<<

REACH Program 2015, Princeton University Lee Wing Hang Randall

Joseph Henry Project – Thermoelectric Battery Supervisor: Prof. Micheal Littman Lee Wing Hang

1. Introduction Seebeck discovered that heating junction of loop composed of dissimilar can deflect a compass in the middle in 1821. Although he did not really notice a voltage hence current was actually generated, the phenomenon is named after him – the Seebeck Effect. Joseph Henry was the first one who predicted the production of sparks by thermoelectricity with connection to ribbon coil. In his lecture on Natural Philosophy in College of New Jersey, he demonstrated the production of sparks with the use of a thermoelectric battery. The battery was kept as an artifact left in Physics Department of Princeton University. From the description in student notebook in his course, the thermoelectric battery is heated from above by a hot iron and cooled from below by freezing mixture to produce a spark.

Figure 1 Thermoelectric battery of Joseph Henry

Figure 2 Student Notebook by Daniel Ayres Jr. The thermoelectric battery is made of 25 pairs of and plates connected in a zig- zag manner with paper intervening each pair. REACH Program 2015, Princeton University Lee Wing Hang Randall

A study to understand the situation when Joseph Henry did the experiment is done. For example, we want to know what was the voltage created and what coil is needed for production of a spark. The study began with theoretical deduction, free gas theory is again verified to be bad theory to explain the seebeck effect. The method to arrive at the Mott Expression of thermopower by Boltzmann Transport Equation is reviewed as well. Study with nearly free electron and semiclassical transport theory would be time-consuming. Collection of experimental data is therefore carried out to deduce the seebeck voltage produced at that time.

2. Experimental Determination The artifact should not be harmed with extreme temperature for historical purposes. Therefore, relevant data regarding the of study on seebeck coefficients of bismuth and antimony are collected from reliable sources. A uniform temperature gradient is assumed so that the seebeck coefficient at certain temperature difference is determined by an average (integration of seebeck coefficient over temperature). Data are collected from papers cautiously and a fitting equation was created for each data set. The data from G.A. Saunders and O. Oktu (1967) showed clear experimental measurement regarding thermopower of antimony. The data are reproduced for analysis as below.

Figure 3 Curve Fitting of Antimony Data (y: Thermopower uV/K, x: Temperature K) REACH Program 2015, Princeton University Lee Wing Hang Randall

Figure 4 Curve Fitting of Antimony with labels For Bismuth, since the absolute thermopower (reference to ) is not available. However, The data for bismuth thermopower relative to and absolute thermopower of copper were found. So The difference between the two was used to compute the absolute thermopower of bismuth.

Figure 5 Curve Fitting of Bismuth Data (copper as reference, y: Thermopower uV/K, x: Temperature K)

REACH Program 2015, Princeton University Lee Wing Hang Randall

Figure 6 Fitting of Copper Data (, y: Thermopower uV/K, x: Temperature K) Fitting equations were created for each curve. The difference between the fitting equations on figure 5 and figure 6 was used to model the absolute thermopower. The computation made on matlab explains it.

Figure 7 Computation for Seebeck Voltage created for connecting 25 pairs of bismuth and antimony

Reference for this section: The Seebeck Coefficient of Bismuth Single Crystals, B.S. Chandrasekhar (1959) Seebeck effect in heavy rare earth single crystals, P40, Larry Robert Sill (1964) The Seebeck Coefficient and the Fermi Surface of Antimony Single Crystals, G.A. Saunders and O. Oktu (1967) REACH Program 2015, Princeton University Lee Wing Hang Randall

3. Free Electron Gas Model This model includes both free electron assumption and independent electron assumption. That means we assume no electron-ion and electron-electron interactions. With this simple model, the relationship between seebeck effect and temperature is sought. 3.1 Drude Theory In 1900, Paul Drude applied the Boltzmann’s kinetic theory of free gases to study the motion within a . There are three main assumptions regarding the Drude Theory. 1. have a scattering time τ. The probability of scattering within a time interval dt is dt/τ. 2. Electron returns to momentum p = 0 after scattering (p as a vector so on average, it vanishes) 3. In between scattering events, the electrons are subjected to Lorentz force. So the expectation of p after dt can be written as: 푑푡 < 풑(푡 + 푑푡) > = (1 − ) (풑(푡) + 푭푑푡) + ퟎ푑푡 휏 F is the Lorentz force which is –e(E + v x B). (1-dt/τ) is the probability that the electron is not scattered so that the resultant momentum changed by the Lorentz Force. If the electron is scattered (with probability dt/τ), the resultant momentum is 0 (with assumption 2). The term with dt2 is dropped and therefore we have, 푑풑 풑 = 푭 − 푑푡 휏 which is a first order differential equation with respect to p, so we expect

풑 = 푨푒−푡/휏 where A is some initial value for momentum. Under only the influence of electric field, can be expressed as 풋 = −푒푛풗 with e as charge of carrier, n as the charge density per unit volume of solid and v as the drift velocity of electron. At steady state, dp/dt is zero, so we have 풑 푚풗 푭 = −푒푬 = = 휏 휏 푒푬휏 풗 = − 푚 푒2푛휏 풋 = − 푬 푚 with conductivity ϭ = -e2nτ/m. REACH Program 2015, Princeton University Lee Wing Hang Randall

Because electrons move only due to electric field according to Drude Model (neglected , charge concentration), voltage difference along the metal is the electrical potential. Seebeck Coefficient is defined as the electric field created caused by temperature gradient. 푬 = 푄훻푻 where Q is the thermopower / Seebeck Coefficient. Consider a one-dimensional model of metal bar. The mean electronic velocity at a point x due to the temperature gradient is 푣2 1 푑푣 푑( ) 푣푞 = [푣(푥 − 푣휏) − 푣(푥 + 푣휏)] = −휏푣 = −휏 2 2 푑푥 푑푥 Generalizing to 3D we have: 휏 푑(푣2) 휏 푑(푣2) 푣푒 = − = − 훻푻 6 푑푥 6 푑푇 For equilibrium under no current flowing, vq + ve = 0: 1 푑 푚푣2 푐 푄 = − ( ) ( ) = − 푣 3푒 푑푇 2 3푛푒 where c is the heat capacity for the electron gases. Classically, c is 3nk/2 where k is the Boltzmann Constant. So the value of Q: 푘 푄 = − , 푤ℎ𝑖푐ℎ 𝑖푠 푎 푐표푛푠푡푎푛푡 2푒 The Drude model makes no sense in explaining the seebeck effect as it is saying that the thermopower is a universal constant for any materials.

Reference of this section: Solid States Physics, Ashcroft/Mermin, Chapter 1 The Oxford Solid State Basics, Steven H. Simon, Chapter 3 3.2 Sommerfeld Theory Sommerfeld incorporated the Fermi Statistics into the Drude’s Theory of metal. Since we consider free electron gas, the Hamiltonian only needs not to include the potential part (which is one of the reason the Sommerfeld model does not explain well since electrons in lattice experience periodic potential). The time-independent Schrodinger Equation is: ℎ2 휕2 휕2 휕2 − ( + + ) ψ(풓) = 휀ψ(풓) 2푚 휕푥2 휕푦2 휕푧2 with periodic boundary condition, ψ(x, y, 푧 + 퐿푧) = ψ(x, y, 푧) REACH Program 2015, Princeton University Lee Wing Hang Randall

ψ(x, y + Ly, 푧) = ψ(x, y, 푧) ψ(x + Lx, y, 푧) = ψ(x, y, 푧) where L’s are the dimensions of solid. The periodic boundary condition gives the quantum confinement of the system. That makes k, the wave vector, to be 2πn/L, where n is integer for the 3 dimensions. Volume of one cell in the k- space is 8π3/L. So the k-space density is V/8π3 where V is the volume of solid. The Fermi Function tells us the probability of a state at certain energy being occupied by fermions (electrons). 1 푓(휀) = 푒(휀−휇)/푘푇 + 1 Because of having 2-spin in one state, so the number of electrons is ∞ 푉 푁 = 2 ∫ 3 푓(휀(풌)) 푑풌 0 8휋 So the electron density per unit volume ∞ 푓(휀(풌)) 푛 = ∫ 3 푑풌 0 4휋 And the energy density is ∞ 푓(휀(풌)) 푢 = ∫ 3 휀(풌) 푑풌 0 4휋 For free electrons, the energy is simply related to its momentum by ħ2k2 휀(풌) = 2푚 And for large k, the volume in k-space forms a sphere, so we can rearrange the integrals in this way ∞ 푓(휀(풌)) ∞ 푘2푑푘푓(휀(풌)) ∞ 푛 = ∫ 3 푑풌 = ∫ 2 = ∫ 푑휀푔(휀)푓(휀(풌)) 0 4휋 0 휋 0 ∞ 푢 = ∫ 푑휀푔(휀)휀푓(휀(풌)) 0 where g(ε) is the per unit volume per small interval of energy.

푚 2푚휀 푔(휀) = √ ħ2휋2 ħ2 REACH Program 2015, Princeton University Lee Wing Hang Randall

Evaluating the integral for electron density yields 휋2 (푘 푇)2푔(휀 ) 2 6 퐵 퐹 1 휋푘퐵푇 휇 = 휀퐹 − ′(휀 ) = 휀퐹(1 − ( ) ) 푔 퐹 3 2휀퐹 And by differentiating the integral for energy density we get the specific heat of the electron gas

2 2 휕푢 휋 2 휋 푘퐵푇 푐푣 = ( )푛 = 푘퐵푇푔(휀퐹) = ( )푛푘퐵 휕푇 3 2 휀퐹 Using this equation of specific heat for the thermopower Q considered in the Drude Model, we have

2 휋 푘퐵푇 푐 2 ( 휀 ) 푛푘퐵 휋2 푘 푇 푘 푄 = − 푣 = − 퐹 = − ( 퐵 )( 퐵) 3푛푒 3푛푒 6 휀퐹 푒 Since the thermopower is temperature dependent now, the writer of this report applied the Sommerfeld theory to bismuth and antimony and compared the results with the experimental data in Section 2 of this report. It was found that free electron gas theory works badly for Seebeck Effect prediction.

The fermi energies εF of bismuth and antimony are 9.90 eV and 10.9 eV respectively. So the computation is simple to compute from the equation of Q above

Figure 8 Computation of thermopower of Bismuth and Antimony REACH Program 2015, Princeton University Lee Wing Hang Randall

Figure 9 Sommerfeld Theory prediction on thermopower of bismuth and antimony The Sommerfeld model predicts a linear relationship between thermopower and temperature, which is not what we observed (see figure 4 and 6). Only fermi energies of metals affect the thermopowers. And this theory does not explain positivity of antimony thermopower.

Reference of this section: Solid States Physics, Ashcroft/Mermin, Chapter 2 The Oxford Solid State Basics, Steven H. Simon, Chapter 4

4. Semi-Classical Transport Theory The semi-classical transport theory, which is derived from Boltzmann Equation, is credible for predicting thermoelectric parameters and the theoretical derived result matches the Law of Widedemann and Franz. The derivation for the seebeck coefficient from semi-classical transport can be referenced to a passage named: Charge and Heat Transport Properties of Electrons. Here we extracted the part related to Seebeck Effect. Consider an electron under a small electric field, temperature gradient, and concentration gradient along the conductor. Consider an infinitesimal point at Z. At this point, the distribution function of electrons is f, and the number of electrons with an energy between E and E+dE is fD(E)dE (where D(E) is the density of states per unit of energy, so D(E)dE gives the number of states within the small energy range). Since the electric field, temperature gradient and concentration gradient are small, these electrons will have almost the same probability to move toward any direction. Also because the solid angle of a sphere is 4, the probability for an electron to move in the (, ) REACH Program 2015, Princeton University Lee Wing Hang Randall direction within a solid angle d s i n dd ) will be d/4. A charge q ( = -e for electrons and +e for holes) moving in the (, ) direction within a solid angle d causes a charge flux of qvcos and energy flux Evcosalong the conduction, where is defined as the angle between the velocity vector and along the conduction. Hence, the charge flux and energy flux in the Z direction carried by all electrons moving toward the entire sphere surrounding the point are respectively

d  2 1   J  ( fD(E))(qv cos )dE  d sin  cosd fD(E)qvdE (1a) Z      4 4 E 0  0 4  0 E 0

d  2 1   J  ( fD(E))(Ev cos )dE  d sin  cosd fD(E)EvdE (1b) E Z      4 4 E 0  0 4  0 E 0 With the relaxation-time approximation, the Boltzmann Transport Equation for electrons take the following form: f   f f  f  v  f  qE    0 (2) t p  where q=-e for electrons and +e for holes. For the steady state case with small temperature/concentration gradient and electric field only, the variation of the distribution function f  f in time is much smaller than that in space, or  v  f , so that we can assume ~0. The t t temperature gradient and electric field is small so that the deviation from equilibrium distribution f f f dE  f f is small, i.e. f  f << f , so we have f  f , and   0  0   v 0 . With these 0 0 0 0 p p E dp E assumptions, eqn. 2 becomes   f f  f v [f  qE 0 ]  0 (3) 0 E  The equilibrium distribution of electrons is the Fermi-Dirac distribution  1 1 E   f0(k )    ;   (4) E(k )   exp() 1 k T exp( ) 1 B kBT where  is the chemical potential that depends strongly on carrier concentration and weakly on temperature (See section 4.2 for the Sommerfeld theory) temperature. Both E and are measured from the band edge (e.g. EC, for the bottom of conduction bands). This reference system essentially sets EC = 0 at different locations although the absolute value of EC measured from a global reference varies at different location as shown in Fig. 6.9b in Chen ( =EF in Chen, (Nanoscale Energy Transport and Conversion: A Parallel Treatment of Electrons, Molecules, Phonons, and  Photons). In this reference system corresponding to Fig. 6.9b in Chen, the same quantum state k 2 2 2 2    (kx  k y  kz ) = (kx, ky, kz) has the same energy E(k )  E(k )  E  at different locations. C 2m REACH Program 2015, Princeton University Lee Wing Hang Randall

 Hence this reference system yields the gradient E(k )  0 (the writer of this report interprets it as constant band structure), simplifying the following derivation. If we use a global reference level  as our zero energy reference point as in Fig. 6.9a in Chen, the same quantum state k = (kx, ky, kz) 2 2 2 2   (kx  k y  kz ) has different energy E(k )   EC because EC changes with locations. In this case,  2m E(k )  EC  0 , making the following derivation somewhat inconvenient. However, both reference systems will yield the same result. From eq. 4,

f0 df0  df0 1 df0 f0     kBT (5) E d E d kBT d E From eq. 5, df f f  0   k T 0  (6) 0 d B E

Also because for the reference system that we are using, from (4)

1  E   1 E     (E(k )  )  T     T (7) 2 2 kBT kBT kBT kBT From eqs. 6-7, f E   f   0 (  T ) (8) 0 E T Combine eqs. 3 and 8, we obtain  E    f f  f v [  T  qE] 0  0 (9) T E  Note that  E  e (10) where  is the electrostatic potential (also called electrical potential, which is the potential e  energy per unit of charge associated with a time-invariant electric field E ); From eqns. 9-10, we obtain  E   f f  f v [  T  q ] 0  0 (11) T e E  From eqn. 11, we obtain REACH Program 2015, Princeton University Lee Wing Hang Randall

 E   f f  f v  [  T ] 0 (12) 0 T E where     qe , is the electrochemical potential that combines the chemical potential and electrostatic potential energy (two reasons to flow of carriers, diffusion current is the part that caused by chemical potential). Electrochemical potential is the (total) driving force for current flow, which can be caused by the gradient in either chemical potential (e.g. due to the gradient in carrier concentration) or the gradient in electrostatic potential (i.e. electric field). When you measure voltage V across a solid using a voltmeter, you actually measured the electrochemical potential difference  per unit charge between the two ends of the solid, i.e. V   / q . If there is no temperature gradient or concentration gradient in the solid, the measured voltage equals e .  In the current case all the gradients and E are in the Z direction, so from eq. 12, d E   dT f d E   dT f f  f v cos[   qE ] 0  f v cos[  ] 0 0 dZ T dZ z E 0 dZ T dZ E (13) Combine eqns. 1 and 13, we obtain the charge flux and energy flux respectively

2 1   J  d sin  cosd f D(E)qvdE Z    0  0 4  0 E 0 (14a) 2   1 2 f0 2 d E   dT   d  sin  cos d  D(E)qv  (   qEz )dE  0 4  0 E 0 E dZ T dZ and

2 1   J  d sin  cosd f D(E)EvdE E Z    0  0 4  0 E 0 (14b) 2   1 2 f0 2 d E   dT   d  sin  cos d  D(E)Ev  (   qEz )dE  0 4  0 E 0 E dZ T dZ

Note that the first term in the right hand of eqn. 14 side is zero and the second term yields

1  f d E   dT J  0 D(E)qv2 (   qE )dE (15a) Z  z 3 E 0 E dZ T dZ

1  f d E   dT J  0 D(E)Ev2 (   qE )dE (15b) E Z  z 3 E 0 E dZ T dZ Note that 1 E  mv2 (16) 2 REACH Program 2015, Princeton University Lee Wing Hang Randall

Use eqn. 16 to eliminate v in eq. 15, we obtain

2q  f d E   dT J   0 D(E)E (   qE )dE Z 3m E dZ T dZ z E 0 (17a) 2q  f d E   dT   0 D(E)E (  )dE 3m E 0 E dZ T dZ

2  f d E   dT J  0 D(E)E 2 (   qE )dE (17b) E Z  z 3m E 0 E dZ T dZ The energy flux from Eq. 17b can be broken up into two terms as following 2  f d E   dT J  0 D(E)E 2 (   qE )dE E Z  z 3m E 0 E dZ T dZ  2 f0 d E   dT   D(E)E(E  ) (   qEz )dE 3m E 0 E dZ T dZ  2 f0 d E   dT    D(E)E (   qEz )dE (18) 3m E 0 E dZ T dZ  2 f0 d E   dT J Z   D(E)E(E  ) (   qEz )dE  3m E 0 E dZ T dZ q 2  f d E   dT J   0 D(E)E(E  ) (  )dE  Z 3m E 0 E dZ T dZ q where JZ is the current density or charge flux given by eq. 17a. At temperature T = 0 K, the first term in the right hand side of eq. 18 is zero, so that the energy flux at T = 0 K is

J Z J E (T  0K)  (19) Z q Because electrons don’t carry any thermal energy at T = 0 K, the thermal energy flux or heat flux carried by the electrons at T ≠ 0 is

J q (T )  J E (T )  J E (T  0) Z Z Z 2  f d E   dT (20)   0 D(E)E(E  ) (  )dE 3m E 0 E dZ T dZ Equations 17a and 20 can be rearranged as 1 d dT J  L ( )  L ( ) (21a) Z 11 q dZ 12 dZ 1 d dT J q  L21( )  L22 ( ) (21b) Z q dZ dZ REACH Program 2015, Princeton University Lee Wing Hang Randall where

2  2q f0 L11    D(E)EdE (22a) 3m E 0 E

 2q f0 L12    D(E)E(E  )dE (22b) 3mT E 0 E

 2q f0 L21    D(E)E(E  )dE  TL12 (22c) 3m E 0 E

 2 f0 2 L22    D(E)E(E  ) dE (22d) 3mT E 0 E Writer of this report remarks that this is the Onsager Relation (Coupled Current Equation). Electrical conductivity: dT d In the case of zero temperature gradient and zero chemical potential,  0 and  0 , eqn. dZ dZ 21a becomes 1 d 1 d J  L ( )  L (  E )  L E (23) Z 11 q dZ 11 q dZ z 11 z The electrical conductivity is defined as

J 2  Z 2q f0    L11    D(E)EdE (24) Ez 3m E 0 E Equation 24 will be simplified further in the following section on Wiedemann-Franz law. Seebeck Coefficient: In the case of non-zero temperature gradient along the Z direction (along the conductor), a thermoelectric voltage can be measured between the two ends of the solid with an open loop electrometer, i.e. J Z  0 (Seebeck voltage is measured when there is not electrical flow). Hence from eq. 21a we obtain 1 d dT J  L ( )  L ( ) =0 (25) Z 11 q dZ 12 dZ Therefore

 d    dZ qL     12 (26)  dT  L   11  dZ  REACH Program 2015, Princeton University Lee Wing Hang Randall

As discussed above, the voltage that the electrometer measure between the two ends of the solid is V   / q . Similarly, dV  d / q . The Seebeck coefficient is defined as the ratio between the voltage gradient and the temperature gradient for an open loop configuration with zero net current flow

 dV   d    f        0 D(E)E(E  )dE  dZ 1 dZ L 1 E S          12   E0   dT  q  dT  L qT   f      11   0 D(E)EdE   dZ   dZ  E  E0  (27)   f 2    0 D(E)E dE  1 E     E0  qT   f    0 D(E)EdE   E0 E 

1 d dT Combine eq. 27, 24, and 21 a, we can write J   ( ) S( ) Z q dZ dZ The scattering mean free time depends on the energy, and we can assume

r   0E (28) where 0 is a constant independent of E. When E is measured from the band edge for either electrons or holes, the density of states

(2m)3/ 2 D(E)  E1/ 2 (29) 2 23

Combine eqns. 27 & 29

   f 2   f 2r1/ 2    0 D(E)E dE    0 E dE  1 E 1 E S     E 0      E 0  (30)   qT  f  qT  f 1r1/ 2    0 D(E)EdE    0 E dE   E 0 E   E 0 E  The integrals in Eq. 30 can be simplified using the product rule

   f0 s s  s1 s1  E dE  f0E |0 s  f0E dE  s  f0E dE (31) E 0 E E 0 E 0 Using eq. 31 to reduce eq. 30 to REACH Program 2015, Princeton University Lee Wing Hang Randall

  r3/ 2   r  5/ 2  f E dE  1 0 S     E 0  (32)  qT  r1/ 2   r  3/ 2  f0E dE   E 0 

The two integrals in eq. 32 can be simplified with the reduced energy   E / kBT

  n n1 n n1  f0 (E, )E dE  kBT   f0 ( ,) d  kBT  Fn ();    / kBT (33) E 0 0 where the Fermi-Dirac integral is defined as

 n Fn ()   f0 ( ,) d (34) 0 Use Eq. 33 to reduce eq. 32 to

  5     5    r  F ()   r  F ()  1 2 r3/ 2 k 2 r3/ 2 S     k T      B      (35) qT  B  3   q   3    r  Fr1/ 2 ()   r  Fr1/ 2 ()    2     2   Seebeck coefficient for metals:

For metals with    / kBT  0 , the Fermi-Dirac integral can be expressed in the form of a rapidly converging series

 1  f F ()  f  nd   0 n1d n  0  0 n 1 0 

   m n1 m  1 f n1 d   ( )   0   d n 1     d m m!  0  m1    (36)  2 1 f0  n1 n n1 ( )       (n 1) ( )  (n 1)n  ...d n 1 0   2   n1  2   n n1  ... n 1 6 If we use only the first two terms of eq. 36 to express the two Fermi-Dirac integrals in eq. 35, we obtain the following (q = -e for electrons in metals) REACH Program 2015, Princeton University Lee Wing Hang Randall

  5     3   5    r  F ()  r  F ()  r  F ()  k 2 r3/ 2 k 2 r1/ 2 2 r3/ 2 S   B       B       q   3   e   3    r  Fr1/ 2 ()   r  Fr1/ 2 ()    2     2           r3/ 2 2   r5/ 2 2    3    1    5    3   r    r   r1/ 2   r    r   r1/ 2     2  3  2  6   2  5  2  6    r    r   kB   2   2      (37) e  3  r3/ 2  r      2  3   r    2     2k k T 3   B ( B )(  r) 3e  2

Note that this is the Mott Expression for seebeck coefficient of metals. This value can be either positive or negative depending on r, or how the scattering rate depends on electron energy. We can ignore the weak temperature dependence of  and assume  = EF, the that is the highest energy occupied by electrons at 0 K in a metal. To apply the theory of semiclassical transport, the band structure of bismuth and antimony need to be obtained. Scattering data (form factor) and crystal structure need to be understood in advance to that. The writer of this report will put it in another document for detailed analysis Reference of this section: Charge and Heat Transport Properties of Electrons, Li Shi, University of Texas at Austin