• No results found

Properties in n-component London Superconductors

N/A
N/A
Protected

Academic year: 2021

Share "Properties in n-component London Superconductors"

Copied!
47
0
0

Loading.... (view fulltext now)

Full text

(1)

Master Thesis

Properties in n-component London Superconductors

Sergio Ampuero Felix

Condensed Matter Physics, Department of Physics, School of Engineering Sciences

Royal Institute of Technology, SE-106 91 Stockholm, Sweden Stockholm, Sweden 2017

(2)

TRITA-FYS 2017:83 ISSN 0280-316X

ISRN KTH/FYS/–17:83—SE

Sergio Ampuero Felix, December 2017c

Printed in Sweden by Universitetsservice US AB, Stockholm December 2017

(3)

Abstract

The purpose of this thesis is to analytically investigate n-component London super- conductor properties that have previously been investigated analytically in cases involving a lower amount of components.

For vortex excitations in n-component London superconductor, we obtain ex- pressions for magnetic flux, magnetic vector potential, microscopic magnetic field and circulation of velocity.

Furthermore, the critical applied field for vortex formation is found.

Finally, the response of a cylindrical specimen to rotation is derived both in terms of a n-component analogue of the London field but also by finding the critical angular frequency for vortex formation.

iii

(4)
(5)

Preface

This thesis assumes that the reader has some familiarity with quantum physics, functional derivation, the concept of minimizing appropriate free energy and the Ginzburg-Landau model for superconductivity.

The result of work done from roughly Mars 2017 to November 2017 is man- ifested through this thesis. I would like to take this opportunity to express my utmost gratitude to my thesis supervisor Egor Bavaev for his patience and excel- lent guidance.

v

(6)
(7)

Contents

Abstract . . . iii

Preface v Contents vii 1 Introduction 1 2 General Considerations 3 2.1 The n-component free energy density Fs . . . 3

2.2 Deriving the Ginzburg-Landau equations . . . 4

2.2.1 Variation with respect to ψl . . . 4

2.2.2 Variation with respect to ~A . . . 5

2.3 London limit . . . 7

2.4 Supercurrents in the London limit . . . 7

3 Vortices 9 3.1 Definitions . . . 9

3.2 Vortex flux . . . 9

3.3 Calculating the magnetic vector potential in the London limit . . . 10

3.4 Calculating the microscopic field in the London limit . . . 12

3.5 Determining the constant C2 . . . 13

3.6 Calculating the circulation of velocity in the London limit . . . 15

3.6.1 2-component example . . . 16

3.6.2 The special case of exponentially decaying supercurrents . . 16

3.7 Free energy of a vortex in the London limit . . . 17

3.7.1 2-component example . . . 20

3.7.2 The special case of exponentially decaying supercurrents . . 21

4 Response to Applied Field ~H 23 4.1 Minimizing Gibb’s free energy . . . 23

4.1.1 2-component example . . . 25 vii

(8)

5 Rotational Response 27

5.1 The London field . . . 27

5.1.1 1-component example . . . 28

5.1.2 2-component example . . . 28

5.2 Vortex formation condition . . . 28

5.2.1 Calculating the z-component of the angular momentum in the London limit . . . 28

5.2.2 Finding the critical angular frequency . . . 31

6 Summary and Conclusions 35

Bibliography 37

(9)

Chapter 1

Introduction

Multicomponent London superconductivity has attracted interest for an extended period of time. See for example [4] and [5] for discussion of flux fractionalization and neutral sector.

The existence of a neutral sector motivated comparison to established results from analysis of neutral one-component superfluids, in particular Onsager’s [12, year 1949] and Feynman’s [14, year 1955] quantization of vortices and the associated circulation quanta m~. Notably, an investigation of the two-component equivalent was published in 2007 [9].

A distinct concern is how superconductors respond to rotation. In 1950, Lon- don provided pioneering analysis of rotating one-component superconductors [13].

Consequently, a dissection of the two-component correspondent was promulgated in 2007 [9]. In this thesis, the cynosure is an analytical examination of how n- component London superconductors respond to application of an external magnetic field and rotation respectively.

Among germane work, one finds an article from 2004 [11] in which expressions for the n-component magnetic flux Φ and current ~J were partial results.

Throughout this thesis, absence of interband couplings is assumed. An article [9] proposes that a physical example of a two-component version of a such system could be liquid metallic deuterium, liquid metallic hydrogen [15] and condensates in neutron stars [16] [23] [24]. Moreover, [11] advocates that higher component cognates should exist in ”metallic phases of light atoms under extreme pressure”.

Due to metallic hydrogen requiring extreme pressures in order to be produced, the task of creating and studying such matter through experiments is a difficult one. This is a subject of broad current research efforts. For example: a prominent article [21] published in the current year, 2017, provides information to ”stimulate theoretical predictions of how to retain metastably hydrogenous materials made at high pressure P on release to ambient”.

1

(10)

A different article [22] aids by ”As such, the present study bridges the important but less well-explored intermediate regime between warm dense fluid and the solid phases at high P-T conditions.”.

A third article [20] analyzes the closely related topic of semimetallic hydrogen.

Most spectacular is an article describing a realized production of solid metallic hydrogen in a laboratory setting [17]. That claim was heavily criticized through arguments indicating that there existed severe experimental flaws [18]; as a reply the authors defended their claim [19].

These current experimental efforts motivate theoretical studies of multicompo- nent superconducting systems that are considered in this thesis.

(11)

Chapter 2

General Considerations

2.1 The n-component free energy density F

s

We will investigate the n-component superconductor using the Ginzburg-Landau and London models.

Following the well-known procedure by Ginzburg and Landau [25], by introduc- ing a macroscopic effective wave function ψ as an order parameter that through

|ψ|2 corresponds to number density, the free energy density of a normal specimen in the absence of a magnetic field Fn, two temperature-dependent parameters α and β, charge q, mass m, magnetic vector potential ~A and magnetic field ~h one can express the free energy density of a 1-component superconductor as following:

Fs= Fn+ α|ψ|2

2|ψ|4+ 1 2m|(~

i

∇ − q ~~ A)ψ|2+ ~h2

0 (2.1)

An important remark is that ~h refers to the microscopic magnetic field. The magnetic flux density ~B is an average of ~h over a suitable volume.

Equation 2.1 is easily generalized to n components as following:

Fs= Fn+X

l

"

αll|2l

2|ψl|4+ 1 2ml

|(~ i

∇ − q~ lA)ψ~ l|2

# + ~h2

0

(2.2)

It is instructive to compare that equation to the one found at [11, equation 1]

and the more general one at [10, p. 167]. As an example one can mention the liquid metallic hydrogen case [15] where different components correspond to condensates of protonic and electronic cooper pairs, respectively. A third and fourth component can arise through cooper pairs of tritium and Bose-Einstein condensation of deu- terium nuclei [11]. In a neutron star example [16] [23] [24], the different components correspond to condensates of neutronic and protonic cooper pairs, respectively.

3

(12)

2.2 Deriving the Ginzburg-Landau equations

Recall following from your course in electrodynamics:

~h = ~∇ × ~A (2.3)

We will obtain the Ginzburg-Landau equations by varying the free energy func- tional with respect to the order parameter Ψ and the magnetic vector potential A.~

2.2.1 Variation with respect to ψ

l

We will first minimize

R3d3rFs with respect to a variation of ψl:

0 =



R3

d3r

"

αlψlδψll

2|ψl|2lδψl

+(−i~∇ − q~ lA)ψ~ l· (~i∇δψ~ l− qlAδψ~ l) 2ml

# (2.4)

Next, recall that given any scalar φ and any vector ~T , we have following vector identity:

∇(φ ~~ T ) = ~∇φ · ~T + φ ~∇ · ~T (2.5)

Utilizing equation 2.5 in 2.4 for φ := ~iδψl and ~T := (

~

−i∇−q~ lA)~

2ml ψl, we then obtain:

0 =



R3

d3rδψl

"

αlψl + βll|2ψl+(−i~∇ − q~ lA)ψ~ l· (−qlA)~ 2ml

# +



R3

d3r ~∇ ·

"

~ iδψl

(−i~∇ − q~ lA)~ 2ml

ψl

#



R3

d3r~ iδψl∇ ·~

"

(−i~∇ − q~ lA)~ 2ml

ψl

# (2.6)

Keep now in mind that equation 2.6 is valid as well if we change integration region from R3 to the superconducting specimen (denoted by ”super”) since the order parameter ψl is 0 outside the specimen. We want to do this change because δψl is also 0 outside the specimen and we want to utilize the condition that we can choose arbitrary δψlinside the specimen to find useful equations.

(13)

2.2. Deriving the Ginzburg-Landau equations 5 Furthermore, recall the generalized divergence theorem [1, p. 93]:



d3r ~∇(...) =



d ~S(...) (2.7)

We can now rewrite equation 2.6 as:

0 =



super

d3rδψl

"

αlψl + βll|2ψl+(−i~∇ − q~ lA)~ 2 2ml

ψl

# +



∂super

δψld ~S ·

"

~ i

(−i~∇ − q~ lA)~ 2ml

ψl

# (2.8)

Which yields a Ginzburg-Landau equation and a boundary condition for each component of the superconductor:

(~i∇ − q~ lA)~ 2 2ml

ψl+ αlψl+ βll|2ψl= 0 (2.9)

d ~S · (~ i

∇ − q~ lA)ψ~ l= 0 (2.10) Recall the expression for the canonical momentum in the case of a particle in a magnetic field [2, p. 308]:

~

p = m~v + q ~A (2.11)

Also recall following expression for the momentum operator:

~ˆ p = ~

i

∇~ (2.12)

We can use equations 2.11 and 2.12 to express 2.10 as following:

d ~S · mlv~l= 0 (2.13)

Which means that the normal component of ~vl through the surface of the su- perconducting specimen is 0.

2.2.2 Variation with respect to ~ A

Next we will utilize equation 2.3 to express ~h in terms of ~ A and then minimize

R3d3rFswith respect to a variation of ~A:

(14)

0 =X

l

"

R3

d3r

−qlψlδ ~A · (−i~∇ − q~ lA)~ 2ml

ψl+−qlψlδ ~A · (~i∇ − q~ lA)~ 2ml

ψl

#

+



R3

d3r

"

2( ~∇ × ~A) · ( ~∇ × δ ~A) 2µ0

#

(2.14)

Recall now that for any ~C and ~D, the following vector identity holds:

∇ · ( ~~ C × ~D) = ~D · ( ~∇ × ~C) − ~C · (∇ × ~D) (2.15) By choosing ~C := µ~h

0 = ∇× ~~µA

0 and ~D := δ ~A and using equation 2.15 in 2.14, we obtain:

0 =X

l

"

R3

d3r−ql

2ml

δ ~A ·

 ψl( ~

−i∇ − q~ lA)ψ~ l + ψl(~ i

∇ − q~ lA)ψ~ l

# +



R3

d3r

"

δ ~A · (∇ × ~h~ µ0

) − ~∇ ·(~h × δ ~A) µ0

#

(2.16)

Next, we again use the general divergence theorem shown at equation 2.7, which yields:

0 =



R3

d3rδ ~A ·

"

X

l

−ql

2ml

 ψl( ~

−i∇ − q~ lA)ψ~ l + ψl(~ i

∇ − q~ lA)ψ~ l

 +

∇ × ~h~ µ0

# +



∂R3

d ~S ·

"

−(~h × δ ~A) µ0

#

(2.17)

Assuming that ~h × δ ~A decays faster than r12 as r → ∞, the surface integral vanishes as r → ∞. The remaining integral yields a Ginzburg-Landau equation:

∇ × ~h~ µ0

=X

l

"

ql~ 2iml

l∇ψ~ l− ψl∇ψ~ l) − ql2 ml

l|2A~

#

(2.18)

One should also keep in mind that when the displacement current is negligible and by defining ~J := ~Jf ree+ ~Jbound , a Maxwell equation yields [3, p. 330]:

J =~

∇ × ~h~

µ0 (2.19)

(15)

2.4. Supercurrents in the London limit 7

2.3 London limit

As commonly taught in literature regarding superconductivity, the conventional parametrization of the order parameter ψ is ψ = |ψ|e.

In the London limit, one models |ψ| to be constant with respect to space co- ordinates throughout the superconducting region and zero in normal regions [13].

The London limit is reasonable when the magnetic penetration length (length scale over which the magnetic field is reduced to a small fraction of its maximum value) is much larger than the coherence lengths (length scales over which the order pa- rameters regain their corresponding bulk values) [27].

We can express the Ginzburg-Landau equations 2.9 and 2.18 as following in the London limit:

(~~∇θl− qlA)~ 2

2ml ψl+ αlψl+ βll|2ψl= 0 (2.20) J =~ X

l

"

ql~

mll|2∇θ~ l− ql2 mll|2A~

#

(2.21) Furthermore we can also simplify the expression for the free energy Fsat equa- tion 2.2 as following in the London limit:

Fs= Fn+X

l

"

αll|2l

2|ψl|4+ 1

2ml(~~∇θl− qlA)~ 2l|2

# + ~h2

0

(2.22)

The part of Fsthat is relevant in the context of vortex formation in the London limit can be described as following:

s:=X

l

"

1

2ml(~~∇θl− qlA)~ 2l|2

# + ~h2

0 (2.23)

2.4 Supercurrents in the London limit

A few words of caution: what one means by ”current” is a matter of convention.

Griffith [3, p. 212] refers to ~J as ”volume current density”, but we will refer to it as ”current” throughout this text.

Recall equation 2.21:

J =~ X

l

"

ql~

mll|2∇θ~ l− ql2 mll|2A~

#

(2.24)

Following the procedure for two-component superconductors done at [4], we define supercurrents ~jl as following:

(16)

~jl:= ql~ ml

l|2∇θ~ l− q2l ml

l|2A~ (2.25)

Given that definition it is easy to notice that following holds:

J =~ X

l

~jl (2.26)

I.e. the current ~J that appears in Maxwell’s equations is in this case the sum of all supercurrents.

The definition makes sense from another perspective as well. Using equations 2.11, 2.12 and ψl= |ψl|el we obtain:

~ vl= ~

ml

∇θ~ l− ql

ml

A~ (2.27)

In general a charged current is obtained by multiplying a velocity with a charge and a density. Choosing |ψl|2as the density, ~vlas the velocity and qlas the charge one obtains:

~jl= qlv~ll|2= ql~

mll|2∇θ~ l− q2l

mll|2A~ (2.28) Which is the same expression as in equation 2.25.

(17)

Chapter 3

Vortices

3.1 Definitions

[26, p. 202] essentially provides following definitions:

Vorticity is defined as the curl of the fluid velocity. A line drawn in R3 such that it is at each point parallel to the vorticity is called a ”vortex-line”. If we draw a closed curve in R3 and at each of its points draw a vortex line we have drawn a tube which is called a ”vortex-tube”. The fluid contained in a vortex-tube is called a ”vortex-filament” and is abbreviated as ”vortex”.

3.2 Vortex flux

Using the typical argument that ~J vanishes at large distances and using equation 2.24, we obtain following result for a contour sufficiently far away:

0 =

J · ~~ dl = X

l

"

ql~ ml

l|2

∇~~θl · ~dl

#

− X

l

"

q2l ml

l|2

A · ~~ dl

# (3.1)

Recall the definition of magnetic flux and Stokes’ theorem to obtain:

Φ :=



B · ~~ dS =



( ~∇ × ~A) · ~dS =

A · ~~ dl (3.2) Also recall following:

∇~~θl· ~dl = "

dx ∂

∂x + dy ∂

∂y + dz ∂

∂z

#

θl = ∆θl = 2πNl, Nl ∈ Z (3.3) Nl is the so called ”winding number” which arises as a consequence of the single-valuedness of the order parameter ψl.

9

(18)

Using equations 3.1, 3.2 and 3.3, we obtain a result which matches the one found at [11, equation 3]:

Φ = P

l ql~

mll|22πNl P

j qj2 mjj|2

(3.4)

3.3 Calculating the magnetic vector potential in the London limit

We will proceed similar to how it is done in [5].

From [3, p. 44] we have for cylindrical parametrization (s, ϕ, z) following equation:

∇ × ~~ A = (1 s

∂Az

∂ϕ −∂Aϕ

∂z ) ˆes+ (∂As

∂z −∂Az

∂s ) ˆeϕ+1

s(∂(sAϕ)

∂s −∂As

∂ϕ) ˆez (3.5) We will be looking at two cases in particular: one where one applies an external field in the ˆezdirection and one where one maintains a rotation in the ˆezdirection.

Therefore we can set ~h = hzz. One way to achieve it is by following magnetic vector potential choice:

A = A~ ϕ(s, ϕ) ˆeϕ (3.6)

Equations 2.3, 3.5 and 3.6 together yield:

~h = 1 s

∂(sAϕ)

∂s eˆz (3.7)

From [3, p. 44], we have:

∇ · ~~ A = 1 s

∂(sAs)

∂s +1 s

∂Aϕ

∂ϕ +∂Az

∂z (3.8)

Using equation 3.6 in 3.8, we then obtain:

∇ · ~~ A = 1 s

∂Aϕ

∂ϕ (3.9)

Choosing the gauge ~∇ · ~A = 0 then leads to that Aϕ is independent of ϕ:

A = A~ ϕ(s) ˆeϕ (3.10)

Through equation 3.7 we see that it means that ~h does not depend on ϕ. Using that together with 3.5 where we replace ~A with ~h, we obtain:

(19)

3.3. Calculating the magnetic vector potential in the London limit 11

J =~

∇ × ~h~ µ0

=−1 µ0

∂hz

∂s eˆϕ=−1 µ0

∂s(1 s

∂(sAϕ)

∂s ) ˆeϕ (3.11) One can use following ansatz as hinted by [5]:

∇~~θl=Nl

s eˆϕ (3.12)

We verify that it yields correct winding number:

∇~~θl· ~dl = Nl

s eˆϕ· sdϕ ˆeϕ= 2πNl (3.13) Using equations 2.24, 3.11 and 3.12 we obtain:

−1 µ0

∂s(1 s

∂(sAϕ)

∂s ) ˆeϕ=X

l

"

ql~ ml

l|2Nl s − q2l

ml

l|2Aϕ

# ˆ

eϕ (3.14)

Recall the modified Bessel differential equation:

x2d2y dx2 + xdy

dx− (x2+ n2)y = 0 (3.15) It is well-known that it has following general solution [7, equation 9.6.1]:

y = C1In(x) + C2Kn(x) (3.16) A relevant property is [7, equation 9.6.19] which shows that In(x) grows too quickly to be a valid solution for our purposes:

In(x) = 1 π

 π 0

ex cos θcos(nθ)dθ (3.17)

A difficulty lies in connecting the flux Φ to the constant C2. It is done by Annett [6, p. 63] for following equation:

s2d2hz

ds2 + sdhz

ds − (s2

λ2 + 02)hz= 0 (3.18) That yielded following:

hz(s) = Φ 2πλ2K0(s

λ) (3.19)

Analogously, we can put equation 3.14 in a more useful form:

(20)

s2d2Aϕ

ds2 + sdAϕ ds −

"

µ0X

l

 ql2 ml

l|2s2

 + 12

#

Aϕ+ µ0X

l

 ql~ ml

l|2Nls



= 0 (3.20)

That looks similar to the modified Bessel differential equation for n = 1. We will make following ansatz [5, equation 7] where D is a constant and following choice for λ:

Aϕ(s) = C2K1(s λ) +D

s (3.21)

λ := 1

q µ0P

l q2l mll|2

(3.22)

Inserting those in equation 3.20 yields:

s2d2(C2K1(sλ))

ds2 + sd(C2K1(λs))

ds −

"

s2 λ2 + 12

#

C2K1(s λ)+

s22D

s3 + s−D s2

"

s2 λ2 + 12

#D

s + µ0X

l

 ql~ ml

l|2Nls



= 0 (3.23)

By realizing that the first row of equation 3.23 satisfies the modified Bessel differential equation and rearranging the 2nd row we obtain:

−s2 λ2

D s + µ0

X

l

 ql~ ml

l|2Nls



= 0 (3.24)

Which yields:

D = λ2µ0X

l

 ql~ ml

l|2Nl



(3.25) Inserting the result for D in the equation 3.21 for Aϕyields:

Aϕ(s) = C2K1(s

λ) +λ2µ0 s

X

l

 ql~ ml

l|2Nl



(3.26)

3.4 Calculating the microscopic field in the London limit

The modified Bessel function of the second kind, Kn(x), has following two proper- ties [7, equation 9.6.28]:

(21)

3.5. Determining the constant C2 13

d

dx(xnKn(x)) = −xnKn−1(x) (3.27) d

dx(x−nKn(x)) = −x−nKn+1(x) (3.28) Recall equation 3.7 which is valid for our cases:

~h = 1 s

∂(sAϕ)

∂s eˆz (3.29)

Using equation 3.26 for Aϕtogether with 3.29 in 3.27 for n = 1, yields:

~h = C2

s

d(sK1(λs))

ds eˆz= −s λ

C2

s K0(s

λ) ˆez=−C2

λ K0(s

λ) ˆez (3.30)

3.5 Determining the constant C

2

We will follow a similar approach to the one by de Gennes [8, p. 58-59]. We will first look at how ~∇ × (~∇ × ~h) looks like when we insert our expression 3.30 for ~h and use equation 3.5 to express derivatives in cylindrical coordinates:

∇ × ~h = −~ ∂hz

∂s eˆϕ=C2

λ

∂K0(λs)

∂s eˆϕ (3.31)

Using equation 3.28 for the case n = 0 in 3.31 yields:

∇ × ~h = −~ C2

λ2K1(s

λ) ˆeϕ (3.32)

Using equation 3.5 again yields:

∇ × (~~ ∇ × ~h) = 1 s

∂(−sCλ22K1(λs))

∂s eˆz (3.33)

Using equations 3.27, 3.30 and 3.33 for the case n = 1 then yields:

∇ × (~~ ∇ × ~h) = 1 s s λ

C2

λ2K0(s

λ) ˆez= C2

λ3K0(s

λ) ˆez= −~h

λ2 (3.34)

We recognize that as the well-known London equation. However, in the London model for vortices, there is a small cylindrical core of radius ξ << λ that is a normal region rather than a superconducting one. The symmetry axis of the core is parallel to ˆez. From Griffiths [3, p. 274] we have:

~h||above− ~h||below = µ0( ~K × ˆn) (3.35) It is an equation at the boundary between the core and the rest of the specimen where ~K is the surface current and ˆn is the normal direction. Furthermore, Griffiths [3, p. 267] provides following equation for the bound surface current ~Kb:

(22)

K~b= ~M × ˆn (3.36) Assuming that the magnetization ~M just like ~h only has a z-component, we obtain:

K~b = Mzz× − ˆes= −Mzϕ (3.37) Inserting that in equation 3.35 yields:

~h||above− ~h||below= µ0( ~Kf ree× − ˆes) + µ0(−Mzϕ× − ˆes) (3.38) Assuming that there is no free surface current and that the magnetization is zero at the boundary to the core, we thus obtain that the z-component of ~h and thus ~h itself is continuous at the boundary.

Furthermore, since the order parameter ψlis 0 in the normal region, then ~J is ~0 according to equation 2.21. Next to satisfy that result for ~J through ~∇ × ~h = µ0J ,~ we will make the ansatz that ~h is constant in the normal region. Then due to the continuity of ~h and equation 3.30 following holds in the normal region of the vortex:

~h = −C2

λ K0(

λ) ˆez (3.39)

That expression can be simplified by using the low-argument limit of K0 [7, equation 9.6.8]:

~h = −C2 λ ln(λ

) ˆez (3.40)

Due to the discontinuity of ~∇ × ~h = µ0J at the core boundary, the London~ Equation 3.34 only holds in the superconducting region. To generalize the London equation to also hold in the normal region one can approximate the core with a point and use:

λ2∇ × (~~ ∇ × ~h) + ~h = Φδ(s)

2πseˆz (3.41)

Outside of the core we can identify that as the regular London Equation. To make sure that it is a proper expression in the context of the normal region we integrate over a surface within the superconducting region but that is at a distance from the core much larger than λ:



dS · λ~ 2∇ × (µ~ 0J ) +~



dS · ~h =~



dS · Φ~ δ(s)

2πseˆz (3.42) Using Stokes’ theorem and using the expression for flux we obtain:

dl · λ~ 2µ0J + Φ =~



sdϕdsΦδ(s)

2πs (3.43)

Just like in 3.1, the circulation of the current disappears at large distances, which then as expected yields:

(23)

3.6. Calculating the circulation of velocity in the London limit 15

0 + Φ = Φ (3.44)

Now that we have established that the suggested generalized London Equation is reasonable, we will look at 3.42 again but integrate this time in a circle just outside the core:

dl · λ~ 2µ0J +~



dS · ~h = Φ~ (3.45)

Since ξ << λ, the contribution from dS · ~h is negligible. Furthermore we can~ use 3.32 to express ~J explicitly. That yields:

ξdϕλ2−C2

λ2 K1

λ) = Φ (3.46)

Using that K1(λs) = λs for small arguments [7, equation 9.6.9], that yields:

C2= − Φ

2πλ (3.47)

Next using the equation 3.30 for ~h, equation 3.26 for Aϕ, equation 3.22 for λ, equation 3.4 for Φ, equation 3.32 for ~∇ × ~h = µ0J we obtain following useful~ expressions:

A =~

"

Φ 2πs− Φ

2πλK1(s λ)

# ˆ

eϕ (3.48)

~h = Φ 2πλ2K0(s

λ) ˆez (3.49)

µ0J = ~~ ∇ × ~h = Φ 2πλ3K1(s

λ) ˆeϕ (3.50)

3.6 Calculating the circulation of velocity in the London limit

Recall equation 2.27:

~ vl= ~

ml

∇θ~ l− ql ml

A~ (3.51)

Using equation 3.48 for ~A and 3.12 for ~∇~θl, we can calculate the circulation of the velocity within the superconducting region:

(24)

~

vl· ~dl = 1 ml

sdϕ

"

~Nl

s −qlΦ 2πs+ qlΦ

2πλK1(s λ)

#

= 2π~

ml

"

Nl− qlΦ 2π~

# + 1

ml

sdϕ

"

qlΦ 2πλK1(s

λ)

#

(3.52)

For a large contour, the 2nd term vanishes due to the exponential decay of K1(sλ) [7, equation 9.7.2]. For that case, using equation 3.4 for Φ, we obtain:

~

vl· ~dl = 2π~

ml

"

Nl−ql P

n qn

mnn|2Nn

P

j q2j mjj|2

#

=2π~

ml

P

n qn

mnn|2(Nlqn− Nnql) P

j q2j mjj|2

(3.53)

3.6.1 2-component example

To test the derived equation 3.53, we will compare to the result for the 2-component case at [9, equation 2]. Therefore, we will choose N1 = 0, N2 = 1, q1 = e and q2= −e and a large contour, which yields:

~

v1· ~dl = 2π~

m1

h e

m11|2(0 − 0) +−em

22|2(0 − e)i

e2

m11|2+me2

22|2 =2π~

m1

2|2 m2

1|2 m1 +m2|2

2

(3.54)

~v2· ~dl = 2π~

m2 h e

m11|2(e − 0) +m−e

22|2(−e + e)i

e2

m11|2+me2

22|2 = 2π~

m2

1|2 m1

1|2 m1 +m2|2

2

(3.55)

That result agrees with [9, equation 2]!

3.6.2 The special case of exponentially decaying supercurrents

Looking at equation 2.28 we see that the supercurrent ~jl is proportional to the velocity ~vl, hence such velocities being exponentially decaying within a supercon- ducting region means that corresponding supercurrents within the same region are exponentially decaying as well.

Looking at equation 3.52 and 3.53, we notice that if Nlqn − Nnql = 0 for all choices of n and l, then ~vl is exponentially decaying and hence ~jl is exponentially decaying as well. Evidently such exponential decay means that for a large contour the circulation of ~vl is zero.

The condition can be rephrased as: Nql

l =Nqn

n for all choices of n and l.

(25)

3.7. Free energy of a vortex in the London limit 17

3.7 Free energy of a vortex in the London limit

Using the equation 2.23 for ˜Fs, 3.48 for ~A and 3.12 for ~∇θl, we obtain:

s:=X

l

"

1 2ml ~

Nl

s − ql

 Φ 2πs− Φ

2πλK1(s λ)

!2

l|2

# + ~h2

0 = X

l

"

1 2ml



~Nl

s − ql

Φ 2πs

2

+ qlΦ 2πλK1(s

λ)

2 +

2



~ Nl

s − ql

Φ 2πs

 qlΦ 2πλK1(s

λ)



l|2

# + ~h2

0

(3.56)

Utilizing equation 3.4 for Φ then yields:

s=X

l

"

l|2 2ml

~2 s2

 Pn qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2

2

+ qlΦ 2πλK1(s

λ)

2 +

2~

s

 P

n qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

 qlΦ 2πλK1(s

λ)

# + ~h2

0

(3.57)

Next, we want to integrate ˜Fs over all space. However, since the order param- eters ψl is 0 outside of the superconducting region and ~h decays exponentially for large distances, it is enough that we integrate over the superconducting region and the vortex core, i.e. over the specimen:



specimen

d3r ˜Fs=



specimen

d3rX

l

"

l|2 2ml

~2 s2

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2

+ qlΦ 2πλK1(s

λ)

2

+2~

s

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

 qlΦ 2πλK1(s

λ)

# +



specimen

d3r~h20

(3.58)

To simplify the expression, we want to use the generalized London Equation 3.41. To do so we got to simplify our expression. Something to also keep in mind is that ~J = ~0 and ψl = 0 holds true inside the core. We will start by using equation 3.50 for K1(λs) and equation 3.22 for λ:

(26)



specimen

d3r ˜Fs=



specimen

d3rX

l

"

l|2 2ml

~2 s2

 P

n qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2

+



qlλ2∇ × ~h~

2 +2~

s

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

 qlΦ 2πλK1(s

λ)

# +



specimen

d3r~h20

=



specimen

d3rX

l

"

l|2~2 2mls2

 P

n qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2

2#

+



specimen

d3rX

l

"

l|2~ mls

 Pn qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2

 qlΦ 2πλK1(s

λ)

# +



specimen

d3r λ20

( ~∇ × ~h)2+



specimen

d3r~h20

(3.59)

For the 3rd integral in equation 3.59, we can use the vector identity described by equation 2.15 for the case ~C := ~h and ~D := ~∇ × ~h as well as Gauss’ divergence theorem:



specimen

d3r λ20

( ~∇ × ~h)2=



specimen

d3r λ20

∇ · (~h × (~~ ∇ × ~h)) +



specimen

d3r λ20

~h · (~∇ × (~∇ × ~h)) =



∂specimen

d ~S · λ2

0(~h × ( ~∇ × ~h)) +



specimen

d3r λ2

0~h · (~∇ × (~∇ × ~h)) (3.60) The surface integral in equation 3.60 disappears since both ~h and ~∇ × ~h decay exponentially at large distances. Combining the remaining integral in 3.60 with the 4th integral in equation 3.59 and using the generalized London Equation 3.41, we obtain:



specimen

d3r λ2

0( ~∇ × ~h)2+



specimen

d3r~h20 =



specimen

d3r 1 2µ0~h ·h

λ2∇ × (~~ ∇ × ~h) + ~hi

=



specimen

d3r 1

0~h · Φδ(s)

2πseˆz (3.61) Assuming cylindrical specimen with radius R and height ∆z as well as a vortex core radius ξ and then using equation 3.49 for ~h and the low-argument limit for K0 as shown by equation 3.40, then yields:

(27)

3.7. Free energy of a vortex in the London limit 19



specimen

d3r λ2

0( ~∇ × ~h)2+



specimen

d3r~h20 =



sdϕdsdz 1

0~h · Φδ(s)

2πseˆz= ∆zΦhz(ξ)

0 = ∆z Φ2 4πµ0λ2ln(λ

ξ) (3.62)

We recognize that result as being in the form analogous to the vortex energy for the 1-component case as shown at [10, equation 5.85]. Next we will turn our attention to the 2nd integral in equation 3.59, use equation 3.28 for the case n = 0 to express K1as a derivative:



specimen

d3rX

l

"

l|2~ mls

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

 qlΦ 2πλK1(s

λ)

#

=



sdϕdsdzX

l

"

l|2~ mls

 P

n qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2



−qlΦ 2π

d ds

K0(s λ)#

= 2π∆zX

l

"

l|2~ ml

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2



−qlΦ 2π

 K0(R

λ)−K0(ξ λ)#

(3.63)

We can simplify the result by again using the low-argument limit for K0as shown by equation 3.40 and utilizing that K0decays exponentially for large arguments [7, equation 9.7.2]:

2π∆zX

l

"

l|2~ ml

 P

n qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2



−qlΦ 2π(K0(R

λ) − K0(ξ λ)

#

=

∆zΦ ln(λ ξ)~X

l

"

l|2ql

ml

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

#

(3.64)

Finally, we will look at the first integral of equation 3.59:

(28)



specimen

d3rX

l

"

l|2~2 2mls2

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2#

=

2π∆z~2 2

 ds s

X

l

"

l|2 ml

 P

n qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2#

=

π∆z~2lnR ξ

X

l

"

l|2 ml

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2#

(3.65)

Combining the results from equations 3.65, 3.64 and 3.62 we finally obtain:



specimen

d3r ˜Fs= π∆z~2lnR ξ

X

l

"

l|2 ml

 Pn qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2

2# +

∆zΦ ln(λ ξ)~X

l

"

l|2ql

ml

 P

n qn

mnn|2(qnNl− qlNn) P

j qj2 mjj|2

#

+ ∆z Φ2 4πµ0λ2ln(λ

ξ) (3.66)

Something to keep in mind is that the right-hand side of equation 3.66 is positive due to the squares in the expression 2.23 for ˜Fs.

3.7.1 2-component example

Like in our previous examples, we will choose N1= 0, N2= 1, q1= e and q2= −e and a large radius R. Since the first term in equation 3.66 grows the fastest with R, we will investigate it first since if it is non-zero the remaining terms would be small compared for it:

(29)

3.7. Free energy of a vortex in the London limit 21

π∆z~2lnR ξ

X

l

"

l|2 ml

 Pn qn

mnn|2(qnNl− qlNn) P

j q2j mjj|2

2#

=

π∆z~2lnR ξ

"

1|2 m1

 P

n qn

mnn|2(0 − eNn) P

j q2j mjj|2

2

+|ψ2|2 m2

 P

n qn

mnn|2(qn+ eNn) P

j qj2 mjj|2

2#

= π∆z~2lnR ξ

"

1|2 m1

−e

m22|2(−e) P

j q2j mjj|2

2 +|ψ2|2

m2

 e

m11|2(e) +m−e

22|2(−e + e) P

j qj2 mjj|2

2#

= π∆z~2lnR ξ

1|2 m1

2|2 m2

" 2|2 m2

(P

j

j|2 mj )2 +

1|2 m1

(P

j

j|2 mj )2

#

=

π∆z~2lnR ξ

1|2 m1

2|2 m2

1 (m1|2

1 +m2|2

2 ) (3.67) That result is non-zero and using the argument above we do not need to calculate other contributions. This result is also the same as in [9]!

3.7.2 The special case of exponentially decaying supercurrents

Just like in chapter 3.6.2, we will look at the special case Nlqn− Nnql= 0 for all choices of l and n at large distances R. Looking at the equation 3.66 for the free energy of a vortex, we notice that the two first terms are zero for this case while the third remains:



specimen

d3r ˜Fs= ∆z Φ2 4πµ0λ2ln(λ

ξ) (3.68)

This scales as O(R0) rather than O(ln(R)) which would have been the case if the first term of equation 3.66 did not vanish.

(30)

References

Related documents

ABSTRACT The overall aim of this thesis was to explore the role of Magnetic resonance imaging MRI of the prostate as an adjunct to the prostate-specific antigen PSA-test in

The overall aim of this thesis was to explore the role of Magnetic resonance imaging (MRI) of the prostate as an adjunct to the prostate-specific antigen (PSA)-test in

If the austenitic stainless steels construction process is unknown, it contradicts the statement that wave power is sustainable as the structure material is an essential

For all contrast agents tested, the liver R2* values returned to baseline values at earlier time points when lower doses were administered, thereby indicating that degradation

Most phase transitions from a paramagnetic to a ferromagnetic state is of second order. In that case, the susceptibility diverges at the critical tempera- ture. The transition

In a theoretical context one can compute vortex interactions and conduct molecular dynamics/MC simulations using the resulting interaction potentials, or determine the ground state of

The recorded XRD patterns collected as a function of both temperature and applied magnetic field could, as well as the conventional diffraction pattern, Fig. 1, be refined in

The ab initio and first principles theo- ries were employed to investigate the vibrational effects on the isotropic hyperfine coupling constant (HFCC) known as the critical parameter