Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Patterson, Bailey - Solid State Physics Introduction to theory

.pdf
Скачиваний:
1134
Добавлен:
08.01.2014
Размер:
7.07 Mб
Скачать

Problems 457

(with A, B, C, and u0 being positive constants). Generalization to higher dimension have been made. The solitary wave owes its stability to the competition of dispersion and nonlinear effects (such as a tendency to steepen waves). The solitary wave propagates with a velocity that depends on amplitude.

Problems

7.1Calculate the demagnetization factor of a sphere.

7.2In the mean-field approximation in dimensionless units for spin 1/2 ferromagnets the magnetization (m) is given by

m = tanh m

t ,

where t = T/Tc and Tc is the Curie temperature. Show that just below the Curie temperature t < 1,

m = 3 1t .

7.3Evaluate the angular momentum L and magnetic moment μ for a sphere of mass M (mass uniformly distributed through the volume) and charge Q (uniformly distributed over the surface), assuming a radius r and an angular velocity ω. Thereby, obtain the ratio of magnetic moment to angular momentum.

7.4Derive Curie’s law directly from a high-temperature expansion of the partition function. For paramagnets, Curie’s law is

χ= CT (The magnetic susceptibility),

where Curie’s constant is

C = μ0 Ng 2μB2 j( j +1) .

3k

N is the number of moments per unit volume, g is Lande’s g factor, μB is the Bohr magneton, and j is the angular momentum quantum number.

7.5Prove (7.155).

7.6Prove (7.156).

458 7 Magnetism, Magnons, and Magnetic Resonance

7.7 In one spatial dimension suppose one assumes the Heisenberg Hamiltonian

H = −

1

R,RJ (R R)SR SR, J( 0 ) = 0,

 

2

 

where R R′ = ±a for nearest neighbor and J1 Ja) > 0, J2 J(±2a) = − J1/2 with the rest of the couplings being zero. Show that the stable ground state is helical and find the turn angle. Assume classical spins. For simplicity, assume the spins are confined to the (x,y)-plane.

7.8Show in an antiferromagnetic spin wave that the neighboring spins precess in the same direction and with the same angular velocity but have different amplitudes and phases. Assume a one-dimensional array of spins with near- est-neighbor antiferromagnetic coupling and treat the spins classically.

7.9Show that (7.163) is a consistent transformation in the sense that it obeys a relation like (7.175), but for Sj.

7.10Show that (7.138) can be written as

H= −J jδ [S jz S j +δ ,z + 12 (S j S +j +δ + S +j S +j +δ )] 2μ0μH j S jz .

7.11Using the definitions (7.179), show that

[bk ,bk] = δkk, [bk ,bk] = 0, [bk,bk] = 0 .

7.12 (a) Apply Hund’s rules to find the ground state of Nd3+ (4f 35s2p6).

(b) Calculate the Lande g-factor for this case.

7.13 By use of Hund’s rules, show that the ground state of Ce3+ is 2F5/2, of Pm3+ is 5I4, and of Eu3+ is 7F0.

7.14Explain what the phrases “3d1 configuration” and “2D term” mean.

7.15Give a rough order of magnitude estimate of the magnetic coupling energy

of two magnetic ions in EuO (Tc 69 K). How large an external magnetic field would have to be applied so that the magnetic coupling energy of a sin-

gle ion to the external field would be comparable to the exchange coupling energy (the effective magnetic moment of the magnetic Eu2+ ions is 7.94 Bohr magnetons)?

7.16Estimate the Curie temperature of EuO if the molecular field were caused by magnetic dipole interactions rather than by exchange interactions.

7.17Prove the Bohr–van Leeuwen theorem that shows the absence of magnetism with purely classical statistics. Hint – look at Chap. 4 of Van Vleck [7.63].

8 Superconductivity

8.1 Introduction and Some Experiments (B)

In 1911 H. Kamerlingh Onnes measured the electrical resistivity of mercury and found that it dropped to zero below 4.15 K. He could do this experiment because he was the first to liquefy helium and thus he could work with the low temperatures required for superconductivity. It took 46 years before Bardeen, Cooper, and Schrieffer (BCS) presented a theory that correctly accounted for a large number of experiments on superconductors. Even today, the theory of superconductivity is rather intricate and so perhaps it is best to start with a qualitative discussion of the experimental properties of superconductors.

Superconductors can be either of type I or type II, whose different properties we will discuss later, but simply put the two types respond differently to external magnetic fields. Type II materials are more resistant, in a sense, to a magnetic field that can cause destruction of the superconducting state. Type II superconductors are more important for applications in permanent magnets. We will introduce the Ginzburg–Landau theory to discuss the differences between type I and type II.

The superconductive state is a macroscopic state. This has led to the development of superconductive quantum interference devices that can be used to measure very weak magnetic fields. We will briefly discuss this after we have laid the foundation by a discussion of tunneling involving superconductors.

We will then discuss the BCS theory and show how the electron–phonon interaction can give rise to an energy gap and a coherent motion of electrons without resistance at sufficiently low temperatures.

Until 1986 the highest temperature that any material stayed superconducting was about 23 K. In 1986, the so-called high-temperature ceramic superconductors were found and by now, materials have been discovered with a transition temperature of about 140 K (and even higher under pressure). Even though these materials are not fully understood, they merit serious discussion. In 2001 MgB2, an intermetallic material was discovered to superconduct at about 40 K and it was found to have several unusual properties. We will also discuss briefly so-called heavyelectron superconductors.

Besides the existence of superconductivity, Onnes further discovered that a superconducting state could be destroyed by placing the superconductor in a large enough magnetic field. He also noted that sending a large enough current through the superconductor would destroy the superconducting state. Silsbee later suggested that these two phenomena were related. The disruption of the superconductive state is caused by the magnetic field produced by the current at the surface of

460 8 Superconductivity

the wire. However, the critical current that destroys superconductivity is very structure sensitive (see below) so that it can be regarded for some purposes as an independent parameter. The critical magnetic field (that destroys superconductivity) and the critical temperature (at which the material becomes superconducting) are related in the sense that the highest transition temperature occurs when there is no external magnetic field with the transition temperature decreasing as the field increases. We will discuss this a little later when we talk about type I and type II superconductors. Figure 8.1 shows at low temperature the difference in behavior of a normal metal versus a superconductor.

Fig. 8.1. Electrical resistivity in normal and superconducting metals (schematic)

Hc(T)

Normal

Superconducting

T

Tc

Fig. 8.2. Schematic of critical field vs. temperature for Type I superconductors

In 1933, Meissner and Ochsenfeld made another fundamental discovery. They found that superconductors expelled magnetic flux when they were cooled below the transition temperature. This established that there was more to the superconducting state than perfect conductivity (which would require E = 0); it is also a state of perfect diamagnetism or B = 0. For a long, thin superconducting specimen, B = H

+4πM (cgs). Inside B = 0, so H + 4πM = 0 and Hin = Ba (the externally applied B field) by the boundary conditions of H along the length being continuous. Thus, Ba

+4πM = 0 or χ = M/Ba = −1/(4π), which is the case for a perfect diamagnet. Exclusion of the flux is due to perfect diamagnetism caused by surface currents, which are always induced so as to shield the interior from external magnetic fields. A simple application of Faraday’s law for a perfect conductor would lead to a constant

8.1 Introduction and Some Experiments (B) 461

flux rather than excluded flux. A plot of critical field versus temperature typically (for type I as we will discuss) looks like Fig. 8.2. The equation describing the critical fields dependence on temperature is often empirically found to obey

 

 

 

 

 

 

 

T

 

2

 

 

 

 

 

 

 

1

 

.

 

H

c

(T ) = H

c

(0)

 

 

 

(8.1)

 

 

 

 

 

 

T

 

 

 

 

 

 

 

 

 

 

 

c

 

 

 

In 1950, H. Fröhlich discussed the electron–phonon interaction and considered the possibility that this interaction might be responsible for the formation of the superconducting state. At about the same time, Maxwell and Reynolds, Serin, Wright, and Nesbitt found that the superconducting transition temperature depended on the isotopic mass of the atoms of the superconductor. They found MαTcconstant. This experimental result gave strong support to the idea that the elec- tron–phonon interaction was involved in the superconducting transition. In the simplest model, α = 1/2.

In 1957, Bardeen, Cooper, and Schrieffer (BCS) finally developed a formalism that contained the correct explanation of the superconducting state in common superconductors. Their ideas had some similarity to Fröhlich’s. A key idea of the BCS theory was developed by Cooper in 1956. Cooper analyzed the electron– phonon interaction in a different way from Fröhlich. Fröhlich had discussed the effect of the lattice vibrations on the self-energy of the electrons. Cooper analyzed the effect of lattice vibrations on the effective interaction between electrons and showed that an attractive interaction between electrons (even a very weak attractive interaction at low enough temperature) would cause pairs of the electrons (the Cooper pairs) to form bound states near the Fermi energy (see Sect. 8.5.3). Later, we will discuss the BCS theory and show the pairing interaction causes a gap in the density of single-electron states.

Type I

 

Type II

 

|4πM|

 

|4πM|

 

 

Super.

Normal

Super.

Vortex

Normal

Hc

Ba

 

Hc1

Hc2 Ba

(a)

 

 

(b)

 

Fig. 8.3. (a) Type I and (b) Type II superconductors

As we have mentioned a distinction is made between type I and type II superconductors. Type I have only one critical field while type II have two critical fields. The idea is shown in Fig. 8.3a and b. 4πM is the magnetic field produced by the surface superconducting currents induced when the external field is applied. Type I superconductors either have flux penetration (normal state) or flux exclusion

462 8 Superconductivity

(superconductivity state). For type II superconductors, there is no flux penetration below Hc1, the lower critical field, and above the upper critical field Hc2 the material is normal. But, between Hc1 and Hc2 the superconductivity regions are threaded by vortex regions of the flux penetration. The idea is shown in Fig. 8.4.

Type I and type II behavior will be discussed in more detail after we discuss the Ginzburg–Landau equations for superconductivity. We now mention some experiments that support the theories of superconductivity.

Fig. 8.4. Schematic of flux penetration for type II superconductors. The gray areas represent flux penetration surrounded by supercurrent (vortex). The net effect is that the superconducting regions in between have no flux penetration

n /α

s α

1.0

 

 

 

 

0.9

[110]

10 Mc

[111]

30.5 Mc

0.8

26 Mc

50 Mc

 

45 Mc

 

 

0.7

0.6

 

 

 

 

 

n

 

 

 

 

 

0.5

 

 

 

 

α/

 

 

 

 

 

 

 

 

 

s

 

 

 

 

 

 

 

 

 

 

α

 

 

 

 

 

0.4

 

 

 

 

 

 

 

 

 

 

0.3

 

 

 

 

 

 

 

 

 

 

0.2

 

 

 

 

 

 

 

 

 

 

0.1

 

 

 

 

 

 

 

 

 

 

0

0.6

0.7

0.8

0.9

1.0

0.6

0.7

0.8

0.9

1.0

0.5

 

 

 

T/Tc

 

0.5

 

 

T/T

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

c

 

 

Fig. 8.5. Absorption coefficients ultrasonic attenuation in Pb (αn refers to the normal state, αs refers to the superconducting state, and Tc is the transition temperature). The dashed curve is derived from BCS theory and it uses an energy gap of 4.2 kTc. [Reprinted with permission from Love RE and Shaw RW, Reviews of Modern Physics 36(1) part 1, 260 (1964). Copyright 1964 by the American Physical Society.]

8.1 Introduction and Some Experiments (B) 463

8.1.1 Ultrasonic Attenuation (B)

The BCS theory of the ratio of the normal to the superconducting absorption coefficients (αn to αs) as a function of temperature variation of the energy gap (discussed in detail later) can be interpreted in such a way as to give information on the temperature variation of the energy gap. Some experimental results on (αn/αs) versus temperature are shown in Fig. 8.5. Note the close agreement of experiment and theory, and that the absorption of superconductors is much lower than for the normal case when well below the transition temperature.

8.1.2 Electron Tunneling (B)

There are at least two types of tunneling experiments of interest. One involves tunneling from a superconductor to a superconductor with a thin insulator separating the two superconductors. Here, as will be discussed later, the Josephson effects are caused by the tunneling of pairs of electrons. The other type of tunneling (Giaever) involves tunneling of single quasielectrons from an ordinary metal to a superconducting metal. As will be discussed later, these measurements provide information on the temperature dependence of the energy gap (which is caused by the formation of Cooper pairs in the superconductor), as well as other features.

8.1.3 Infrared Absorption (B)

The measurement of transmission or reflection of infrared radiation through thin films of a superconductor provides direct results for the magnitude of the energy gap in superconductors. The superconductor absorbs a photon when the photon’s energy is large enough to raise an electron across the gap.

8.1.4 Flux Quantization (B)

We will discuss this phenomenon in a little more detail later. Flux quantization through superconducting rings of current provides evidence for the existence of paired electrons as predicted by Cooper. It is found that flux is quantized in units of h/2e, not h/e.

8.1.5 Nuclear Spin Relaxation (B)

In these experiments, the nuclear spin relaxation time T1 is measured as a function of temperatures. The time T1 depends on the exchange of energy between the nuclear spins and the conduction electrons via the hyperfine interaction. The data for T1 for aluminum looks somewhat as sketched in Fig. 8.6. The rapid change of T1 near T = Tc can be explained, at least quantitatively, by BCS theory.

464 8 Superconductivity

T1

 

1

T/Tc

 

Fig. 8.6. Schematic of nuclear spin relaxation time in a superconductor near Tc

8.1.6 Thermal Conductivity (B)

A sketch of thermal conductivity K versus temperature for a superconductor is shown in Fig. 8.7. Note that if a high enough magnetic field is turned on, the material stays normal—even below Tc. So, a magnetic field can be used to control the thermal conductivity below Tc.

K

Normal (field)

Normal

Super

T

Tc

Fig. 8.7. Effect of magnetic field on thermal conductivity K

All of the above experiments have tended to confirm the BCS ideas of the superconducting state. A central topic that needs further elaboration is the criterion for occurrence of superconductivity in any material. We would like to know if the BCS interaction (electrons interacting by the virtual exchange of phonons) is the only interaction. Could there be, for example, superconductivity due to magnetic interactions? Over a thousand superconducting alloys and metals have been found, so superconductivity is not unusual. It is, perhaps, still an open question as to how common it is.

In the chapter on metals, we have mentioned heavy-fermion materials. Superconductivity in these materials seems to involve a pairing mechanism. However, the most probable cause of the pairing is different from the conventional BCS theory. Apparently, the nature of this “exotic” pairing has not been settled as of this writing, and reference needs to be made to the literature (see Sect. 8.7).

8.2 The London and Ginzburg–Landau Equations (B) 465

For many years, superconducting transition temperatures (well above 20 K) had never been observed. With the discovery of the new classes of high-temperature superconductors, transition temperatures (well above 100 K) have now been observed. We will discuss this later, also. The exact nature of the interaction mechanism is not known for these high-temperature superconductors, either.

8.2 The London and Ginzburg–Landau Equations (B)

We start with a derivation of the Ginzburg–Landau (GL) equations, from which several results will follow, including the London equations. Originally, these equations were proposed on intuitive, phenomenological lines. Later, it was realized they could be derived from the BCS theory. Gor’kov showed the GL theory was a valid description of the BCS theory near Tc. He also showed that the wave function ψ of the GL theory was proportional to the energy gap. Also, the density of superconducting electrons is |ψ|2. Due to spatial inhomogeneities ψ = ψ(r), where ψ(r) is also called the order parameter. This whole theory was developed further by Abrikosov and is often known as the Ginzburg, Landau, Abrikosov, and Gor’kov theory (for further details, see, e.g., Kuper [8.20]

Near the transition temperature, the free energy density in the phenomenological GL theory is assumed to be (gaussian units)

 

2

 

1

 

 

4

 

1

 

 

 

qA

 

2

h2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

FS(r) =αψ

 

+

 

β

ψ

 

+

 

 

 

 

 

 

ψ

 

+ FN +

 

,

(8.2)

 

2

 

2mS

i

c

8π

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where N and S refer to normal and superconducting phases. The coefficients α and β are phenomenological coefficients to be discussed. h2/8π is the magnetic energy density (h = h(r) is local and the magnetic induction B is determined by the spatial average of h(r), so A is the vector potential for h, h = × A). m* = 2m (for pairs of electrons), q = 2e is the charge and is negative for electrons, and ψ is the complex superconductivity wave function. Requiring (in the usual calculus of variations procedure) δFS/δψ* to be zero (δFS/δψ = 0 would yield the complex conjugate of (8.3)), we obtain the first Ginzburg–Landau equation

 

1

 

 

qA

2

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+α + β

ψ

 

(8.3)

 

 

 

 

 

 

ψ = 0 .

 

2m

i

 

c

 

 

 

 

 

 

 

 

S

 

 

 

 

 

 

 

 

 

 

FS can be regarded as a functional of ψ and A, so requiring δFS/δA = 0 we obtain the second GL equation for the current density:

 

c

 

 

 

 

 

 

q

 

 

 

 

 

2

 

 

j =

 

 

×h =

 

 

(ψ ψ ψ ψ )

q

ψ ψ A

 

4π

 

 

 

 

 

 

 

 

 

 

 

 

2m i

 

 

m c

(8.4)

 

q

 

 

 

 

2

 

 

 

q

 

 

 

 

=

 

 

ψ

 

φ

 

 

 

 

 

 

 

 

 

 

 

A

,

 

 

 

m

 

c

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

466 8 Superconductivity

where ψ = |ψ|eiφ. Note (8.3) is similar to the Schrödinger wave equation (except for the term involving β) and (8.4) is like the usual expression for the current density. Writing nS = |ψ|2 and neglecting, as we have, any spatial variation in |ψ|, we find

 

 

q2n

 

 

 

 

 

q

2n

× J = −

 

 

 

× A = −

 

S

B ,

 

m c

 

 

 

 

 

 

 

 

 

m c

so,

 

 

 

 

 

 

 

 

 

 

 

× J = −

c

B ,

(8.5)

2

 

 

 

 

 

 

4πλL

 

 

 

where

 

 

 

 

 

 

 

 

 

 

 

 

λ2L =

 

m c2

 

 

 

 

 

 

,

 

(8.6)

 

 

4πn q2

 

 

 

 

 

 

 

S

 

 

 

where λL is the London penetration depth. Equation (8.5) is London’s equation. Note this is the same for a single electron (where m* = m, q = e, nS = ordinary den-

sity) or a Cooper pair (m* = 2m, q = 2e, nS n/2).

Let us show why λL is called the London penetration depth. At low frequencies, we can neglect the displacement current in Maxwell’s equations and write

× B =

4π

J .

(8.7)

c

 

 

 

Combining with (8.5) that we assume to be approximately true, we have

×( × B) =

4π

 

× J = −

1

B ,

(8.8)

 

 

 

c

 

 

 

λ2L

 

or using B = 0, we have

 

 

 

 

 

 

 

2B =

 

1

B .

 

 

(8.9)

 

2

 

 

 

 

 

λL

 

 

 

For a geometry with a normal material for x < 0 and a superconductor for x > 0, if the magnetic field at x = 0 is B0, the solution of (8.9) is

B(x) = B0 exp(x / λL ) .

(8.10)

Clearly, λL is a penetration depth. Thus, if we have a very thin superconducting film (with thickness << λL), we really do not have a Meissner effect (flux exclusion). Magnetic flux will penetrate the surface of a superconductor over a distance approximately equal to the London penetration depth λL 100 to 1000 Å. Actually, λL is temperature dependent and can be well described by

 

λL

 

2

 

1

 

 

 

 

=

 

,

(8.11)

 

 

(T / T )4

 

λL

0

 

1

 

 

 

 

 

 

 

c

 

 

Соседние файлы в предмете Химия