<<

arXiv:1502.05697v2 [cond-mat.mtrl-sci] 2 Apr 2016 oeo h lcrceia aueo h thermoelectric the Apertet Y. of nature electrochemical the power on note A edopitrqet to requests offprint Send 2 1 edwti lsdeetia ici aeo w dis- two of made circuit electrical closed a magnetic a within of field appearance the More [3]. described per- latter Seebeck by this precisely, a across simply equilibrium establishing circuit thermal the metallic at results turbing a aiming the in experiments polarization published his magnetic he of 1821, analysis In and Seebeck. to attributed the- [1,2]. of development the irreversible in in useful ories extremely kilo- are sys- that model several tems provides to also orders microwatts thermoelectricity ten Further, from watts. spanning typically range magnitude: a over of pur- powers specific for involving heat designed waste poses be applications from may production promising devices power thermoelectric the electrical as to of thermoelec- field due the in mostly in interest is large systems The tric chem- engineering. physical of physics, and area of active istry, fields very various still covering yet research mature a is Thermoelectricity Introduction 1 3 5 4 wl eisre yteeditor) the by inserted be (will No. manuscript EPJ a usa unu etr 0 oaaSre,Sokv,Mosco Skolkovo, Street, Novaya 100 Center, Quantum Russian France Pont-Audemer, Pr´evert,Lyc´ee F-27500 Jacques F age iatsErn`rs nvri´ eCe Norm Etrang`eres, Caen Universit´e Vivantes de Langues UFR ntttdEetoiu odmnae Universit´e Paris Fondamentale, d’Electronique Institut (LI France Demain Paris, de 75013 Mann, Energies des Interdisciplinaire Laboratoire mi address: Email h icvr ftetemeeti ffc susually is effect thermoelectric the of discovery The olcrcadtemmgei effects thermomagnetic and moelectric PACS. equation. drift-diffusion liter res the the current from in electrical obtained sometimes the non-deg seen of a expression as of phenomenological potential coefficient contact Seebeck s a the we introduce derive clarified, the may are the one the in potentials potential, to electrochemical potentials relevant and turn the chemical then and We power thermoelectr electr determined. the thermoelectric its experimentally of of is definition gradient thermodynamic coefficient the the to tra clarify or heat to potential and whic chemical coefficient, transport system’s charge Seebeck the electric whether between interaction literature the in clear Abstract. date version: Revised / date Received: 1a .Ouerdane H. , 46.bTemeeti,eetoadnmcadohrdir other and electrogasdynamic Thermoelectric, 84.60.Rb [email protected] hl hrolcrctasotter swl established well is theory transport thermoelectric While .Apertet Y. : 23 .Goupil C. , 4 n h Lecoeur Ph. and , Sd NS M 62 -10 ra,France Orsay, F-91405 8622, UMR CNRS, -Sud, D,UR83 nvri´ ai ieo,CR,5ReThomas Rue 5 CNRS, Diderot, Universit´e Paris 8236 UMR ED), ni,Epaaed aPi 43 an France Caen, 14032 Paix la de Esplanade andie, 5 rfiietmeaue lcrceia oeta,, potential, electrochemical zero temperature, at distinc- finite energy Fermi clear or between a associations an- establish relevant or to to tion straightforward discipline always it a electrochemistry, not and is from physics solid-state vary e.g., say, may terminol- other, vari- the conventions as the and Indeed (see, with ogy variations relationship their [10,11,12,13,14,15,16]). refs. its still and e.g., there potentials on yet, thermodynamic consensus of ous clear as experi- However, no materials. first is [6,7,8, of microscopic the and properties from [5] 9] thermodynamic derived both that to ments from extended been as coined been also is has coefficient”. system which “Seebeck the power, across thermoelectric difference the [4]. temperature electro- the difference an and electromotive temperature force this of between the presence indi- coefficient proportionality by the an The in was induced field originated force magnetic it motive the as case effect this rect and work in thermomag- Seebeck’s that reexamined a showed soon as Oersted effect, phenomenon netic observed Seebeck the While interpreted temperatures. materials different these at between maintained junctions were the as materials similar ein132,RsinFederation Russian 143025, region w ligfo hrolcrceet a edirectly be may effects thermoelectric from ulting h ento ftetemeeti opighslater has coupling thermoelectric the of definition The o ht ihapoe osdrto feach of consideration proper a with that, how sot st erltdt h rdeto the of gradient the to related be to is nsport, ceia oeta.Tepeetatceaims article present The potential. ochemical tr.Frhroe edmntaeta the that demonstrate we Furthermore, ature. samaueo h tegho h mutual the of strength the of measure a is h cculn.Frt ercl o h Seebeck the how recall we First, coupling. ic meeti ytm stedfiiin fthe of definitions the As system: rmoelectric nrt eiodco ihu h edto need the without enerate nlsso h eainhpbtenthe between relationship the of analysis c nrycneso 22.aTher- 72.20.Pa – conversion energy ect n ieyapid ti o always not is it applied, widely and 2 Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power

The coefficient αAB, obtained for the whole circuit, is re- lated to the Seebeck coefficient of each material through [17]: αAB = αB αA, (2) − where αA and αB are the Seebeck coefficients of the ma- terials A and B respectively. From an experimental viewpoint, the presence of the material B (= A) is mandatory as it is associated with the probe’s wires6 (see, e.g., ref. [18]). However, if its See- beck coefficient αB is sufficiently small to be neglected, the measurement may be used to determine directly the Seebeck coefficient of material A. In this case, one gets:

V2 V1 αA = − . (3) −T2 T1 − Note the presence of a minus sign in the expression above: It is often overlooked in the literature but, fortunately, Fig. 1. Determination of the Seebeck coefficient for a circuit that omission is most of the time compensated by the ab- composed of two dissimilar materials. sence of a clear sign convention for the measured voltage. Let us now turn to the analysis of the measured quanti- ties. While the temperature is not subject to questioning, Fermi level relative either to the conduction band mini- the voltage obtained from a voltmeter must be defined mum or to the vacuum, and . unambiguously. Indeed, it appears that its connection to In this article, we discuss the definition of the See- the microscopic and thermodynamic properties of mate- beck coefficient focusing particularly on the distinction rials has remained unclear for quite some time, leading between chemical and electrochemical potentials. First, in Riess to publish in 1997, hence fairly recently, an article sect. 2, we address the experimental determination of the untitled “What does a voltmeter measure ?” [19]. In that Seebeck coefficient in order to identify the quantities of in- paper, Riess demonstrated that the voltage measured by terest. Next, the purpose of sect. 3 is to demonstrate that a a voltmeter between two points in a circuit is the differ- clear physical picture of thermoelectric phenomena at the ence of electrochemical potentials µ at the two considered microscopic scale may be obtained on the condition that points divided by the elementary electric charge e, but not the potentials are carefully introduced. For this purpose, the difference between the electrostatice potentials ϕ alone. we review the standard definitions given in the literature The potential V might thus be defined as V = µ/e. This to remove any confusion between the chemical and elec- result is recovered when one measures the voltage− at the trochemical potentials before we present and discuss our ends of a pn junction at equilibrium: While there ise a built- derivation of the Seebeck coefficient for a non-degenerate in electric field associated with the depletion layer, the semiconductor. measured voltage remains zero. The Seebeck coefficient thus appears as a link between the applied temperature difference and the resulting difference of electrochemical 2 Experimental determination of the potential between the two junctions. thermoelectric power The simple technique presented here is not the only one used to determine the thermoelectric power of a given material. Indeed, since the measurement always involves The determination of the Seebeck coefficient traditionally a couple of materials, the absolute Seebeck coefficient of involves components made of dissimilar materials, which the second material has to be known accurately. To obtain we label A and B respectively. The two materials are com- this value, it is possible to use low temperature measure- bined to obtain two junctions as depicted in fig. 1. These ment to reach superconducting state where α = 0 and then junctions are then brought to different temperatures T1 derive higher temperatures values using the Thomson co- and T2. An isothermal voltage measurement at a tem- efficient that can be measured for a single material. For a perature T3, is performed between the free ends of the detailed presentation of the Seebeck coefficient metrology, component B. The voltage thus measured is V2 V1 (this the reader may refer to the instructive review by Martin − notation allows to clearly define a direction for the volt- et al. [20]. age) and the Seebeck coefficient αAB associated with the global system, i.e. the couple AB, is defined as the pro- portionality coefficient between the resulting voltage and 3 Relationship between the thermoelectric the applied temperature difference: power and the electrochemical potential

V2 V1 αAB = − . (1) In order to better understand the influence of each po- T2 T1 tential, we identify the respective effects of temperature − Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power 3 bias, concentration difference, and electric charge, and we discuss the relationship between chemical potential, elec- trochemical potential and the of materials. We then derive the Seebeck coefficient in the simple case of a non-degenerate semiconductor to illustrate the con- tribution of each potential.

3.1 Definition of the thermopower

The Seebeck coefficient may be obtained from a micro- scopic analysis of the considered materials, with the local version of Eq. (3), in open-circuit condition, i.e., with a vanishing electrical current:

µ α = ∇ , (4) e T ∇e where µ and T are respectively the local electrochemical Fig. 2. Energy levels in an n-type semiconductor highlighting potential and temperature, defined at each point of the the notations used in this article (adapted from ref. [8]). The energy EG refers to the bandgap energy. system.e The notation is associated with the gradient of each quantity. In the following,∇ for the sake of simplicity, we consider a unidimensional system so that the spatial gradient reduces to its x-component: . ∇x µ = µ + µe. (5) Note that the quantitiese we just referred to as potentials 3.2 Distinction between the potentials are actually energies. The electrical contribution µe may be expressed as a function of the electrostatic potential ϕ (a genuine potential contrary to µ and µ) so that the Consider a semiconductor sample at thermal equilibrium electrochemical potential reads: and characterized by a spatially inhomogeneous . e As the carrier concentration is nonuniform, a particle cur- µ = µ + qϕ, (6) rent takes place from the region of higher concentration to that of lower concentration: This is the diffusion process where q is the electricale charge of the considered particle. associated with the variation of the carriers’ chemical po- When used in solid state physics, these quantities have tential across the system. This type of electrical current to be related to an energy band diagram. This correspon- is referred to as the diffusion current. The inhomogeneous dence may be found for example in the book of Heikes and electron population in the system thus generates an elec- Ure [8]: Considering the example of an n-doped semicon- tric potential difference and hence a built-in electric field ductor, the electrochemical potential µ corresponds to the which influences the electrons’ motion in such a fashion Fermi level, the electrostatic energy eϕ corresponds to that it tends to curb the diffusion current. The electron the energy level of the bottom of thee− conduction band motion driven by the built-in electric field is the drift cur- while the chemical potential µ corresponds to the dif- rent, which, at thermal equilibrium, exactly cancels the ference between these two quantities and is often called diffusion current, in accordance with the principle of Le These notations are summarized on fig. (2). Chatelier and Braun. In this case, the measured voltage The difference between Fermi level (µ) and Fermi energy across the system always remains zero and there is no (µ) was already highlighted by Wood [9]: “The differ- net electrical current even if the system is short-circuited: ence between the Fermi energy and thee Fermi level should The electric field associated with the electrical potential be noted. The Fermi energy is generally measured from variation is obviously not an electromotive field. However, the adjacent conducting band edge (valence or conduction if the electrons are placed in a non-equilibrium situation band for holes or electrons, respectively), i.e. a reference caused by a thermal bias applied across the system, a non level which may vary in energy, whereas the Fermi level is vanishing may be obtained when the cir- measured from some arbitrary fixed energy level”. This last cuit is closed. This current obviously stems from the un- remark stresses the importance of the choice of an energy compensated contributions of both the diffusion and drift reference, which is a key parameter: To express energies of charge carriers, and it is traditionally related to the in a semiconductor, the bottom of the conduction band gradient of the temperature and to the gradient of the is often used as the reference [10]; however, for studies electrochemical potential. of non-equilibrium phenomena such as thermoelectricity, The electrochemical potential µ of a population of elec- it is mandatory to define an arbitrary fixed energy refer- trically charged particles is the sum of a chemical con- ence independent of the position within the material since tribution µ, the chemical potential,e and of an electrical both µ and µ may vary along the system. It seems the contribution µe [17]: only way to correctly describe the relative displacement e 4 Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power

µ EC µ n(T )= N exp − = N exp , (7)  kBT  kBT  e with

3/2 2πmeff kBT N =2 , (8)  h2 

and where EC is the energy level of the bottom of the conduction band, meff is the electron effective mass, kB is the and h is the Planck constant. The Seebeck coefficient is associated with non-equilibrium phenomena, and, as such, it is tightly linked to transport properties of electrons inside the material. To take account Fig. 3. Schematic illustration (adapted from Ref. [21]) of the of these properties, we build on the drift-diffusion equation variations of the bottom of the conduction band, EC , the top used to obtain the net electrical current density J : of the valence band, EV , the Fermi level, EF , and the vac- x uum level just outside the material, ǫS , all along the circuit J = enM + eD n, (9) depicted in Fig. 1. The slopes of the lines have been greatly x nEx n∇x exaggerated for clarity, and band bending at the interfaces has where Mn and Dn are the electron mobility and diffusivity, been neglected. and where the electric field is related to the energy level Ex EC through: of these energies. Note that the vacuum level infinitely far xEC xEC x = ∇ = ∇ . (10) from the system, E∞, might be a good and meaningful E − q e energy reference. At first, we assume a situation where the electron diffusiv- Figure 3 illustrates the variations of the different en- ity Dn does not depend on the other parameters, including ergies around the circuit depicted in Fig. 1 in the case of the position. The variation of Dn will be discussed further semiconductor materials. It highlights the difference be- below. tween the slope of the bottom of the conduction band and The Seebeck coefficient is obtained setting Jx = 0. the slope of the Fermi level: The variation of the chemical However this current density should be related first to potential thus differs from the variation of the electro- xT and xµ rather than to x and xn. To do so, we chemical potential. Distinguishing these two energies is, evaluate∇ the∇ gradient of the electronE density∇ given by eq. therefore, crucial to properly evaluate the Seebeck coeffi- (7) consideringe that EC, µ and T may vary along the mate- cient. Figure 3 also displays the vacuum level ǫS just out- rial. This approach is seldom found in the literature as one side the material (different from E∞). This vacuum level often sets EC = 0, thus consideringe the bottom of the con- is related to the bottom of the conduction band through duction as the reference everywhere in the nonequilibrium the affinity χ of the material. The discontinuities in ǫS system. As already stressed, this viewpoint is misleading at the interfaces might be seen as contact potentials. On for thermoelectric phenomena. From eq. (7), the gradient the contrary, the Fermi level EF is continuous along the of electron density reads: system, even at the interfaces. Its variation however un- dergoes a sudden change at the interface, reflecting both changes in temperature gradient (assumed constant in a 3 n xT n xn = ∇ + 2 [T ( xµ e x) µ xT ] . (11) given material) and in Seebeck coefficient from a material ∇ 2 T kBT ∇ − E − ∇ to an other. The thermopower is indeed associated with e bulk material but not to interfaces. A similar figure for a We then use this equality along with Einstein’s relation system made of can be found in Ref. [21]. between the electron mobility Mn to the electron diffusiv- ity Dn: M e n = (12) 3.3 From potentials to thermoelectric power: the Dn kBT illustrative case of a non-degenerate semiconductor to modify eq. (9) as follows:

We emphasise the importance of the distinction between xµ kB 3 µ Jx = enMn ∇ + xT . (13) µ and µ on the derivation of the thermoelectric power us-  e e 2 − kBT  ∇  ing the example of a non-degenerate semiconductor doped e withe electrons. In this case, the expression of the carrier Now, setting Jx = 0 and using the definition given in concentration n is rather simple: eq. (4), we find: Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power 5

Scattering mechanism Exponent s in the conduction band and may thus be approximated by Acoustic phonon -1/2 its average value, i.e., 3/2kBT . Replacing τ in eq. (17), we Ionized impurity (strongly screened) -1/2 obtain: Neutral impurity 0 Piezoelectric +1/2 D (D )= n (1 + s) (T ) . (18) Ionized impurity (weakly screened) +3/2 ∇x n T ∇x Table 1. Values of the exponent s for different scattering Finally, inserting eq. (18) and eq. (11) in eq. (16) yields: mechanisms (adapted from Ref. [22]).

xµ kB 5 µ Jx = enMn ∇ + + s xT , (19)  e e 2 − kBT  ∇  kB 3 µ e α = , (14) − e 2 − kBT  and consequently: with a constant electron diffusivity, which is the expected kB 5 µ α = + s . (20) expression for a non-degenerate semiconductor. Further, − e 2 − kBT  this result may also be interpreted by looking at the net thermal energy transported by each carrier transported The contribution of the diffusivity gradient to the thermo- electric power is (1 + s)kB/e and hence depends only on inside the material, i.e., qΠ, where Π is the Peltier coef- − ficient [17]. For electrons, this energy is the scattering parameter s. This term has also been recov- ered by Cai and Mahan [12] using a Boltzmann equation. 3 Note that this term was also introduced by Ioffe [23] with eΠ = kBT µ, (15) − 2 − the notation αD. However, Ioffe used a different power law: He assumed that the carrier’s mean free path l is r since it corresponds to the energy above the Fermi level proportional to (E EC) . He consequently found that µ and hence to the sum of the average thermal energy for − αD = (1/2+ r)kB/e. This discrepancy is quite easy to free electrons and of the energy between the Fermi level − understand since τ is proportional to l/√E EC. ande the bottom of the conduction band, i.e., µ. We thus − recover the second Kelvin relation relating− the Seebeck and Peltier coefficients: Π = αT . 4 Discussion

3.4 Taking into account diffusivity variation 4.1 An unusual derivation While the result given in Eq. (14) is well-known, its deriva- If we relax the assumption of constant diffusivity Dn, this latter becomes a function of the spatial coordinate x and tion presented here is quite original. Indeed, it was di- we end up with the so-called Stratton equation [22]: rectly obtained from the drift-diffusion equation. Thus, the phenomenological equation associated with thermo- electric transport: Jx = enMn x + e x (Dnn) E ∇ xµ = enM + eD (n)+ en (D ) , (16) Jx = σ ∇ σα xT, (21) nEx n∇x ∇x n e − ∇ e It corresponds to a more general form of the drift-diffusion where σ = enMn is the electrical conductivity, is identi- equation, which contains a third contribution to the car- cal to eq. (19) (or to eq. (13) depending on the hypothesis rier motion, directly linked to the gradient of diffusivity made). This latter appears as a modified form of the drift- along the system. To evaluate its effect on the thermo- diffusion equation, which accounts for the couple of vari- electric power, we may reexpress it as a function of the ables [µ, T], or more precisely their gradients, rather than temperature gradient using the relation between the dif- the traditional couple [n, ϕ]. This modification puts forth fusivity Dn and the relaxation time of the carriers τ. Since the facte that the first term of the right hand side of eq. Mn = eτ/meff , the Einstein relation reads: (21) does not correspond any longer to the genuine local τ form of Ohm’s law since it does not involve the electrical Dn = kBT . (17) field x. In this case, the true is given meff by theE gradient of the electrochemical potential as carriers To keep the calculations on an analytical level, we assume experience both diffusion and effects of the electric field. that we deal with low-energy conduction electrons, and The simple derivation of Eq. (19) has been allowed by we express the relaxation time using a power law of the the use as a reference of a fixed energy level, arbitrarily form: τ (E EC)s, where E is the total energy of the chosen but independent of the position along the material, carrier and∝ s −is a characteristic exponent depending on rather than the bottom of the conduction band. This ap- the scattering mechanisms [22]. Some typical values for proach demonstrates that a particular knowledge of both this exponent are given in Table 1. Note that the energy n and is not mandatory to determine the Seebeck ∇x Ex E EC corresponds to the thermal energy of the carriers coefficient of a non-degenerate semiconductor. Indeed, in − 6 Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power this case, these two contributions to the electrochemical i.e., to eq. (14). This latter term is identical to the so-called gradient seem to always compensate each other in such Kelvin formula of the thermopower [13]. As discussed by a way that the resulting electromotive power is indepen- Shastry [26], this contribution “captures the many body dent of specific assumptions, for example a constraint on enhancements, while missing velocity and the carrier concentration. Equation (19) is thus valid for a relaxation contributions”. It thus justifies the introduction wide range of temperatures: It is correct for the extrinsic of the coefficient S to take into account dynamical effects. regime, i.e., when the carrier concentration is fixed by the We believe however that this coefficient should not be pre- concentration of impurities, but it also remains valid in the sented as effective since it does not reflect the appearance freeze-out regime and in the intrinsic regime where addi- of the electromotive force due to the temperature gradi- tional carriers are thermally generated. However, in this ent. It represents only one of the possible contributions latter regime, the contribution to thermo- to this electromotive force. From a practical viewpoint, it electric power should also be considered as these minority has recently been demonstrated that the contribution to carriers may no longer be negligible. One may also refer the thermopower from the diffusivity gradient might be to Ref. [24] in which the authors derive the Seebeck co- significant [27]. efficient focusing only on potentials and the electric field rather than using a statistical approach. 4.4 On the contact potentials

4.2 Link with non-equilibrium thermodynamics Finally, we want to point out the inappropriate use of the contact potentials in the derivation of the thermoelectric While Eq. (21) is widely used in solid-state physics, its power sometimes found in the literature. For example, in formulation is slightly different in non-equilibrium ther- Ref. [23], Ioffe obtains Eq. (19) splitting the Seebeck co- modynamics as general forces are traditionally computed efficient into three separate terms, one being αD while 1 from the gradients µ and (1/T ), instead of µ the two others, αn and αϕ, are associated respectively T ∇x ∇x ∇x and xT [17,25]. So, one should then rewrite Eq. (21) to to concentration gradient and to the “temperature depen- get: ∇ e e dence of the contact potential”. However, as demonstrated later by Chambers [21], contact potentials are irrelevant σT 1 2 1 to thermoelectric effects. This latter term is indeed intro- Jx = xµ + σαT x . (22) e · T ∇ · ∇ T  duced only to compensate the erroneous expression of αn e stemming from the confusion between ϕ and µ/e. From an With this form, it is possible to identify each term with experimental viewpoint, contact potentials are irrelevant the canonical expression [17], since they cannot be probed by a voltmeter: Ase depicted in Fig. 3, these energy discontinuities concern only the bot- tom of the conduction band E (or identically the vacuum Jx 1 1 C JN = = L11 xµ + L12 x , (23) level just outside the material ǫ ) but not the Fermi level − e · T ∇ · ∇ T  S EF . e to recover the expressions of the kinetic coefficients in the 2 2 thermoelectric case, i.e., L11 = σT/e and L12 = σαT /e. 5 Conclusion

4.3 On the so-called effective Seebeck coefficient In this article, we have discussed the definition of the thermoelectric power with a special emphasis on its re- Let us now turn to the previous analysis of the thermo- lationship to the electrochemical potential. A proper con- electric power in non-degenerate semiconductor. In ref. sideration of all potentials inside the material has led to [11], Mahan introduces an effective Seebeck coefficient S, demonstrate that the phenomenological equation for the distinct from the genuine thermoelectric power eq. (3) ob- electrical current involving thermoelectric coefficients may tained from measurements. In a subsequent article with be derived directly from the drift-diffusion equation. We Cai [12], this effective coefficient is presented as the ra- also shed light on the physical interpretation of the effec- tio between the electric field and the temperature gradi- tive Seebeck coefficient defined by Mahan, showing that it ent. These two different Seebeck coefficients are related is actually related to the gradient of diffusivity along the through the following relation [12]: system.

1 ∂µ α = S + , (24) e ∂T n References A comparison of eq. (16) with the equation (20) of ref. 1. S. R. de Groot Thermodynamics of Irreversible Processes [11] leads to identify the effective Seebeck coefficient to (Interscience Publishers Inc., New York, 1958). the contribution of the diffusivity gradient, i.e., αD. The 2. Y. Apertet, H. Ouerdane, C. Goupil, and Ph. Lecoeur, second term of the right hand side of eq. (24) should then Phys. Rev. E 85, 031116 (2012). be associated with the assumption of constant diffusivity, 3. T. J. Seebeck, Abh. K. Akad. Wiss. Berlin 289 (1821). Y. Apertet et al.: A note on the electrochemical nature of the thermoelectric power 7

4. H. C. Oersted, Annales de Chimie et de Physique 22, 199 (1823). 5. H. B. Callen, Phys. Rev. 73, 1349 (1948). 6. C. Herring, Phys. Rev. 96, 1163 (1954). 7. P. J. Price, Phys. Rev. 104, 1223 (1956). 8. R. R. Heikes and R. W. Ure, Thermoelectricity: Science and Engineering (Interscience Publishers Inc., New York, 1961). 9. C. Wood, Reports on Progress in Physics 51, 459 (1988). 10. C. Kittel, Introduction to Solid State Physics, 8th edition (John Wiley & Sons, New York, 2005). 11. G. D. Mahan, J. Appl. Phys. 87, 7326 (2000). 12. J. Cai and G. D. Mahan, Phys. Rev. B 74, 075201 (2006). 13. M. R. Peterson and B. S. Shastry, Phys. Rev. B 82, 195105 (2010). 14. A. A. Varlamov and K. V. Kavokin, EPL 103, 47005 (2013). 15. H. Ouerdane, A. A. Varlamov, A. V. Kavokin, C. Goupil, and C. B. Vining, Phys. Rev. B 91, 100501(R) (2015). 16. K. Behnia, Fundamentals of Thermoelectricity (Oxford University Press, Oxford,2015). 17. H. B. Callen, Thermodynamics and an introduction to Thermostatistics (John Wiley & Sons, New York, 1985). 18. Z. Zhou and C. Uher, Rev. Sci. Instrum. 76, 023901 (2005). 19. I. Riess. Solid State Ionics 95, 327 (1997). 20. J. Martin, T. Tritt, and C. Uher, J. Appl. Phys. 108, 121101 (2010); and references therein. 21. R. G. Chambers, Physics Education 12, 374 (1977). 22. M. Lundstrom, Fundamentals of carrier transport, second edition (Cambridge University Press, Cambridge, 2009). 23. A. F. Ioffe, Physics of (Infosearch, Ltd., London, 1960). 24. F. W. G. Rose, E. Billig, and J. E. Parrott, J. Electron. and Control 3, 481 (1957). 25. C. Goupil, W. Seifert, K. Zabrocki, E. Muller, and G. J. Snyder, Entropy 13, 1481-1517 (2011). 26. B. S. Shastry, Thermopower in Correlated Systems, in New Materials for Thermoelectric Applications: Theory and Ex- periment (Springer Netherlands, Dordrecht, 2013). 27. P. Sun, B. Wei, J. Zhang, J. M. Tomczak, A. M. Strydom, M. Sondergaard, B. B. Iversen, and F. Steglich, Nature Communications 6, 7475 (2015).