Open access peer-reviewed chapter

Quantum Theory of the Seebeck Coefficient in YBCO

Written By

Shigeji Fujita and Akira Suzuki

Submitted: 05 November 2018 Reviewed: 16 April 2019 Published: 27 May 2019

DOI: 10.5772/intechopen.86378

From the Edited Volume

Advanced Thermoelectric Materials for Energy Harvesting Applications

Edited by Saim Memon

Chapter metrics overview

896 Chapter Downloads

View Full Metrics

Abstract

The measured in-plane thermoelectric power (Seebeck coefficient) S ab in YBCO below the superconducting temperature T c ( ∼ 94 K) S ab is negative and T -independent. This is shown to arise from the fact that the “electrons” (minority carriers) having heavier mass contribute more to the thermoelectric power. The measured out-of-plane thermoelectric power S c rises linearly with the temperature T . This arises from moving bosonic pairons (Cooper pairs), the Bose-Einstein condensation (BEC) of which generates a supercurrent below T c . The center of mass of pairons moves as bosons. The resistivity ρ ab above T c has T -linear and T -quadratic components, the latter arising from the Cooper pairs being scattered by phonons.

Keywords

  • Seebeck coefficient
  • in-plane thermoelectric power
  • out-of-plane thermoelectric power
  • moving bosonic pairons (Cooper pairs)
  • Bose-Einstein condensation
  • supercurrent
  • YBCO

1. Introduction

In 1986, Bednorz and Müller [1] reported their discovery of the first of the high- T c cuprate superconductors (La-Ba-Cu-O, T c > 30 K). Since then many investigations [2, 3] have been carried out on high- T c superconductors (HTSC) including Y-Ba-Cu-O (YBCO) with T c 94 K [4]. These compounds possess all of the main superconducting properties, including zero resistance, Meissner effect, flux quantization, Josephson effect, gaps in the excitation energy spectra, and sharp phase transition. In addition these HTSC are characterized by (i) two-dimensional (2D) conduction, (ii) short zero-temperature coherence length ξ 0 ( 10 Å), (iii) high critical temperature T c ( 100 K), and (iv) two energy gaps. The transport behaviors above T c are significantly different from those of a normal metal.

YBCO has a critical (superconducting) temperature T c 94 K, which is higher than the liquid nitrogen temperature (77 K). This makes it a very useful superconductor. Terasaki et al. [5, 6] measured the resistivity ρ , the Hall coefficient R H , and the Seebeck coefficient (thermoelectric power) S in YBCO above the critical temperature T c . A summary of the data is shown in Figure 1. In-plane Hall coefficient R ab H is positive and temperature T -independent, while in-plane Seebeck coefficient S ab is negative and T -independent (anomaly). Thus, there are different charge carriers for the Ohmic conduction and the thermal diffusion. We know that the carrier’s mass is important in the Ohmic currents. Lighter mass particles contribute more to the conductivity. The T independence of R ab H and S ab suggests that “electrons” and “holes” are responsible for the behaviors. We shall explain this behavior, by assuming “electrons” and “holes” as carriers and using statistical mechanical calculations. Out-of-plane Hall coefficient R c H is negative and temperature-independent, while out-of-plane Seebeck coefficient S c is roughly temperature T -linear. We shall show that the pairons, whose Bose condensation generates the supercurrents below T c , are responsible for this strange T -linear behavior. The in-plane resistivity appears to have T -linear and T -quadratic components. We discuss the resistivity ρ above the critical temperature T c in Section 6.

Figure 1.

Normal-state transport of highly oxygenated YBa2Cu3 O 7 δ after Terasaki et al.’s [5, 6]. Resistivities (top panel); Hall coefficients (middle panel); Seebeck coefficient (bottom panel). The subscripts ab and c denote in-copper plane and out-of-plane directions, respectively.

In this paper we are mainly interested in the sign and the temperature behavior of the Seebeck coefficient in YBCO. But we discuss the related matter for completeness. There are no Seebeck currents in the superconducting state below the critical temperature ( S = 0 ).

Advertisement

2. The crystal structure of YBCO: two-dimensional conduction

HTSC have layered structures such that the copper planes comprising Cu and O are periodically separated by a great distance (e.g., a = 3.88 Å, b = 3.82 Å, c = 11.68 Å for YBCO). The lattice structure of YBCO is shown in Figure 2. The succession of layers along the c -axis can be represented by CuO–BaO–CuO2–Y-CuO2–BaO-CuO–[CuO–BaO–…]. The buckled CuO2 plane where Cu-plane and O-plane are separated by a short distance as shown is called the copper planes. The two copper planes separated by yttrium (Y) are about 3 Å apart, and they are believed to be responsible for superconductivity.

Figure 2.

Arrangement of atoms in a crystal of YBa2Cu3O7.

The conductivity measured is a few orders of magnitude smaller along the c -axis than perpendicular to it [7]. This appears to contradict the prediction based on the naive application of the Bloch theorem. This puzzle may be solved as follows [8]. Suppose an electron jumps from one conducting layer to its neighbor. This generates a change in the charge states of the layers involved. If each layer is macroscopic in dimension, we must assume that the charge state Q n of the n th layer can change without limits: Q n = , 2 , 1, 0, 1, 2 , in units of the electron charge (magnitude) e . Because of unavoidable short circuits between layers due to lattice imperfections, these Q n may not be large. At any rate if Q n are distributed at random over all layers, then the periodicity of the potential for electron along the c -axis is destroyed. The Bloch theorem based on the electron potential periodicity does not apply even though the lattice is periodic along the c -axis. As a result there are no k -vectors along the c -axis. This means that the effective mass in the c -axis direction is infinity, so that the Fermi surface for a layered conductor is a right cylinder with its axis along the c -axis. Hence a 2D conduction is established.

Since electric currents flow in the copper planes, there are continuous k -vectors and Fermi energy ε F . Many experiments [1, 2, 3, 9] indicate that a singlet pairs with antiparallel spins called Cooper pairs (pairons) form a supercondensate below T c .

Let us first examine the cause of electron pairing. We first consider attraction via the longitudinal acoustic phonon exchange. Acoustic phonons of lowest energies have long wavelengths λ and a linear energy-momentum ( ε k ) relation:

ε = c s k , E1

may be assumed, where c s is the sound speed. The attraction generated by the exchange of longitudinal acoustic phonons is long-ranged. This mechanism is good for a type I superconductor whose pairon size is of the order of 10 4 Å. This attraction is in action also for a HTSC, but it alone is unlikely to account for the much smaller pairon size.

Second we consider the optical phonon exchange. Roughly speaking each copper plane has Cu and O, and 2D lattice vibrations of optical modes are expected to be important. Optical phonons of lowest energies have short wavelengths of the order of the lattice constants, and they have a quadratic dispersion relation:

ε = ε 0 + A 1 k 1 π a 1 2 + A 2 k 2 π a 2 2 , E2

where ε 0 , A 1 , and A 2 are constants. The attraction generated by the exchange of a massive boson is short-ranged just as the short-ranged nuclear force between two nucleons generated by the exchange of massive pions, first shown by Yukawa [10]. Lattice constants for YBCO are given by a 1 a 2 = 3.88,3.82 Å, and the limit wavelengths λ min at the Brillouin boundary are twice these values. The observed coherence length ξ 0 is of the same order as λ min :

ξ 0 λ min 8 Å . E3

Thus an electron-optical phonon interaction is a viable candidate for the cause of the electron pairing. To see this in more detail, let us consider the copper plane. With the neglect of a small difference in lattice constants along the a - and b -axes, Cu atoms form a square lattice of a lattice constant a 0 = 3.85 Å, as shown in Figure 3. Twice as many oxygen (O) atoms as copper (Cu) atoms occupy midpoints of the nearest neighbors (Cu, Cu) in the x 1 x 2 plane.

Figure 3.

The idealized copper plane contains twice as many O’s as Cu’s.

First, let us look at the motion of an electron wave packet that extends over more than one Cu-site. This wave packet may move easily in 110 rather than the first neighbor directions 100 and 010 . The Bloch wave packets are superposable; therefore, the electron can move in any direction characterized by the two-dimensional k -vectors with bases taken along 110 and 1 1 ¯ 0 . If the number density of electrons is small, the Fermi surfaces should then be a small circle as shown in the central part in Figure 4.

Figure 4.

The two-dimensional Fermi surface of a cuprate model has a small circle (electrons) at the center and a set of four small pockets (holes) at the Brillouin boundary. Exchange of a phonon can create the electron pairon at B B and the hole pairon at A A . The phonon must have a momentum p k , with k being greater than the distance between the electron circle and the hole pockets.

Second, we consider a hole wave packet that extends over more than one O-site. It may move easily in 100 because the Cu-sublattice of a uniform charge distribution favors such a motion. If the number of holes is small, the Fermi surface should consist of the four small pockets shown in Figure 4. Under the assumption of such a Fermi surface, pair creation of ± pairons via an optical phonon may occur as shown in the figure. Here a single-phonon exchange generates an electron transition from A in the O-Fermi sheet to B in the Cu-Fermi sheet and another electron transition from A to B , creating the pairon at B B and the +pairon at A A . From momentum conservation the momentum (magnitude) of a phonon must be equal to times the k -distance AB , which is approximately equal to the momentum of an optical phonon of the smallest energy. Thus an almost insulator-like layered conductor should have a Fermi surface comprising a small electron circle and small hole pockets, which are quite favorable for forming a supercondensate by exchanging an optical phonon.

Advertisement

3. Quantum statistical theory of superconductivity

Following the Bardeen, Cooper, and Schrieffer (BCS) theory [11], we regard the phonon-exchange attraction as the cause of superconductivity. Cooper [12] solved Cooper’s equation and obtained a linear dispersion relation for a moving pairon:

ε = w 0 + 1 2 v F p , E4

where w 0 is the ground-state energy of the Cooper pair (pairon) and v F is the Fermi speed. This relation was obtained for a three-dimensional (3D) system. For a 2D system, we obtain

ε = w 0 + 2 π v F p . E5

The center of mass (CM) motion of a composite is bosonic (fermionic) according to whether the composite contains an even (odd) number of elementary fermions. The Cooper pairs, each having two electrons, move as bosons. In our quantum statistical theory of superconductivity [13], the superconducting temperature T c is regarded as the Bose-Einstein condensation (BEC) point of pairons. The center of mass of a pairon moves as a boson [13]. Its proof is given in Appendix for completeness. The critical temperature T c in 2D is given by

k B T c = 1.24 v F n 1 / 2 , E6

where n is the pairon density. The inter-pairon distance

r 0 n 1 / 2 = 1.24 v F k B T c 1 E7

is several times greater than the BCS pairon size represented by the BCS coherence length:

ξ 0 0.181 v F k B T c 1 . E8

Hence the BEC occurs without the pairon overlap. Phonon exchange can be repeated and can generate a pairon-binding energy ε b of the order of k B T b :

ε b k B T b , T b 1000 K . E9

Thus, the pairons are there above the superconducting temperature T c . The angle-resolved photoemission spectroscopy (ARPES) [14] confirms this picture.

In the quantum statistical theory of superconductivity, we start with the crystal lattice, the Fermi surface and the Hamiltonian and calculate everything, using statistical mechanical methods. The details are given in Ref. [15].

Loram et al. [15] extensively studied the electronic heat capacity of YBa2Cu O 6 + δ with varying oxygen concentrations 6 + δ . A summary of their data is shown in Figure 5. The data are in agreement with what is expected of a Bose-Einstein (B-E) condensation of free massless bosons in 2D, a peak with no jump at T c with the T 2 -law decline on the low-temperature side. The maximum heat capacity at T c with a shoulder on the high-temperature side can only be explained naturally from the view that the superconducting transition is a macroscopic change of state generated by the participation of a great number of pairons with no dissociation. The standard BCS model regards their T c as the pair dissociation point and predicts no features above T c .

Figure 5.

Electronic heat capacity C el plotted as C el / T vs. temperature T after Loram et al. [15] for YBa2Cu3 O 6 + δ with the δ values shown.

The molar heat capacity C for a 2D massless bosons rises like T 2 in the condensed region and reaches 4.38 R at T = T c ; its temperature derivative C T n / T jumps at this point. The order of phase transition is defined to be that order of the derivative of the free energy F whose discontinuity appears for the first time. Since C V = T S / T V = T 2 F / T 2 , C V / T = T 3 F / T 3 2 F / T 2 , the B-E condensation is a third-order phase transition. The temperature behavior of the heat capacity C in Figure 6 is remarkably similar to that of YBa2Cu3 O 6.92 (optimal sample) in Figure 5. This is an important support for the quantum statistical theory. Other support is discussed in Sections 5 and 6.

Figure 6.

The molar heat capacity C for 2D massless bosons rise like T 2 , reaches 4.38 R at the critical temperature T c , and then decreases to 2 R in the high-temperature limit.

Our quantum statistical theory can be applied to 3D superconductors as well. The linear dispersion relation (4) holds. The superconducting temperature T c in 3D is given by

k B T c = 1.01 v F n 1 3 , E10

which is identified as the BEC point. The molar heat capacity C for 3D bosons with the linear dispersion relation ε = cp rises like T 3 and reaches 10.8 R , R = gas constant, at T c = 2.02 cn 0 1 / 3 . It then drops abruptly by 6.57 R and approaches 3 R in the high-temperature limit. This temperature behavior of C is shown in Figure 7. The phase transition is of second order. This behavior is good agreement with experiments, which supports the BEC picture of superconductivity.

Figure 7.

The molar heat capacity C for 3D massless bosons rises like T 3 and reaches 10.8 R at the critical temperature T c = 2.02 cn 0 1 / 3 . It then drops abruptly by 6.57 R and approaches the high-temperature limit 3 R .

Advertisement

4. In-plane Seebeck coefficient above the critical temperature

4.1 Seebeck coefficient for conduction electrons

When a temperature difference is generated and/or an electric field E is applied across a conductor, an electromotive force (emf) is generated. For small potential and temperature gradients, the linear relation between the electric current density j and the gradients

j = σ V + A T = σ E A T E11

holds, where E = V is the electric field and σ is the conductivity. If the ends of the conducting bar are maintained at different temperatures, no electric current flows. Thus from Eq. (11), we obtain

σ E S A T = 0 , E12

where E S is the field generated by the thermal emf. The Seebeck coefficient S , also called the thermoelectric power or the thermopower, is defined through

E S = S T , S A / σ . E13

The conductivity σ is always positive, but the Seebeck coefficient S can be positive or negative depending on the materials. We present a kinetic theory to explain Terasaki et al.’s experimental results [5, 6] for the Seebeck coefficient in YBa2Cu3 O 7 δ , reproduced in Figure 1.

We assume that the carriers are conduction electrons (“electron,” “hole”) with charge q ( e for “electron,” + e for “hole”) and effective mass m . At a finite temperature T > 0 , “electrons” (“holes”) are excited near the Fermi surface if the surface curvature is negative (positive) [16]. The “electron” (“hole”) is a quasi-electron which has an energy higher lower than the Fermi energy ε F and which circulates clockwise (counterclockwise) viewed from the tip of the applied magnetic field vector. “Electrons” (“holes”) are excited on the positive (negative) side of the Fermi surface with the convention that the positive normal vector at the surface points in the energy-increasing direction. The number of thermally excited “electrons” N ex , having energies greater than the Fermi energy ε F , is defined and calculated as

N ex ε F d ε D ε f ε T μ D ε F ε F d ε 1 e β ε μ + 1 ln 2 k B T D ε F , E14

where D ε is the density of states. This formula holds for 2D and 3D in high degeneracy. The density of thermally excited “electrons,”

n ex = N ex / A , A = planer area , E15

is higher at the high-temperature end, and the particle current runs from the high- to the low-temperature end. This means that the electric current runs toward (away from) the high-temperature end in an “electron” (“hole”)-rich material. After using Eqs. (13) and (14), we obtain

S = < 0 for electrons > 0 for holes E16

The Seebeck current arises from the thermal diffusion. We assume Fick’s law:

j = q j particle = qD n ex , E17

where D is the diffusion constant, which is computed from the standard formula:

D = 1 d v = 1 d v F 2 τ , v = v F , = , E18

where v F is the Fermi velocity and τ the relaxation time of the charged particles. The symbol d denotes the dimension. The density gradient n ex is generated by the temperature gradient T and is given by

n ex = ln 2 A d k B D ε F T , E19

where Eq. (14) is used. Using Eqs. (17)(19) and (11), we obtain the thermal diffusion coefficient A as

A = ln 2 2 A qv F 2 k B D ε F τ . E20

We divide A by the conductivity

σ = nq 2 τ / m , E21

and obtain the Seebeck coefficient S [see Eq. (13)]:

S A / σ = ln 2 k B ε F nq D ε F A , ε F 1 2 m v F 2 . E22

The relaxation time τ cancels out from numerator and denominator. This result is independent of the temperature T .

4.2 In-plane thermopower for YBCO

We apply our theory to explain the in-plane thermopower data for YBCO. For YBa2Cu3 O 7 δ (composite), there are “electrons” and “holes”. The “holes”, having smaller m and higher v F 2 ε F / m 1 / 2 , dominate in the Ohmic conduction and also in the Hall voltage V H , yielding a positive Hall coefficient R ab H (see Figure 1). But the experiments indicate that the in-plane thermopower S ab is negative. This puzzle may be solved as follows.

We assume an effective mass approximation for the in-plane “electrons”:

ε = p x 2 + P y 2 / 2 m . E23

The 2D density of states including the spin degeneracy is

D = m A / π 2 , E24

which is independent of energy. The “electrons” (minority carriers), having heavier mass m 1 , contribute more to A , and hence the thermopower S ab can be negative as shown below.

When both “electrons” (1) and “holes” (2) exist, their contributions to the thermal diffusion are additive. Using Eqs. (20) and (24), we obtain

A ab = e ln 2 k B 2 π 2 v F 1 2 m 1 τ 1 + e ln 2 k B 2 π 2 v F 2 2 m 2 τ 2 = e ln 2 k B ε F π 2 τ 1 τ 2 . E25

If phonon scattering is assumed, then the scattering rate is given by

Γ τ 1 = n ph v F s , E26

where s is the scattering diameter and n ph denotes the phonon population given by the Planck distribution function:

n ph = exp ε ph / k B T 1 1 , E27

where ε ph is a phonon energy. We then obtain

τ 1 τ 2 = 1 / Γ 1 1 / Γ 2 = n ph v F 1 s 1 n ph v F 2 s 1 = 1 n ph s 1 v F 1 1 v F 2 > 0 , v F 1 < v F 2 . E28

The total conductivity is

σ = σ 1 + σ 2 = e 2 n 1 m 1 τ 1 + e 2 n 2 m 2 τ 2 = e 2 n 1 m 1 v F 1 n ph s + e 2 n 2 m 2 v F 2 n ph s . E29

Using Eqs. (25)(29), we obtain the in-plane thermopower S ab above the critical temperature as

S ab A ab σ ab = ln 2 k B ε F πe 2 1 v F 1 1 v F 2 n 1 m 1 v F 1 + n 2 m 2 v F 2 1 . E30

The factors n ph s drop out from numerator and denominator. The obtained Seebeck coefficient S ab is negative and T -independent, in agreement with experiments in YBa2Cu3 O 7 δ , reproduced in Figure 1.

Advertisement

5. Out-of-plane thermopower

Terasaki et al. [17, 18] and Takenaka et al. [19] measured the out-of-plane resistivity ρ c in YBa2Cu3Ox. In the range 6.6 < x < 6.92 , the data for ρ c can be fitted with

ρ c = C 1 ρ ab + C 2 / T , E31

where C 1 and C 2 are constants and ρ ab is the in-plane resistivity. The first term C 1 ρ ab arises from the in-plane conduction due to the (predominant) “holes” and + pairons. The second term C 2 / T arises from the pairons’ quantum tunneling between the copper planes [20]. Pairons move with a linear dispersion relation [21]:

ε = 2 π v F p cp , p < p 0 w 0 / c 0 , otherwise E32

with w 0 being the binding energy of a pairon. The Hall coefficient R c H (current along the c -axis) is observed to be negative, indicating that the carriers have negative charge (see Figure 1).

The tunneling current is calculated as follows. A pairon arrives at a certain lattice-imperfection (impurity, lattice defect, etc.) and quantum-jumps to a neighboring layer with the jump rate given by the Dirac-Fermi golden rule

w = 2 π p f U p i 2 δ ε f ε i 2 π M 2 δ ε f ε i , E33

where p i p f and ε i ε f are, respectively, the initial (final) momentum and energy and U is the imperfection-perturbation. We assume a constant absolute squared matrix-elements M 2 . The current density j c i along the c -axis due to a group of particles i having charge q i and momentum-energy p ε is calculated from

j c i = j c , H i j c , L i = q i a 0 wn i v H i v L i , E34

where n i is the 2D number density, a 0 the interlayer distance, and j c , H i j c , L i represents the current density from the high (low)-temperature end. Pairons move with the same speed c = 2 / π v F , but the velocity component v x is

v x = ε p x = cp x p = c 2 ε p x . E35

Lower-energy (smaller p ) pairons are more likely to get trapped by the imperfection and going into tunneling. We represent this tendency by K = B / ε , where B is a constant having the dimension of energy/length. Since the thermal average of the v is different, a steady current is generated. The temperature difference Δ T = T H T L causes a change in the B-E distribution F :

F ε e ε μ β + 1 1 , β k B T 1 , E36

where μ is the chemical potential. We compute the current density j c from

j c = 2 e M 2 B 3 c 2 a 0 Δ T k B T 2 0 cp 0 d ε d F ε d β , E37

assuming a small Δ T . Not all pairons reaching an imperfection are triggered into tunneling. The factor B contains this correction.

At the BEC temperature T c , the chemical potential μ vanishes:

μ T c = 0 , E38

and

βμ μ / k B T E39

is negative and small in magnitude for T > T c . For high temperature and low density, the B-E distribution function F can be approximated by the Boltzmann distribution function:

F ε f 0 ε = exp μ ε β , E40

which is normalized such that

1 2 π 2 d 2 p f 0 ε = n 0 pairon density . E41

All integrals in (37) and (41) can be evaluated simply by using 0 d x e x x n = n ! . Hence we obtain

d 2 p f 0 ε = 2 πm 0 d ε e βμ e βε = 2 πm e μβ β 1 . E42

The integral in (37) is then calculated as

0 cp 0 d ε d F ε d β 0 cp 0 d ε d d β f 0 ε = e μβ 0 d ε ε e βε = e βμ β 2 . E43

From Eqs. (11) and (37) along with Eq. (43), we obtain

A c 2 e M 2 Bk B a 0 3 c 2 1 , E44

which is T -independent.

Experiments [5] indicate that the first term C 1 ρ ab in (31) is dominant for x > 6.8 :

ρ c C 1 ρ ab T . E45

Hence at x = 7 , we have an expression for the out-of-plane Seebeck coefficient S c above the critical temperature:

S c A c σ c = A c C 1 ρ ab T > 0 . ρ ab T . E46

The lower the temperature of the initial state, the tunneling occurs more frequently. The particle current runs from the low- to the high-temperature end, the opposite direction to that of the conduction in the ab -plane. Hence S c > 0 , which is in accord with experiments (see Figure 1).

Advertisement

6. Resistivity above the critical temperature

We use simple kinetic theory to describe the transport properties [22]. Kinetic theory was originally developed for a dilute gas. Since a conductor is far from being the gas, we shall discuss the applicability of kinetic theory. The Bloch wave packet in a crystal lattice extends over one unit cell, and the lattice-ion force averaged over a unit cell vanishes. Hence the conduction electron (“electron,” “hole”) runs straight and changes direction if it hits an impurity or phonon (wave packet). The electron–electron collision conserves the net momentum, and hence, its contribution to the conductivity is zero. Upon the application of a magnetic field, the system develops a Hall electric field so as to balance out the Lorentz magnetic force on the average. Thus, the electron still move straight and is scattered by impurities and phonons, which makes the kinetic theory applicable.

YBCO is a “hole”-type HTSC in which “holes” are the majority carriers above T c , while Nd 1.84 Ce 0.16 CuO 4 is an “electron”-type HTSC.

6.1 In-plane resistivity

Consider a system of “holes,” each having effective mass m 2 and charge + e , scattered by phonons. Assume a weak electric field E applied along the x -axis. Newton’s equation of motion for the “hole” with the neglect of the scattering is

m 2 m dv x d t = eE . E47

Solving it for v x and assuming that the acceleration persists in the mean-free time τ 2 , we obtain

v d = eE m 2 τ 2 E48

for the drift velocity v d . The current density ( x -component) j is given by

j = en 2 v d = n 2 e 2 τ 2 m 2 E , E49

where n 2 is the “hole” density. Assuming Ohm’s law

j = σE , E50

we obtain an expression for the electrical conductivity:

σ 2 = n 2 e 2 m 2 1 Γ 2 , E51

where Γ 2 τ 2 1 is the scattering rate. The phonon scattering rate can be computed, using

Γ 2 = n ph v F A 2 , E52

where A 2 is the scattering diameter. If acoustic phonons having average energies

ω q α 0 ω D k B T , α 0 0.20 E53

are assumed, then the phonon number density n ph is given by [23].

n ph = n a exp α 0 ω D / k B T 1 1 n a k B T α 0 ω D , E54

where

n a 2 π 2 d 2 k E55

is the small k -space area where the acoustic phonons are located.

Using Eqs. (51), (52), and (54), we obtain

σ 2 = C 2 n 2 e 2 T , C 2 α 0 ω D n a m 2 k B v F A 2 . E56

Similar calculations apply to “electrons.” We obtain

σ 1 = C 1 n 1 e 2 T , C 1 α 0 ω D n a m 1 k B v F A 2 . E57

The resistivity ρ is the inverse of the conductivity σ . Hence the resistivity for YBCO is proportional to the temperature T :

ρ 1 σ T . E58

Let us now consider a system of + pairons, each having charge + 2 e and moving with the linear dispersion relation:

ε = cp . E59

Since

v x = d ε / d p p / p x = c p x / p , E60

Newton’s equation of motion is

p c d v x d t = ε c 2 d v x d t = 2 eE , E61

yielding v x = 2 e c 2 / ε Et + initial velocity. After averaging over the angles, we obtain

v d 3 = 2 ec 2 τ 3 E ε 1 , E62

where τ 3 is the pairon mean free time and the angular brackets denote a thermal average. Using this and Ohm’s law, we obtain

σ 3 = 2 e 2 c ε 1 n 3 Γ 3 1 , Γ 3 τ 3 1 , E63

where n 3 is the pairon density and Γ 3 is the pairon scattering rate. If we assume a Boltzmann distribution for bosonic pairons above T c , then we obtain

ε 1 2 π 2 π 2 0 d p p 1 ε e βcp 2 π 2 π 2 0 d p p e βcp 1 = k B T 1 for T > T c . E64

The rate Γ 3 is calculated with the assumption of a phonon scattering. We then obtain

σ 3 = 4 n 3 e 2 c 2 k B T Γ 3 = 2 n 2 e 2 C 3 T 2 , C 3 8 π 2 α 0 ω D v F n a k B 2 A 3 . E65

The total conductivity σ for YBCO is σ 2 + σ 3 . Thus taking the inverse of σ , we obtain, by using the results (56) and (65):

ρ ab 1 σ = C 2 n 2 e 2 T + C 3 n 3 2 e 2 T 2 1 = T 2 n 2 e 2 C 2 T + 2 C 3 E66

while the conductivity for Nd 1.84 Ce 0.16 CuO 4 is given by σ 1 + σ 3 , and hence the resistivity is similarly given by

ρ ab = T 2 n 1 e 2 C 1 T + 2 C 3 . E67

In Nd 1.84 Ce 0.16 while in YCuO4 system, “electrons” and pairons play an essential role for the conduction. In YBa2Cu3 O 7 δ the “holes” and + pairons are the major carriers in the in-plane resistivity. The resistivity in the plane ( ρ ab ) vs. temperature ( T ) in various samples at optimum doping after Iye [24] is shown in Figure 8. The overall data are consistent with our formula.

Figure 8.

Resistivity in the ab plane, ρ ab vs. temperature T . Solid lines represent data for HTSC at optimum doping and dashed lines data for highly overdosed samples, after Iye [24].

At higher temperature > 160 K , the resistivity ρ ab is linear (see formula (58)):

ρ ab T , T > 160 K , E68

in agreement with experiments (Figure 8). This part arises mainly from the conduction electrons scattered by phonon. At the low temperatures close to the critical temperature T c , the in-plane resistivity ρ ab shows a T -quadratic behavior [see formula (66)]:

ρ ab T 2 near  and  above T c . E69

This behavior arises mainly from the pairons scattered by phonons. The agreement with the data represents one of the most important experimental supports for the BEC picture of superconductivity.

References

  1. 1. Bednorz JG, Müller KA. Zeitschrift für Physik B Condensed Matter. 1986;64:189
  2. 2. Halley JW, editor. Theory of High Temperature Superconductivity. Redwood City, CA: Addison-Wesley; 1988
  3. 3. Lundquist S, editor. Towards the Theoretical Understanding of High-T Superconductivity. Singapore: World Scientific; 1988
  4. 4. Wu MK, Ashburn JR, Torng CJ, Hor PH, Meng RL, Gao L, et al. Physical Review Letters. 1987;58:908
  5. 5. Terasaki I, Saito Y, Tajima S, Miyamoto S, Tanaka S. Physica C. 1994;235-240:1413
  6. 6. Terasaki I, Sato Y, Tajima S, Miyamoto S, Tanaka S. Physical Review B. 1995;52:16246
  7. 7. Farrell DE, Beck RG, Booth MF, Allen CJ, Bukowski ED, Ginsberg DM. Physical Review B. 1990;42:6758
  8. 8. Godoy S, Fujita S. Journal of Engineering Science. 1991;29:1201
  9. 9. Kitazawa K, Ishiguro T, editors. Advances in Superconductivity. Tokyo: Springer; 1989
  10. 10. Yukawa H. Proceedings of the Physico-Mathematical Society of Japan. 1935;17:48
  11. 11. Bardeen J, Cooper LN, Schrieffer JR. Physics Review. 1957;108:1175
  12. 12. Cooper LN. Physics Review. 1956;104:1189
  13. 13. Fujita S, Ito K, Godoy S. Quantum Theory of Conducting Matter, Superconductivity. New York: Springer; 2009
  14. 14. Lanzara A, Bogdanov PV, Zhou XJ, Kellar SA, Feng DL, Lu ED, et al. Nature. 2001;412:510
  15. 15. Loram JW, Mirza KA, Cooper JR, Liang WY. Journal of Superconductivity. 1994;7:347
  16. 16. Fujita S, Godoy S, Nguyen D. Foundations of Physics. 1995;25:1209
  17. 17. Terasaki I, Sato Y, Terajima S. Physical Review B. 1997;55:15300
  18. 18. Terasak I, Sato Y, Tajima S. Journal of the Korean Physical Society. 1994;31:23
  19. 19. Takenaka K, Mizuhashi K, Takagi H, Uchida S. Physical Review B. 1994;50:6534
  20. 20. Fujita S, Tamura Y, Suzuki A. Modern Physics Letters B. 2000;14:30
  21. 21. Fujita S, Godoy S. Theory of High Temperature Superconductivity. The Netherlands: Kluwer Academic; 2001. pp. 65-69, p. 108
  22. 22. Fujita S, Kim SY-G, Okamura Y. Modern Physics Letters B. 2000;14:495
  23. 23. Chien TR, Wang ZZ, Ong NP. Physical Review Letters. 1991;67:2088
  24. 24. Iye Y. In: Ginzberg DM, editor. Physical Properties of High Temperature Superconductors III. Singapore: World Scientific; 1989

Written By

Shigeji Fujita and Akira Suzuki

Submitted: 05 November 2018 Reviewed: 16 April 2019 Published: 27 May 2019