8). In other words, in the traditional “plasma theories” for bulk semiconductor absorption and gain, phase space filling (
is not taken into account. Since in the high density case where gain develops, the Hulthén potential decreases due to screening, which described by large
in Equation (12), there is still good agreement with experiments. See e.g., Refs. [
22,
23].
Equation (11) can now be Fourier-transformed
Here
is the sample volume,
denotes the Dirac delta function. Expanding
in the basis of eigenstates of the Hamiltonian:
,
Thus, Equation (15) can be rewritten as
Projection onto state
yields
Substitution into Equation (16)
Fourier-transforming back to
k-space
From
and
, a closed expression can be obtained.
where a factor 2 for spin has been explicitly written out of the summation over all quantum numbers. Introducing and combining Equations (1), (2), (10) and (21) leads to
where
reduces to a Dirac delta function for
. The velocity matrix element is expressed in terms of the dipole moment matrix element and the fundamental bandgap as
. Next, the Schrödinger Equation for the Hulthén potential must be solved, so that
can be inserted in Equation (23). The first step is to separate the wavefunction in radial and angular parts. The label
thus spans the set
,
The corresponding Schrödinger Equation, which is a generalized Wannier equation [
23] can be cast in the form:
The energy eigenvalues depend only on the
quantum numbers, and thus we can replace
by
. Introducing the 3D Rydberg
and Bohr Radius
, as well
,
and
, leads to
Note that the angular momentum operator has been applied to the wavefunction directly from Equation (25) to Equation (26), i.e.,
Only solutions that do not vanish at
contribute to the emitted power, so
is selected. The labels
will be dropped at the moment to simplify the notation,
Introducing
, and
Equation (27) is transformed into
The auxiliary variables,
and
, lead to the equation
which reduces to the Hypergeometric Equation [
26,
27],
The generalized Wannier Equation, Equation (25), has two types of solutions: bound states for and unbound solutions for . The wavefunctions and eigenvalues are thus different and it makes sense to study each case separately and then add all contributions when a sum over all possible as required from Equation (23).
2.3. Continuum States
The unbound solutions that make a continuum have positive eigenvalues,
, and thus imaginary
. Dropping labels to simplify the development in the next equations yields
which can be written for simplicity as
Note that the transformation
is being used. The solution that will be later inserted in Equation (23), will be normalized in a sphere of radius
and asymptotic solutions, obtained a large radius
will be investigated. Next, Equation (15.3.6) from Ref. [
26] is used, i.e.,
Furthermore, note that
and for all values of
. Thus,
, which, combined with
(Equation (6.1.2.3) of Ref. [
26]), gives in the asymptotic limit
Leading to asymptotic forms of
where
. The normalization constant is thus given by
,
However,
see e.g., Equations (6.1.29) and (6.1.31) of Ref. [
26], plus a little algebra deliver the continuum normalization constant
The required value of the wave function at the origin can thus be expressed as
Next, note that
, but
, leading to
The sum of continuum states becomes an integral,
Introducing 2 for spin and changing variables, the continuum contribution becomes
which combines with the bound states
to deliver the power spectrum
Here, , , and the square of the velocity matrix element is , where the spin orbit shift, the free-carrier bandgap, and renormalized bandgap are given by , and .
Note that this approach does not include cavity effects, which can be introduced in the Photon Green’s functions solution following Ref. [
19]. Quasi-periodic structures can also be addressed by a Green’s functions formalism as shown in Ref. [
28].