fbpx
Wikipedia

Seebeck coefficient

The Seebeck coefficient (also known as thermopower,[1] thermoelectric power, and thermoelectric sensitivity) of a material is a measure of the magnitude of an induced thermoelectric voltage in response to a temperature difference across that material, as induced by the Seebeck effect.[2] The SI unit of the Seebeck coefficient is volts per kelvin (V/K),[2] although it is more often given in microvolts per kelvin (μV/K).

The use of materials with a high Seebeck coefficient[3] is one of many important factors for the efficient behaviour of thermoelectric generators and thermoelectric coolers. More information about high-performance thermoelectric materials can be found in the Thermoelectric materials article. In thermocouples the Seebeck effect is used to measure temperatures, and for accuracy it is desirable to use materials with a Seebeck coefficient that is stable over time.

Physically, the magnitude and sign of the Seebeck coefficient can be approximately understood as being given by the entropy per unit charge carried by electrical currents in the material. It may be positive or negative. In conductors that can be understood in terms of independently moving, nearly-free charge carriers, the Seebeck coefficient is negative for negatively charged carriers (such as electrons), and positive for positively charged carriers (such as electron holes).

Definition edit

One way to define the Seebeck coefficient is the voltage built up when a small temperature gradient is applied to a material, and when the material has come to a steady state where the current density is zero everywhere. If the temperature difference ΔT between the two ends of a material is small, then the Seebeck coefficient of a material is defined as:

 

where ΔV is the thermoelectric voltage seen at the terminals. (See below for more on the signs of ΔV and ΔT.)

Note that the voltage shift expressed by the Seebeck effect cannot be measured directly, since the measured voltage (by attaching a voltmeter) contains an additional voltage contribution, due to the temperature gradient and Seebeck effect in the measurement leads. The voltmeter voltage is always dependent on relative Seebeck coefficients among the various materials involved.

Most generally and technically, the Seebeck coefficient is defined in terms of the portion of electric current driven by temperature gradients, as in the vector differential equation

 

where   is the current density,   is the electrical conductivity,   is the voltage gradient, and   is the temperature gradient. The zero-current, steady state special case described above has  , which implies that the two electrical conductivity terms have cancelled out and so  

Sign convention edit

The sign is made explicit in the following expression:

 

Thus, if S is positive, the end with the higher temperature has the lower voltage, and vice versa. The voltage gradient in the material will point against the temperature gradient.

The Seebeck effect is generally dominated by the contribution from charge carrier diffusion (see below) which tends to push charge carriers towards the cold side of the material until a compensating voltage has built up. As a result, in p-type semiconductors (which have only positive mobile charges, electron holes), S is positive. Likewise, in n-type semiconductors (which have only negative mobile charges, electrons), S is negative. In most conductors, however, the charge carriers exhibit both hole-like and electron-like behaviour and the sign of S usually depends on which of them predominates.

Relationship to other thermoelectric coefficients edit

According to the second Thomson relation (which holds for all non-magnetic materials in the absence of an externally applied magnetic field), the Seebeck coefficient is related to the Peltier coefficient   by the exact relation

 

where   is the thermodynamic temperature.

According to the first Thomson relation and under the same assumptions about magnetism, the Seebeck coefficient is related to the Thomson coefficient   by

 

The constant of integration is such that   at absolute zero, as required by Nernst's theorem.

Measurement edit

Relative Seebeck coefficient edit

In practice the absolute Seebeck coefficient is difficult to measure directly, since the voltage output of a thermoelectric circuit, as measured by a voltmeter, only depends on differences of Seebeck coefficients. This is because electrodes attached to a voltmeter must be placed onto the material in order to measure the thermoelectric voltage. The temperature gradient then also typically induces a thermoelectric voltage across one leg of the measurement electrodes. Therefore, the measured Seebeck coefficient is a contribution from the Seebeck coefficient of the material of interest and the material of the measurement electrodes. This arrangement of two materials is usually called a thermocouple.

The measured Seebeck coefficient is then a contribution from both and can be written as:

 

Absolute Seebeck coefficient edit

 
Absolute Seebeck coefficient of lead at low temperature, according to Christian, Jan, Pearson, Templeton (1958). Below the critical temperature of lead (indicated by the dashed line, approximately 7 K) the lead is superconducting.
 
Absolute Seebeck coefficients of various metals up to high temperatures, mainly from Cusack & Kendall (1958). The data for lead (Pb) is from Christian, Jan, Pearson, Templeton (1958).

Although only relative Seebeck coefficients are important for externally measured voltages, the absolute Seebeck coefficient can be important for other effects where voltage is measured indirectly. Determination of the absolute Seebeck coefficient therefore requires more complicated techniques and is more difficult, but such measurements have been performed on standard materials. These measurements only had to be performed once for all time, and for all materials; for any other material, the absolute Seebeck coefficient can be obtained by performing a relative Seebeck coefficient measurement against a standard material.

A measurement of the Thomson coefficient  , which expresses the strength of the Thomson effect, can be used to yield the absolute Seebeck coefficient through the relation:  , provided that   is measured down to absolute zero. The reason this works is that   is expected to decrease to zero as the temperature is brought to zero—a consequence of Nernst's theorem. Such a measurement based on the integration of   was published in 1932,[4] though it relied on the interpolation of the Thomson coefficient in certain regions of temperature.

Superconductors have zero Seebeck coefficient, as mentioned below. By making one of the wires in a thermocouple superconducting, it is possible to get a direct measurement of the absolute Seebeck coefficient of the other wire, since it alone determines the measured voltage from the entire thermocouple. A publication in 1958 used this technique to measure the absolute Seebeck coefficient of lead between 7.2 K and 18 K, thereby filling in an important gap in the previous 1932 experiment mentioned above.[5]

The combination of the superconductor-thermocouple technique up to 18 K, with the Thomson-coefficient-integration technique above 18 K, allowed determination of the absolute Seebeck coefficient of lead up to room temperature. By proxy, these measurements led to the determination of absolute Seebeck coefficients for all materials, even up to higher temperatures, by a combination of Thomson coefficient integrations and thermocouple circuits.[6]

The difficulty of these measurements, and the rarity of reproducing experiments, lends some degree of uncertainty to the absolute thermoelectric scale thus obtained. In particular, the 1932 measurements may have incorrectly measured the Thomson coefficient over the range 20 K to 50 K. Since nearly all subsequent publications relied on those measurements, this would mean that all of the commonly used values of absolute Seebeck coefficient (including those shown in the figures) are too low by about 0.3 μV/K, for all temperatures above 50 K.[7]

Seebeck coefficients for some common materials edit

In the table below are Seebeck coefficients at room temperature for some common, nonexotic materials, measured relative to platinum.[8] The Seebeck coefficient of platinum itself is approximately −5 μV/K at room temperature,[9] and so the values listed below should be compensated accordingly. For example, the Seebeck coefficients of Cu, Ag, Au are 1.5 μV/K, and of Al −1.5 μV/K. The Seebeck coefficient of semiconductors very much depends on doping, with generally positive values for p doped materials and negative values for n doping.

Material Seebeck coefficient
relative to platinum (μV/K)
Selenium 900
Tellurium 500
Silicon 440
Germanium 330
Antimony 47
Nichrome 25
Iron 19
Molybdenum 10
Cadmium, tungsten 7.5
Gold, silver, copper 6.5
Rhodium 6.0
Tantalum 4.5
Lead 4.0
Aluminium 3.5
Carbon 3.0
Mercury 0.6
Platinum 0 (definition)
Sodium -2.0
Potassium -9.0
Nickel -15
Constantan -35
Bismuth -72

Physical factors that determine the Seebeck coefficient edit

A material's temperature, crystal structure, and impurities influence the value of thermoelectric coefficients. The Seebeck effect can be attributed to two things:[10] charge-carrier diffusion and phonon drag.

Charge carrier diffusion edit

On a fundamental level, an applied voltage difference refers to a difference in the thermodynamic chemical potential of charge carriers, and the direction of the current under a voltage difference is determined by the universal thermodynamic process in which (given equal temperatures) particles flow from high chemical potential to low chemical potential. In other words, the direction of the current in Ohm's law is determined via the thermodynamic arrow of time (the difference in chemical potential could be exploited to produce work, but is instead dissipated as heat which increases entropy). On the other hand, for the Seebeck effect not even the sign of the current can be predicted from thermodynamics, and so to understand the origin of the Seebeck coefficient it is necessary to understand the microscopic physics.

Charge carriers (such as thermally excited electrons) constantly diffuse around inside a conductive material. Due to thermal fluctuations, some of these charge carriers travel with a higher energy than average, and some with a lower energy. When no voltage differences or temperature differences are applied, the carrier diffusion perfectly balances out and so on average one sees no current:  . A net current can be generated by applying a voltage difference (Ohm's law), or by applying a temperature difference (Seebeck effect). To understand the microscopic origin of the thermoelectric effect, it is useful to first describe the microscopic mechanism of the normal Ohm's law electrical conductance—to describe what determines the   in  . Microscopically, what is happening in Ohm's law is that higher energy levels have a higher concentration of carriers per state, on the side with higher chemical potential. For each interval of energy, the carriers tend to diffuse and spread into the area of device where there are fewer carriers per state of that energy. As they move, however, they occasionally scatter dissipatively, which re-randomizes their energy according to the local temperature and chemical potential. This dissipation empties out the carriers from these higher energy states, allowing more to diffuse in. The combination of diffusion and dissipation favours an overall drift of the charge carriers towards the side of the material where they have a lower chemical potential.[11]: Ch.11 

For the thermoelectric effect, now, consider the case of uniform voltage (uniform chemical potential) with a temperature gradient. In this case, at the hotter side of the material there is more variation in the energies of the charge carriers, compared to the colder side. This means that high energy levels have a higher carrier occupation per state on the hotter side, but also the hotter side has a lower occupation per state at lower energy levels. As before, the high-energy carriers diffuse away from the hot end, and produce entropy by drifting towards the cold end of the device. However, there is a competing process: at the same time low-energy carriers are drawn back towards the hot end of the device. Though these processes both generate entropy, they work against each other in terms of charge current, and so a net current only occurs if one of these drifts is stronger than the other. The net current is given by  , where (as shown below) the thermoelectric coefficient   depends literally on how conductive high-energy carriers are, compared to low-energy carriers. The distinction may be due to a difference in rate of scattering, a difference in speeds, a difference in density of states, or a combination of these effects.

Mott formula edit

The processes described above apply in materials where each charge carrier sees an essentially static environment so that its motion can be described independently from other carriers, and independent of other dynamics (such as phonons). In particular, in electronic materials with weak electron-electron interactions, weak electron-phonon interactions, etc. it can be shown in general that the linear response conductance is

 

and the linear response thermoelectric coefficient is

 

where   is the energy-dependent conductivity, and   is the Fermi–Dirac distribution function. These equations are known as the Mott relations, of Sir Nevill Francis Mott.[12] The derivative

 

is a function peaked around the chemical potential (Fermi level)   with a width of approximately  . The energy-dependent conductivity (a quantity that cannot actually be directly measured — one only measures  ) is calculated as   where   is the electron diffusion constant and   is the electronic density of states (in general, both are functions of energy).

In materials with strong interactions, none of the above equations can be used since it is not possible to consider each charge carrier as a separate entity. The Wiedemann–Franz law can also be exactly derived using the non-interacting electron picture, and so in materials where the Wiedemann–Franz law fails (such as superconductors), the Mott relations also generally tend to fail.[13]

The formulae above can be simplified in a couple of important limiting cases:

Mott formula in metals edit

In semimetals and metals, where transport only occurs near the Fermi level and   changes slowly in the range  , one can perform a Sommerfeld expansion  , which leads to

 

This expression is sometimes called "the Mott formula", however it is much less general than Mott's original formula expressed above.

In the free electron model with scattering, the value of   is of order  , where   is the Fermi temperature, and so a typical value of the Seebeck coefficient in the Fermi gas is   (the prefactor varies somewhat depending on details such as dimensionality and scattering). In highly conductive metals the Fermi temperatures are typically around 104 – 105 K, and so it is understandable why their absolute Seebeck coefficients are only of order 1 – 10 μV/K at room temperature. Note that whereas the free electron model predicts a negative Seebeck coefficient, real metals actually have complicated band structures and may exhibit positive Seebeck coefficients (examples: Cu, Ag, Au).

The fraction   in semimetals is sometimes calculated from the measured derivative of   with respect to some energy shift induced by field effect. This is not necessarily correct and the estimate of   can be incorrect (by a factor of two or more), since the disorder potential depends on screening which also changes with field effect.[14]

Mott formula in semiconductors edit

In semiconductors at low levels of doping, transport only occurs far away from the Fermi level. At low doping in the conduction band (where  , where   is the minimum energy of the conduction band edge), one has  . Approximating the conduction band levels' conductivity function as   for some constants   and  ,

 

whereas in the valence band when   and  ,

 

The values of   and   depend on material details; in bulk semiconductor these constants range between 1 and 3, the extremes corresponding to acoustic-mode lattice scattering and ionized-impurity scattering.[15]

 
Seebeck coefficient of silicon at 300 K, calculated from Mott model. The crossover from hole-dominated conduction (positive  ) to electron-dominated conduction (negative  ) happens for Fermi levels in the middle of the 1.1 eV-wide gap.

In extrinsic (doped) semiconductors either the conduction or valence band will dominate transport, and so one of the numbers above will give the measured values. In general however the semiconductor may also be intrinsic in which case the bands conduct in parallel, and so the measured values will be

 

This results in a crossover behaviour, as shown in the figure. The highest Seebeck coefficient is obtained when the semiconductor is lightly doped, however a high Seebeck coefficient is not necessarily useful on its own. For thermoelectric power devices (coolers, generators) it is more important to maximize the thermoelectric power factor  ,[16] or the thermoelectric figure of merit, and the optimum generally occurs at high doping levels.[17]

Phonon drag edit

Phonons are not always in local thermal equilibrium; they move against the thermal gradient. They lose momentum by interacting with electrons (or other carriers) and imperfections in the crystal. If the phonon-electron interaction is predominant, the phonons will tend to push the electrons to one end of the material, hence losing momentum and contributing to the thermoelectric field. This contribution is most important in the temperature region where phonon-electron scattering is predominant. This happens for

 

where   is the Debye temperature. At lower temperatures there are fewer phonons available for drag, and at higher temperatures they tend to lose momentum in phonon-phonon scattering instead of phonon-electron scattering. At lower temperatures, material boundaries also play an increasing role as the phonons can travel significant distances.[18] Practically speaking, phonon drag is an important effect in semiconductors near room temperature (even though well above  ), that is comparable in magnitude to the carrier-diffusion effect described in the previous section.[18]

This region of the thermopower-versus-temperature function is highly variable under a magnetic field.[citation needed]

Relationship with entropy edit

The Seebeck coefficient of a material corresponds thermodynamically to the amount of entropy "dragged along" by the flow of charge inside a material; it is in some sense the entropy per unit charge in the material.[19]

References edit

  1. ^ Thermopower is a misnomer as this quantity does not actually express a power quantity: Note that the unit of thermopower (V/K) is different from the unit of power (watts).
  2. ^ a b Blundell, Stephen; Blundell, Stephen J.; Blundell, Katherine M. (2010). Concepts in Thermal Physics. Oxford University Press. p. 415. ISBN 978-0-19-956210-7.
  3. ^ Joseph R. Sootsman; Duck Young Chung; Mercouri G. Kanatzidis (2009). "New and Old Concepts in Thermoelectric Materials". Angewandte Chemie. 48 (46): 8616–8639. doi:10.1002/anie.200900598. PMID 19866458.
  4. ^ Borelius, G.; Keesom, W. H.; Johannson, C. H.; Linde, J. O. (1932). "Establishment of an Absolute Scale for the Thermo-electric Force". Proceedings of the Royal Academy of Sciences at Amsterdam. 35 (1): 10.
  5. ^ Christian, J. W.; Jan, J.-P.; Pearson, W. B.; Templeton, I. M. (1958). "Thermo-Electricity at Low Temperatures. VI. A Redetermination of the Absolute Scale of Thermo-Electric Power of Lead". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 245 (1241): 213. Bibcode:1958RSPSA.245..213C. doi:10.1098/rspa.1958.0078. S2CID 96708128.
  6. ^ Cusack, N.; Kendall, P. (1958). "The Absolute Scale of Thermoelectric Power at High Temperature". Proceedings of the Physical Society. 72 (5): 898. Bibcode:1958PPS....72..898C. doi:10.1088/0370-1328/72/5/429.
  7. ^ Roberts, R. B. (1986). "Absolute scales for thermoelectricity". Measurement. 4 (3): 101–103. Bibcode:1986Meas....4..101R. doi:10.1016/0263-2241(86)90016-3.
  8. ^ The Seebeck Coefficient, Electronics Cooling.com (accessed 2013-Feb-01)
  9. ^ Moore, J. P. (1973). "Absolute Seebeck coefficient of platinum from 80 to 340 K and the thermal and electrical conductivities of lead from 80 to 400 K". Journal of Applied Physics. 44 (3): 1174–1178. Bibcode:1973JAP....44.1174M. doi:10.1063/1.1662324.
  10. ^ Kong, Ling Bing (2014). Waste Energy Harvesting. Lecture Notes in Energy. Vol. 24. Springer. pp. 263–403. doi:10.1007/978-3-642-54634-1. ISBN 978-3-642-54634-1.
  11. ^ Datta, Supriyo (2005). Quantum Transport: Atom to Transistor. Cambridge University Press. ISBN 978-0-521-63145-7.
  12. ^ Cutler, M.; Mott, N. (1969). "Observation of Anderson Localization in an Electron Gas". Physical Review. 181 (3): 1336. Bibcode:1969PhRv..181.1336C. doi:10.1103/PhysRev.181.1336.
  13. ^ Jonson, M.; Mahan, G. (1980). "Mott's formula for the thermopower and the Wiedemann-Franz law". Physical Review B. 21 (10): 4223. Bibcode:1980PhRvB..21.4223J. doi:10.1103/PhysRevB.21.4223.
  14. ^ Hwang, E. H.; Rossi, E.; Das Sarma, S. (2009). "Theory of thermopower in two-dimensional graphene". Physical Review B. 80 (23): 235415. arXiv:0902.1749. Bibcode:2009PhRvB..80w5415H. doi:10.1103/PhysRevB.80.235415. S2CID 8125966.
  15. ^ Semiconductor Physics: An Introduction, Karlheinz Seeger
  16. ^ Imai, H.; Shimakawa, Y.; Kubo, Y. (10 December 2001). "Large thermoelectric power factor in TiS2 crystal with nearly stoichiometric composition". Physical Review B. 64 (24): 241104. arXiv:cond-mat/0111063. Bibcode:2001PhRvB..64x1104I. doi:10.1103/PhysRevB.64.241104. S2CID 119389373.
  17. ^ G. Jeffrey Snyder, "Thermoelectrics". http://www.its.caltech.edu/~jsnyder/thermoelectrics/
  18. ^ a b Mahan, G. D.; Lindsay, L.; Broido, D. A. (28 December 2014). "The Seebeck coefficient and phonon drag in silicon". Journal of Applied Physics. 116 (24): 245102. Bibcode:2014JAP...116x5102M. doi:10.1063/1.4904925. OSTI 1185754.
  19. ^ Bulusu, A.; Walker, D. G. (2008). "Review of electronic transport models for thermoelectric materials". Superlattices and Microstructures. 44 (1): 1. Bibcode:2008SuMi...44....1B. doi:10.1016/j.spmi.2008.02.008.

seebeck, coefficient, also, known, thermopower, thermoelectric, power, thermoelectric, sensitivity, material, measure, magnitude, induced, thermoelectric, voltage, response, temperature, difference, across, that, material, induced, seebeck, effect, unit, volts. The Seebeck coefficient also known as thermopower 1 thermoelectric power and thermoelectric sensitivity of a material is a measure of the magnitude of an induced thermoelectric voltage in response to a temperature difference across that material as induced by the Seebeck effect 2 The SI unit of the Seebeck coefficient is volts per kelvin V K 2 although it is more often given in microvolts per kelvin mV K The use of materials with a high Seebeck coefficient 3 is one of many important factors for the efficient behaviour of thermoelectric generators and thermoelectric coolers More information about high performance thermoelectric materials can be found in the Thermoelectric materials article In thermocouples the Seebeck effect is used to measure temperatures and for accuracy it is desirable to use materials with a Seebeck coefficient that is stable over time Physically the magnitude and sign of the Seebeck coefficient can be approximately understood as being given by the entropy per unit charge carried by electrical currents in the material It may be positive or negative In conductors that can be understood in terms of independently moving nearly free charge carriers the Seebeck coefficient is negative for negatively charged carriers such as electrons and positive for positively charged carriers such as electron holes Contents 1 Definition 1 1 Sign convention 1 2 Relationship to other thermoelectric coefficients 2 Measurement 2 1 Relative Seebeck coefficient 2 2 Absolute Seebeck coefficient 3 Seebeck coefficients for some common materials 4 Physical factors that determine the Seebeck coefficient 4 1 Charge carrier diffusion 4 1 1 Mott formula 4 1 1 1 Mott formula in metals 4 1 1 2 Mott formula in semiconductors 4 2 Phonon drag 4 3 Relationship with entropy 5 ReferencesDefinition editMain article Seebeck effect One way to define the Seebeck coefficient is the voltage built up when a small temperature gradient is applied to a material and when the material has come to a steady state where the current density is zero everywhere If the temperature difference DT between the two ends of a material is small then the Seebeck coefficient of a material is defined as S D V D T displaystyle S Delta V over Delta T nbsp where DV is the thermoelectric voltage seen at the terminals See below for more on the signs of DV and DT Note that the voltage shift expressed by the Seebeck effect cannot be measured directly since the measured voltage by attaching a voltmeter contains an additional voltage contribution due to the temperature gradient and Seebeck effect in the measurement leads The voltmeter voltage is always dependent on relative Seebeck coefficients among the various materials involved Most generally and technically the Seebeck coefficient is defined in terms of the portion of electric current driven by temperature gradients as in the vector differential equation J s V s S T displaystyle mathbf J sigma boldsymbol nabla V sigma S boldsymbol nabla T nbsp where J displaystyle scriptstyle mathbf J nbsp is the current density s displaystyle scriptstyle sigma nbsp is the electrical conductivity V displaystyle scriptstyle boldsymbol nabla V nbsp is the voltage gradient and T displaystyle scriptstyle boldsymbol nabla T nbsp is the temperature gradient The zero current steady state special case described above has J 0 displaystyle scriptstyle mathbf J 0 nbsp which implies that the two electrical conductivity terms have cancelled out and so V S T displaystyle boldsymbol nabla V S boldsymbol nabla T nbsp Sign convention edit The sign is made explicit in the following expression S V l e f t V r i g h t T l e f t T r i g h t displaystyle S frac V rm left V rm right T rm left T rm right nbsp Thus if S is positive the end with the higher temperature has the lower voltage and vice versa The voltage gradient in the material will point against the temperature gradient The Seebeck effect is generally dominated by the contribution from charge carrier diffusion see below which tends to push charge carriers towards the cold side of the material until a compensating voltage has built up As a result in p type semiconductors which have only positive mobile charges electron holes S is positive Likewise in n type semiconductors which have only negative mobile charges electrons S is negative In most conductors however the charge carriers exhibit both hole like and electron like behaviour and the sign of S usually depends on which of them predominates Relationship to other thermoelectric coefficients edit Main article Thomson relations According to the second Thomson relation which holds for all non magnetic materials in the absence of an externally applied magnetic field the Seebeck coefficient is related to the Peltier coefficient P displaystyle scriptstyle Pi nbsp by the exact relation S P T displaystyle S frac Pi T nbsp where T displaystyle T nbsp is the thermodynamic temperature According to the first Thomson relation and under the same assumptions about magnetism the Seebeck coefficient is related to the Thomson coefficient K displaystyle scriptstyle mathcal K nbsp by S K T d T displaystyle S int frac mathcal K T dT nbsp The constant of integration is such that S 0 displaystyle scriptstyle S 0 nbsp at absolute zero as required by Nernst s theorem Measurement editRelative Seebeck coefficient edit See also Thermocouple In practice the absolute Seebeck coefficient is difficult to measure directly since the voltage output of a thermoelectric circuit as measured by a voltmeter only depends on differences of Seebeck coefficients This is because electrodes attached to a voltmeter must be placed onto the material in order to measure the thermoelectric voltage The temperature gradient then also typically induces a thermoelectric voltage across one leg of the measurement electrodes Therefore the measured Seebeck coefficient is a contribution from the Seebeck coefficient of the material of interest and the material of the measurement electrodes This arrangement of two materials is usually called a thermocouple The measured Seebeck coefficient is then a contribution from both and can be written as S A B S B S A D V B D T D V A D T displaystyle S AB S B S A Delta V B over Delta T Delta V A over Delta T nbsp Absolute Seebeck coefficient edit nbsp Absolute Seebeck coefficient of lead at low temperature according to Christian Jan Pearson Templeton 1958 Below the critical temperature of lead indicated by the dashed line approximately 7 K the lead is superconducting nbsp Absolute Seebeck coefficients of various metals up to high temperatures mainly from Cusack amp Kendall 1958 The data for lead Pb is from Christian Jan Pearson Templeton 1958 Although only relative Seebeck coefficients are important for externally measured voltages the absolute Seebeck coefficient can be important for other effects where voltage is measured indirectly Determination of the absolute Seebeck coefficient therefore requires more complicated techniques and is more difficult but such measurements have been performed on standard materials These measurements only had to be performed once for all time and for all materials for any other material the absolute Seebeck coefficient can be obtained by performing a relative Seebeck coefficient measurement against a standard material A measurement of the Thomson coefficient K displaystyle mathcal K nbsp which expresses the strength of the Thomson effect can be used to yield the absolute Seebeck coefficient through the relation S T 0 T K T T d T displaystyle S T int 0 T mathcal K T over T dT nbsp provided that K displaystyle mathcal K nbsp is measured down to absolute zero The reason this works is that S T displaystyle S T nbsp is expected to decrease to zero as the temperature is brought to zero a consequence of Nernst s theorem Such a measurement based on the integration of K T displaystyle mathcal K T nbsp was published in 1932 4 though it relied on the interpolation of the Thomson coefficient in certain regions of temperature Superconductors have zero Seebeck coefficient as mentioned below By making one of the wires in a thermocouple superconducting it is possible to get a direct measurement of the absolute Seebeck coefficient of the other wire since it alone determines the measured voltage from the entire thermocouple A publication in 1958 used this technique to measure the absolute Seebeck coefficient of lead between 7 2 K and 18 K thereby filling in an important gap in the previous 1932 experiment mentioned above 5 The combination of the superconductor thermocouple technique up to 18 K with the Thomson coefficient integration technique above 18 K allowed determination of the absolute Seebeck coefficient of lead up to room temperature By proxy these measurements led to the determination of absolute Seebeck coefficients for all materials even up to higher temperatures by a combination of Thomson coefficient integrations and thermocouple circuits 6 The difficulty of these measurements and the rarity of reproducing experiments lends some degree of uncertainty to the absolute thermoelectric scale thus obtained In particular the 1932 measurements may have incorrectly measured the Thomson coefficient over the range 20 K to 50 K Since nearly all subsequent publications relied on those measurements this would mean that all of the commonly used values of absolute Seebeck coefficient including those shown in the figures are too low by about 0 3 mV K for all temperatures above 50 K 7 Seebeck coefficients for some common materials editFor the thermoelectric properties of high performance thermoelectric materials see Thermoelectric materials In the table below are Seebeck coefficients at room temperature for some common nonexotic materials measured relative to platinum 8 The Seebeck coefficient of platinum itself is approximately 5 mV K at room temperature 9 and so the values listed below should be compensated accordingly For example the Seebeck coefficients of Cu Ag Au are 1 5 mV K and of Al 1 5 mV K The Seebeck coefficient of semiconductors very much depends on doping with generally positive values for p doped materials and negative values for n doping Material Seebeck coefficient relative to platinum mV K Selenium 900 Tellurium 500 Silicon 440 Germanium 330 Antimony 47 Nichrome 25 Iron 19 Molybdenum 10 Cadmium tungsten 7 5 Gold silver copper 6 5 Rhodium 6 0 Tantalum 4 5 Lead 4 0 Aluminium 3 5 Carbon 3 0 Mercury 0 6 Platinum 0 definition Sodium 2 0 Potassium 9 0 Nickel 15 Constantan 35 Bismuth 72Physical factors that determine the Seebeck coefficient editA material s temperature crystal structure and impurities influence the value of thermoelectric coefficients The Seebeck effect can be attributed to two things 10 charge carrier diffusion and phonon drag Charge carrier diffusion edit On a fundamental level an applied voltage difference refers to a difference in the thermodynamic chemical potential of charge carriers and the direction of the current under a voltage difference is determined by the universal thermodynamic process in which given equal temperatures particles flow from high chemical potential to low chemical potential In other words the direction of the current in Ohm s law is determined via the thermodynamic arrow of time the difference in chemical potential could be exploited to produce work but is instead dissipated as heat which increases entropy On the other hand for the Seebeck effect not even the sign of the current can be predicted from thermodynamics and so to understand the origin of the Seebeck coefficient it is necessary to understand the microscopic physics Charge carriers such as thermally excited electrons constantly diffuse around inside a conductive material Due to thermal fluctuations some of these charge carriers travel with a higher energy than average and some with a lower energy When no voltage differences or temperature differences are applied the carrier diffusion perfectly balances out and so on average one sees no current J 0 displaystyle scriptstyle mathbf J 0 nbsp A net current can be generated by applying a voltage difference Ohm s law or by applying a temperature difference Seebeck effect To understand the microscopic origin of the thermoelectric effect it is useful to first describe the microscopic mechanism of the normal Ohm s law electrical conductance to describe what determines the s displaystyle scriptstyle sigma nbsp in J s V displaystyle scriptstyle mathbf J sigma boldsymbol nabla V nbsp Microscopically what is happening in Ohm s law is that higher energy levels have a higher concentration of carriers per state on the side with higher chemical potential For each interval of energy the carriers tend to diffuse and spread into the area of device where there are fewer carriers per state of that energy As they move however they occasionally scatter dissipatively which re randomizes their energy according to the local temperature and chemical potential This dissipation empties out the carriers from these higher energy states allowing more to diffuse in The combination of diffusion and dissipation favours an overall drift of the charge carriers towards the side of the material where they have a lower chemical potential 11 Ch 11 For the thermoelectric effect now consider the case of uniform voltage uniform chemical potential with a temperature gradient In this case at the hotter side of the material there is more variation in the energies of the charge carriers compared to the colder side This means that high energy levels have a higher carrier occupation per state on the hotter side but also the hotter side has a lower occupation per state at lower energy levels As before the high energy carriers diffuse away from the hot end and produce entropy by drifting towards the cold end of the device However there is a competing process at the same time low energy carriers are drawn back towards the hot end of the device Though these processes both generate entropy they work against each other in terms of charge current and so a net current only occurs if one of these drifts is stronger than the other The net current is given by J s S T displaystyle scriptstyle mathbf J sigma S boldsymbol nabla T nbsp where as shown below the thermoelectric coefficient s S displaystyle scriptstyle sigma S nbsp depends literally on how conductive high energy carriers are compared to low energy carriers The distinction may be due to a difference in rate of scattering a difference in speeds a difference in density of states or a combination of these effects Mott formula edit The processes described above apply in materials where each charge carrier sees an essentially static environment so that its motion can be described independently from other carriers and independent of other dynamics such as phonons In particular in electronic materials with weak electron electron interactions weak electron phonon interactions etc it can be shown in general that the linear response conductance is s c E d f E d E d E displaystyle sigma int c E Bigg frac df E dE Bigg dE nbsp and the linear response thermoelectric coefficient is s S k B e E m k B T c E d f E d E d E displaystyle sigma S frac k rm B e int frac E mu k rm B T c E Bigg frac df E dE Bigg dE nbsp where c E displaystyle scriptstyle c E nbsp is the energy dependent conductivity and f E displaystyle scriptstyle f E nbsp is the Fermi Dirac distribution function These equations are known as the Mott relations of Sir Nevill Francis Mott 12 The derivative d f E d E 1 4 k B T sech 2 E m 2 k B T displaystyle frac df E dE frac 1 4k rm B T operatorname sech 2 left frac E mu 2k rm B T right nbsp is a function peaked around the chemical potential Fermi level m displaystyle mu nbsp with a width of approximately 3 5 k B T displaystyle 3 5k rm B T nbsp The energy dependent conductivity a quantity that cannot actually be directly measured one only measures s displaystyle sigma nbsp is calculated as c E e 2 D E n E displaystyle c E e 2 D E nu E nbsp where D E displaystyle D E nbsp is the electron diffusion constant and n E displaystyle nu E nbsp is the electronic density of states in general both are functions of energy In materials with strong interactions none of the above equations can be used since it is not possible to consider each charge carrier as a separate entity The Wiedemann Franz law can also be exactly derived using the non interacting electron picture and so in materials where the Wiedemann Franz law fails such as superconductors the Mott relations also generally tend to fail 13 The formulae above can be simplified in a couple of important limiting cases Mott formula in metals edit In semimetals and metals where transport only occurs near the Fermi level and c E displaystyle scriptstyle c E nbsp changes slowly in the range E m k B T displaystyle E approx mu pm k rm B T nbsp one can perform a Sommerfeld expansion c E c m c m E m O E m 2 displaystyle scriptstyle c E c mu c mu E mu O E mu 2 nbsp which leads to S m e t a l p 2 k B 2 T 3 e c m c m O k B T 3 s m e t a l c m O k B T 2 displaystyle S rm metal frac pi 2 k rm B 2 T 3e frac c mu c mu O k rm B T 3 quad sigma rm metal c mu O k rm B T 2 nbsp This expression is sometimes called the Mott formula however it is much less general than Mott s original formula expressed above In the free electron model with scattering the value of c m c m displaystyle scriptstyle c mu c mu nbsp is of order 1 k B T F displaystyle scriptstyle 1 k rm B T rm F nbsp where T F displaystyle T rm F nbsp is the Fermi temperature and so a typical value of the Seebeck coefficient in the Fermi gas is S F e r m i g a s p 2 k B 3 e T T F displaystyle scriptstyle S rm Fermi gas approx tfrac pi 2 k rm B 3e T T rm F nbsp the prefactor varies somewhat depending on details such as dimensionality and scattering In highly conductive metals the Fermi temperatures are typically around 104 105 K and so it is understandable why their absolute Seebeck coefficients are only of order 1 10 mV K at room temperature Note that whereas the free electron model predicts a negative Seebeck coefficient real metals actually have complicated band structures and may exhibit positive Seebeck coefficients examples Cu Ag Au The fraction c m c m displaystyle scriptstyle c mu c mu nbsp in semimetals is sometimes calculated from the measured derivative of s m e t a l displaystyle scriptstyle sigma rm metal nbsp with respect to some energy shift induced by field effect This is not necessarily correct and the estimate of c m c m displaystyle scriptstyle c mu c mu nbsp can be incorrect by a factor of two or more since the disorder potential depends on screening which also changes with field effect 14 Mott formula in semiconductors edit In semiconductors at low levels of doping transport only occurs far away from the Fermi level At low doping in the conduction band where E C m k B T displaystyle scriptstyle E rm C mu gg k rm B T nbsp where E C displaystyle scriptstyle E rm C nbsp is the minimum energy of the conduction band edge one has d f E d E 1 k B T e E m k B T displaystyle scriptstyle frac df E dE approx tfrac 1 k rm B T e E mu k rm B T nbsp Approximating the conduction band levels conductivity function as c E A C E E C a C displaystyle scriptstyle c E A rm C E E rm C a rm C nbsp for some constants A C displaystyle scriptstyle A rm C nbsp and a C displaystyle scriptstyle a rm C nbsp S C k B e E C m k B T a C 1 s C A C k B T a C e E C m k B T G a C 1 displaystyle S rm C frac k rm B e Big frac E rm C mu k rm B T a rm C 1 Big quad sigma rm C A rm C k rm B T a rm C e frac E rm C mu k rm B T Gamma a rm C 1 nbsp whereas in the valence band when m E V k T displaystyle scriptstyle mu E rm V gg kT nbsp and c E A V E V E a V displaystyle scriptstyle c E A rm V E rm V E a rm V nbsp S V k e m E V k B T a V 1 s V A V k B T a V e m E V k B T G a V 1 displaystyle S rm V frac k e Big frac mu E rm V k rm B T a rm V 1 Big quad sigma rm V A rm V k rm B T a rm V e frac mu E rm V k rm B T Gamma a rm V 1 nbsp The values of a C displaystyle scriptstyle a rm C nbsp and a V displaystyle scriptstyle a rm V nbsp depend on material details in bulk semiconductor these constants range between 1 and 3 the extremes corresponding to acoustic mode lattice scattering and ionized impurity scattering 15 nbsp Seebeck coefficient of silicon at 300 K calculated from Mott model The crossover from hole dominated conduction positive S S V displaystyle scriptstyle S approx S rm V nbsp to electron dominated conduction negative S S C displaystyle scriptstyle S approx S rm C nbsp happens for Fermi levels in the middle of the 1 1 eV wide gap In extrinsic doped semiconductors either the conduction or valence band will dominate transport and so one of the numbers above will give the measured values In general however the semiconductor may also be intrinsic in which case the bands conduct in parallel and so the measured values will be S s e m i s C S C s V S V s C s V s s e m i s C s V displaystyle S rm semi frac sigma rm C S rm C sigma rm V S rm V sigma rm C sigma rm V quad sigma rm semi sigma rm C sigma rm V nbsp This results in a crossover behaviour as shown in the figure The highest Seebeck coefficient is obtained when the semiconductor is lightly doped however a high Seebeck coefficient is not necessarily useful on its own For thermoelectric power devices coolers generators it is more important to maximize the thermoelectric power factor s S 2 displaystyle scriptstyle sigma S 2 nbsp 16 or the thermoelectric figure of merit and the optimum generally occurs at high doping levels 17 Phonon drag edit Main article Phonon drag Phonons are not always in local thermal equilibrium they move against the thermal gradient They lose momentum by interacting with electrons or other carriers and imperfections in the crystal If the phonon electron interaction is predominant the phonons will tend to push the electrons to one end of the material hence losing momentum and contributing to the thermoelectric field This contribution is most important in the temperature region where phonon electron scattering is predominant This happens for T 1 5 8 D displaystyle T approx 1 over 5 theta mathrm D nbsp where 8 D displaystyle scriptstyle theta rm D nbsp is the Debye temperature At lower temperatures there are fewer phonons available for drag and at higher temperatures they tend to lose momentum in phonon phonon scattering instead of phonon electron scattering At lower temperatures material boundaries also play an increasing role as the phonons can travel significant distances 18 Practically speaking phonon drag is an important effect in semiconductors near room temperature even though well above 8 D 5 displaystyle scriptstyle theta rm D 5 nbsp that is comparable in magnitude to the carrier diffusion effect described in the previous section 18 This region of the thermopower versus temperature function is highly variable under a magnetic field citation needed Relationship with entropy edit The Seebeck coefficient of a material corresponds thermodynamically to the amount of entropy dragged along by the flow of charge inside a material it is in some sense the entropy per unit charge in the material 19 References edit Thermopower is a misnomer as this quantity does not actually express a power quantity Note that the unit of thermopower V K is different from the unit of power watts a b Blundell Stephen Blundell Stephen J Blundell Katherine M 2010 Concepts in Thermal Physics Oxford University Press p 415 ISBN 978 0 19 956210 7 Joseph R Sootsman Duck Young Chung Mercouri G Kanatzidis 2009 New and Old Concepts in Thermoelectric Materials Angewandte Chemie 48 46 8616 8639 doi 10 1002 anie 200900598 PMID 19866458 Borelius G Keesom W H Johannson C H Linde J O 1932 Establishment of an Absolute Scale for the Thermo electric Force Proceedings of the Royal Academy of Sciences at Amsterdam 35 1 10 Christian J W Jan J P Pearson W B Templeton I M 1958 Thermo Electricity at Low Temperatures VI A Redetermination of the Absolute Scale of Thermo Electric Power of Lead Proceedings of the Royal Society A Mathematical Physical and Engineering Sciences 245 1241 213 Bibcode 1958RSPSA 245 213C doi 10 1098 rspa 1958 0078 S2CID 96708128 Cusack N Kendall P 1958 The Absolute Scale of Thermoelectric Power at High Temperature Proceedings of the Physical Society 72 5 898 Bibcode 1958PPS 72 898C doi 10 1088 0370 1328 72 5 429 Roberts R B 1986 Absolute scales for thermoelectricity Measurement 4 3 101 103 Bibcode 1986Meas 4 101R doi 10 1016 0263 2241 86 90016 3 The Seebeck Coefficient Electronics Cooling com accessed 2013 Feb 01 Moore J P 1973 Absolute Seebeck coefficient of platinum from 80 to 340 K and the thermal and electrical conductivities of lead from 80 to 400 K Journal of Applied Physics 44 3 1174 1178 Bibcode 1973JAP 44 1174M doi 10 1063 1 1662324 Kong Ling Bing 2014 Waste Energy Harvesting Lecture Notes in Energy Vol 24 Springer pp 263 403 doi 10 1007 978 3 642 54634 1 ISBN 978 3 642 54634 1 Datta Supriyo 2005 Quantum Transport Atom to Transistor Cambridge University Press ISBN 978 0 521 63145 7 Cutler M Mott N 1969 Observation of Anderson Localization in an Electron Gas Physical Review 181 3 1336 Bibcode 1969PhRv 181 1336C doi 10 1103 PhysRev 181 1336 Jonson M Mahan G 1980 Mott s formula for the thermopower and the Wiedemann Franz law Physical Review B 21 10 4223 Bibcode 1980PhRvB 21 4223J doi 10 1103 PhysRevB 21 4223 Hwang E H Rossi E Das Sarma S 2009 Theory of thermopower in two dimensional graphene Physical Review B 80 23 235415 arXiv 0902 1749 Bibcode 2009PhRvB 80w5415H doi 10 1103 PhysRevB 80 235415 S2CID 8125966 Semiconductor Physics An Introduction Karlheinz Seeger Imai H Shimakawa Y Kubo Y 10 December 2001 Large thermoelectric power factor in TiS2 crystal with nearly stoichiometric composition Physical Review B 64 24 241104 arXiv cond mat 0111063 Bibcode 2001PhRvB 64x1104I doi 10 1103 PhysRevB 64 241104 S2CID 119389373 G Jeffrey Snyder Thermoelectrics http www its caltech edu jsnyder thermoelectrics a b Mahan G D Lindsay L Broido D A 28 December 2014 The Seebeck coefficient and phonon drag in silicon Journal of Applied Physics 116 24 245102 Bibcode 2014JAP 116x5102M doi 10 1063 1 4904925 OSTI 1185754 Bulusu A Walker D G 2008 Review of electronic transport models for thermoelectric materials Superlattices and Microstructures 44 1 1 Bibcode 2008SuMi 44 1B doi 10 1016 j spmi 2008 02 008 Retrieved from https en wikipedia org w index php title Seebeck coefficient amp oldid 1219812764, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.