1.1. State of the Art
Silicon–germanium alloys of the type
, with composition
changing inside the system, are extensively used in modern technology [
1] as, for instance, in design of thermoelectric energy generators [
2]. The thermoelectric efficiency of energy generators
, with
the electric power output, and
thermal power entering in the system, is an increasing function the so called figure-of-merit
, where
is the Seebeck coefficient,
the electrical conductivity, and
the thermal conductivity [
3,
4]. It is easily verified the the physical dimension of Z is the inverse of temperature, namely
. Hence, since it is sometime convenient to work with non-dimensional quantities, some authors denote as figure-of-merit the quantity ZT, with T the temperature. Critical analysis of assigning the figure-of-merit is given in [
5,
6]. If one remains in the frame of linear thermodynamics, for a system whose two sides are constantly maintained at different temperatures
(the hottest temperature) and
(the coldest one), it can be proved that the maximum efficiency is [
4]
wherein
is the classical Carnot efficiency, and
. Thus, the higher
, (and, hence, the smaller
), the higher
. However, any strategy for the reduction in
should take into account the results of phonon hydrodynamics, wherein it is assumed that the transport of heat is due to the phonons [
7]. Phonons are quasiparticles, which in a solid crystal form a rarefied gas, whose kinetic equation can be obtained similarly to that of an ordinary gas. Phonons interact among themselves and with the crystal lattice through:
Normal (N) processes, conserving the phonon momentum;
Resistive (R) processes, in which the phonon momentum is not conserved.
The frequencies
and
of normal and resistive processes, respectively, determine the characteristic relaxation times
and
. Purely diffusive heat transport takes over when there are many more R processes than N processes, i.e., when
tends to infinity and
tends to zero. If, instead, there are only few R processes and many more N processes, then
tends to zero, and a wavelike energy transport (second sound propagation) may occur. The total relaxation time
can be calculated by according to the Mathiessen rule
, while the thermal conductivity is given by
, where
is the mass density,
the specific heat per unit mass at constant volume, and
is the average of the phonons’ speed. Thus, a reduction in
would produce a reduction of
, i.e., an increment of the phonon scattering [
7,
8], with a consequent increment of dissipation. Thus, numerator and denominator in the expression of
cannot be controlled independently, and one should look for reductions of
which produce moderate increment of phonon scattering.
Materials with composition-dependent thermal conductivity are often used to enhance Z [
2,
9,
10,
11] since, by grading appropriately the stochiometry,
can be reduced, achieving so a consequent increment of Z [
4,
7]. We also observe that a further improvement of Z could be obtained by increasing the product
as function of
c. At the moment we are not aware of the data concerning the dependency of
and
on
c, but such a task could be considered in future studies. Herein, we limit ourselves to take into account the dependency of
and
on temperature reported in [
12], and use three different values of those quantities for three different operational temperatures. Different physical quantities as, for instance, the electronic part of the thermal conductivity, could be considered as well. The role of such material parameter has been considered in [
13,
14]. Indeed, the problem under consideration depends on several parameters of different nature. Herein, in order to obtain applicable results, we limited ourselves to consider a simple but meaningful case.
The thermoelectric efficiency of graded systems has been investigated in Refs. [
15,
16,
17,
18,
19], where the dependency of the performance of a thermoelectric energy generator as function of the two independent parameters
c and
x, with
c as the composition of the system, and
x as the square root of the mean value of the temperature gradient applied to its boundaries, has been investigated extensively. We used the square root of the mean value of the temperature gradient and not the temperature gradient itself, in order to obtain a more manageable expression for the rate of energy dissipated. The first independent parameter, i.e., the composition cannot be tuned externally, since
c is fixed after manufacturing the system. The second parameter, instead related to the applied temperature gradient, can be tuned externally. We have pointed out that the analysis of those systems yields new information on how manufacturing homogeneous thermoelectric generators made by silicon–germanium alloys, by determining the composition and the temperature gradient which optimize their efficiency. It is worth observing that our aim in [
15,
16,
17,
18,
19] was not to obtain explicit values of the efficiency for given independent parameters, because such values depend on several physical quantities which can be changed arbitrarily, such as, for instance, the geometry of the system and the external electric field. This fact motivated our choice of determining the values of the independent physical parameters which minimize the energy dissipated along the thermoelectric process, whatever is the value of such energy. One could wonder if the minimum of energy dissipated corresponds to the optimal efficiency of the thermoelectric process. Our answer is affirmative, and is based on the following considerations. In the following, we disregard all the losses introduced in the production of
and in the management of the generated difference of electrical potential, and we focus only on the thermodynamic process inside a thermoelectric wire of length L. It consists in the generation of an electric potential after that an amount of heat per unit time
entered the system. Such a heat produces dissipation by Joule effect which, in any point
z of the system and at any time
t, is represented by the rate of energy dissipated
. Then, in a thermoelectric process of duration
, the total energy dissipated is given by
so that, being the integrand a positive quantity, the right-hand side of Equation (
2) is minimum if, and only if, the integrand function is minimum in any point of the domain of integration [
19].
We underline again that the previous analysis regards only the process of thermoelectric energy conversion inside the wire. A more complete analysis should include the dissipation due to the production of , and that due to the transport and management of the obtained difference of electric potential, as well as the details of the processing parameters. However, such an analysis is outside the scopes of the present research, and is more pertinent to the field of engineering.
In order to make our investigation applicable to real cases, in [
15,
16,
17,
18,
19] we used the experimental data on thermal conductivity of silicon–germanium alloys as function of the composition, at
,
, and
[
20,
21,
22]. From the disposition of the experimental points in the plane
it is evident that the thermal conductivity is very steep in the two regions close to
, and
. Indeed, around
, due to the small value of
c, the alloy can be considered as doped Ge, while, close
, due to the high value of
c, the alloy can be considered as doped Si. Thus, since doped crystalline semiconductors have reduced thermal conductivity with respect to the alloy, such a rapid decreasing is expected. Between those two steep parts of the curve, the experimental values of
have small variation, so that
presents a kind of wide minimum between the two dilute zones
and
. From the mathematical point of view, such a behavior is well represented by the sum of two exponential functions
with
,
,
,
,
and
as unknown (non independent) parameters, to be determined by NLRM, under the constraints
and
. Thus, as first step, we have looked for a fitting curve represented by the Equation (
3), and applied the following iterative procedure [
23]:
Starting from the disposition of the experimental data in the plane
, we have assigned an initial estimated value of each parameter entering Equation (
3);
We have generated the curve defined by the initial values of the parameters;
We have calculated the sum of the squares (i.e., the sum of the squares of the vertical distances of the experimental points from the curve);
We have adapted the parameters in such a way that the curve was as close as possible to the experimental points;
We have stopped the calculations when we have observed a negligible difference of results in successive iterations.
In this way, we have obtained the best-fit curve of the data as
with
M,
N,
P,
Q four independent material parameters, and
and
two material functions of them. The constraints
and
led to the following form of the functions
and
Some more details of NLRM applied to obtain a fit of the experimental data are given in Ref. [
16]. Therein, it is illustrated the method for minimizing the mean distance between the experimental points and the fitting ones. Of course, such a distance cannot be reduced to zero, and this error affects the coefficients
M,
N,
P,
Q, and, as a consequence, the function
given in Equations (
4) and (
5). We expect that this error will influence also the optimal values of
c, T, and
calculated in
Section 3.
In [
15,
16,
17,
18,
19], we have studied the effects of the action of an electric field
on a graded
wire of length L, crossed by an electric current
. The right-hand side
, was supposed to be at the temperature
, while the left-hand side,
, was supposed to be at the temperature
. We assumed that
was flowing uniformly from left to right, and that the heat rate
was entering uniformly into the hot side of the system, giving rise to a heat flux
.
In our investigations, we applied the basic equations of thermoelectricity, whose physical meaning is discussed, for instance, in [
4,
24]. One of the key quantities in thermoelectricity is the form of the heat flux. Classically, the Fourier-like constitutive equation
where
is the Peltier coefficient, and
is the additional heat flux due to the circulation of the electric current
[
4,
7,
25,
26] is postulated. However, heat transport theory is currently broadening its field of applicability since, owing to the miniaturization, new phenomenologies, beyond the classical Fourier theory of heat conduction, have been discovered [
27,
28]. Those new phenomena depend on the relationship between the mean free path of the heat carriers
ℓ, and the characteristic dimension of the conductor L, expressed by the Knudsen number
. Fourier’s law is valid when
, namely, for
. However, Kn can increase for a reduction in
L, as in miniaturization technologies. Thus, when the mean free path of the heat carriers is comparable to the characteristic dimension of the conductor, i.e.,
, more complicated transport laws for the heat flux are necessary [
19,
27,
28]. In the present investigation, the constitutive equation for
has been supposed to be
where
denotes a characteristic-length vector, and
b,
, is a dimensionless physical parameter entering the effective thermal conductivity
[
25,
26]. In stationary situations, Equation (
7) arises in thermomass (TM) theory [
29,
30] of heat transport. In a non-stationary case, the thermomass heat transport equation reads
wherein
is a relaxation time [
25,
26,
29,
30]. In thermomass description, the heat flux is generated by a gas of heat carriers, characterized by an effective mass density and flowing through the medium under the action of a thermomass–pressure gradient. This gas is made by massive quasi-particles of heat carriers, named thermons, which are nothing but the vibrations of the molecules generated by heating the conductor, with null rest-mass and dynamic mass which may be calculated from the Einstein’s mass–energy duality. In gases and liquids, the thermons are supposed to be attached to the molecules or atoms of the medium. In solids, the thermomass gas coincides with the phonon gas for crystals, attached on the electron gas for pure metals, or just both of them for systems in which the heat carriers are phonons and electrons. The physical parameters entering Equation (
8) are [
19]
with the dimensionless parameter
being the Grüneisen constant,
standing for a dimensionless number which is called thermal Mach number of the drift velocity relative to the thermal-wave speed in the heat-carrier collection, and
a characteristic-length vector. In fact, the physical dimensions of
are meters, as it can be directly inferred by the dimensional analysis of Equation (
7). It characterizes the strength of the non-Fourier effects introduced by Equation (
7) and, for conceivable values of
, attains values which are always much smaller than those of the mean-free path of the thermons [
25,
26].
Equation (
7) is non-linear in the heat flux and also accounts for first-order non-local effects through the term
. In [
15,
16,
17,
18,
19], we applied Equation (
7), since it contains the meaningful concept of effective thermal conductivity, accounting for the experimental evidence that the thermal conductivity is not independent of the heat flux. The constitutive equation for the current density is
with
as the electric field, herein regarded as an external force applied to the system [
4].
Finally, Equations (
7) and (
9) must be coupled with the energy-rate equation [
4]
where
u is the specific internal energy, and the quantity
is the rate of energy production due to the circulation of electric current. According to second law of thermodynamics, for such a system, the energy dissipated along the process, locally can be written as
where
denotes the local entropy production.
Our analysis has been carried out under the hypotheses that
, that both
and
depend only on the position on the longitudinal axis
z, and that
and
are parallel. Then, by using Equation (
11), together with the constitutive equations for the heat flux and for the electric current, after some manipulations the local rate of energy dissipated can be written as follows [
16]:
where
and
denote the mean values on the interval
of
and
, respectively, and
is the square root of the mean value of the temperature gradient. By Equation (
12), we infer that the effective thermal conductivity
influences, in a meaningful way, the rate of energy dissipated. It is worth noticing that in deriving the expression (
12) of the energy dissipated we did not use the second Kelvin relation
because when the heat flux is given by Equation (
7), such a relation could no longer be valid, as proved in [
31].
Then, under the assumption that the optimal thermoelectric energy conversion corresponds to the minimum of the function
, we have determined the couples
which minimize
in different situations. Moreover, in correspondence of each minimum, we have obtained the value
of the thermal conductivity. For the sake of illustration, in
Table 1 are summarized the results obtained in Ref. [
18].
To our best knowledge, there are not similar investigations in literature, so that we cannot compare the present results with pre-existing ones. Their experimental confirmation could follow by their possible application to design and manufacturing of thermoelectric energy converters.
1.2. The Present Research
Another parameter which can be externally controlled is the operational temperature, namely, the temperature of the conductor during the process. Thus, for the same system considered in Ref. [
18], it seems important to obtain the couples
which minimize the energy dissipated, once
x is fixed. To achieve that task, we need to recalculate the energy dissipated as function of
c and T, at constant
x. A direct inspection of Equation (
12) suggests that this can be obtained if, and only if, we are able to express
as function of
c and T. On the other hand, since at different temperatures correspond a different phonon scattering, we expect a different behavior of
with respect to the case of constant temperature. Such a hypothesis seems to be confirmed by
Table 2, wherein it is shown the thermal conductivity of a pure Si and pure Ge wire of length
at the temperatures
,
, and
[
18].
We note a marked difference of the values of
for the different temperatures. Thus, a dependency of
on temperature must be taken into account. Since we have no experimental data of
as function of temperature for
wires of length
, as a first step we look for an approximation of it, in the neighborhood of the reference temperatures
,
, and
. In more detail, we look for
in the form
where
,
,
,
,
, and
are suitable temperature-dependent material functions. For
,
,
, and
, we use a series expansion up to the first order of the terms
M,
N,
P, and
Q in the exponents of Equation (
2), with a suitable correction factor to be determined in order to keep the points of minimum of the temperature close to the reference temperature. For
and
, instead, we use the zeroth order approximation given by Equation (
4). We guess that, since the functions
,
,
, and
enter the exponents appearing in the expression of
, even a small variation of them will produce a sensible variation in
, so that we expect a remarkable difference with respect to the isothermal case. This fact is in accordance with our hypothesis that variations in temperature influence remarkably the phonon scattering and, as a consequence, the thermal conductivity.
Thus, around
, for instance, we write
where
are the correction factors we are looking for, and
,
,
, and
,
, are the coefficients of the expansion. The quantities
do not derive by a theoretical calculation, but are chosen empirically in order to obtain values of
which remain limited for T ranging in the neighborhood of the reference temperatures. In future studies, we plan to investigate the possible existence of optimal values of them. Notice that we do not impose the conditions
,
,
, and
, which would guarantee that for
the fit in Equation (
2) is recovered. We are aware that such a strategy could appear unusual at a first look. Our choice is motivated by the experimental evidence that, in materials in which
depends on composition and temperature, there is a different rate of phonon scattering with respect to materials in which
is insensitive to the temperature variation, so that we foresee different values of
for the same composition. Our hypothesis is confirmed by numerical investigations, because the conditions
,
,
, and
do not produce either an acceptable fit or a minimum value of the heat conductivity. Anyway, we underline again that here we do not obtain any fit of
, because we do not have data on its dependency on T for
wires of length
. Herein, we obtain only a first-order approximation of
as function of T, and investigate how such temperature dependency influences the thermoelectric efficiency. Such a situation will be analyzed in more detail in
Section 3. Then, we do not impose any limitation to Equations (
14)–(
17), and look for correction factors which yield a physically acceptable approximation of
. In this way, we obtain a dependency of
on
c which is qualitatively similar to that obtained in [
18], but having different values. In future studies, we aim at exploring how much the numerical error influences such differences, and if we can reduce them by improving our approximation of
.
The procedure illustrated above is also applied for the other two temperatures under consideration, by using different correction factors. In this way, we obtain an approximation of the function which allows to express the rate of energy dissipated as function of c and T. Then, we determine the conditions ensuring the optimal efficiency of the thermoelectric energy conversion by calculating the points of minimum of . The plots of for different temperatures are obtained as well.
The paper runs as follows.
In
Section 2, we apply the approximation procedure described above to obtain the thermal conductivity of a wire of length
as function of
c and T.
In
Section 3, we study the first and second derivatives of
to calculate its minima around the temperatures under consideration. We prove that for each temperature there exists one, and only one, couple
which minimizes
. Moreover, we discuss the results in view of our approximation of
and explain how they can be used in manufacturing thermoelectric energy converters.